skip to main content
research-article
Open Access

From Non-punctuality to Non-adjacency: A Quest for Decidability of Timed Temporal Logics with Quantifiers

Published:23 June 2023Publication History

Skip Abstract Section

Abstract

Metric Temporal Logic (MTL) and Timed Propositional Temporal Logic (TPTL) are prominent real-time extensions of Linear Temporal Logic (LTL). In general, the satisfiability checking problem for these extensions is undecidable when both the future (Until, U) and the past (Since, S) modalities are used (denoted by MTL[U,S] and TPTL[U,S]). In a classical result, the satisfiability checking for Metric Interval Temporal Logic (MITL[U,S]), a non-punctual fragment of MTL[U,S], is shown to be decidable with EXPSPACE complete complexity. A straightforward adoption of non-punctuality does not recover decidability in the case of TPTL[U,S]. Hence, we propose a more refined notion called non-adjacency for TPTL[U,S] and focus on its 1-variable fragment, 1-TPTL[U,S]. We show that non-adjacent 1-TPTL[U,S] is strictly more expressive than MITL. As one of our main results, we show that the satisfiability checking problem for non-adjacent 1-TPTL[U,S] is decidable with EXPSPACE complete complexity. Our decidability proof relies on a novel technique of anchored interval word abstraction and its reduction to a non-adjacent version of the newly proposed logic called PnEMTL. We further propose an extension of MSO [<] (Monadic Second Order Logic of Orders) with Guarded Metric Quantifiers (GQMSO) and show that it characterizes the expressiveness of PnEMTL. That apart, we introduce the notion of non-adjacency in the context of GQMSO (NA-GQMSO), which is a syntactic generalization of logic Q2MLO due to Hirshfeld and Rabinovich and show the decidability of satisfiability checking for NA-GQMSO.

Skip 1INTRODUCTION Section

1 INTRODUCTION

Metric Temporal Logic (\(\text{MTL}\)) and Timed Propositional Temporal Logic (\(\text{TPTL}\)) are natural extensions of Linear Temporal Logic (\(\text{LTL}\)) for specifying real-time properties [5]. \(\text{MTL}\) extends the Until (\(\mathsf {U}\)) and Since (\(\mathsf {S}\)) modalities of \(\text{LTL}\) by associating a timing interval with these. \(a \mathsf {U}_I b\) describes behaviors modelled as timed words consisting of a sequence of a’s followed by a b, which occurs at a time within (relative) interval I. However, \(\text{TPTL}\) uses freeze quantification to store the current timestamp. A freeze quantifier (also called as Half Order Quantifiers [3]) with clock variable x has the form \(x.\varphi\). When it is evaluated at a point i on a timed word, the timestamp \(\tau _i\) at i is frozen or registered in x, and the formula \(\varphi\) is evaluated using this value for x. Variable x is used in \(\varphi\) in a constraint of the form \(T-x \in I\); this constraint, when evaluated at a point j, checks if \(\tau _j -\tau _i \in I\), where \(\tau _j\) is the timestamp at point j.1 For example, the formula \(\mathsf {F}x.(a\wedge \mathsf {F}(b \wedge T-x \in [1,2] \wedge \mathsf {F}(c \wedge T-x \in [1,2])))\) asserts that there is a point in the future where a holds and in its future within interval \([1,2]\), b and c occur, and b occurs before c. This property is not expressible in \(\text{MTL}[\mathsf {U},\mathsf {S}]\) [8, 39]. Moreover, every property in \(\text{MTL}[\mathsf {U},\mathsf {S}]\) can be expressed in 1 variable fragment of TPTL (1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\)). Thus, 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) is strictly more expressive than \(\text{MTL}[\mathsf {U},\mathsf {S}]\). Unfortunately, both the logics have an undecidable satisfiability checking problem, making automated analysis of these logics difficult (in full generality existence of a sound and complete algorithm is impossible for such problems). It is possible to restrict certain parameters of the behaviors and get terminating algorithms. But that would require prior information about some parameters of the behaviors, which may not be always accessible. Moreover, the complexity of the algorithm often depends on the value of these parameters. For example, if we restrict the models to be k-bounded variable, i.e., models where the number of events within any unit time interval is bounded by \(k,\)2 then the satisfiability checking becomes decidable for these logics [19] but the complexity of this problem depends on k. Moreover, this would require access to this bound k, which is not the case, in general. Exploring natural decidable variants of these logics has been an active area of research since their advent [4, 23, 25, 26, 42, 43, 46]. One line of work restricted the logic to contain future only modality \(\text{MTL}[\mathsf {U}]\) and 1-\(\text{TPTL}[\mathsf {U}]\). Both these logics have been shown to have decidable satisfiability over finite timed words, under a pointwise interpretation [22, 37].3 The complexity, however, is non-primitive recursive. Moreover, these problems become undecidable over infinite timed words. Obtaining an expressive fragment with elementary complexity has been a challenging problem. One of the most celebrated such logics is the Metric Interval Temporal Logic (\(\text{MITL}[\mathsf {U},\mathsf {S}]\)) [1], a subclass of \(\text{MTL}[\mathsf {U},\mathsf {S}]\) where the timing intervals is restricted to be non-punctual, i.e., non-singular (intervals of the form \(\langle x, y \rangle\) where \(x \lt y\)). The satisfiability checking for \(\text{MITL}\) formulae is decidable with EXPSPACE complete complexity [1]. While non-punctuality helps to recover the decidability of \(\text{MTL}[\mathsf {U},\mathsf {S}]\), it does not help in \(\text{TPTL}[\mathsf {U},\mathsf {S}]\). The freeze quantifiers of \(\text{TPTL}\) enable us to trivially express punctual timing constraints using only the non-punctual intervals: For instance, the 1-\(\text{TPTL}\) formula \(x.(a \mathsf {U}(a \wedge T-x \in [1, \infty) \wedge T-x \in [0, 1]))\) uses only non-punctual intervals but captures the \(\text{MTL}\) formula \(a \mathsf {U}_{[1,1]} b\). Thus, a more refined notion of non-punctuality is needed to recover the decidability of 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\).4

Contributions. With the above observations, to obtain a decidable class of 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) akin to \(\text{MITL}[\mathsf {U},\mathsf {S}]\), we revisit the notion of non-punctuality as it stands currently. As our first contribution, we propose non-adjacency, a refined version of non-punctuality. Two intervals \(I_1\) and \(I_2\) are non-adjacent if the supremum of \(I_1\) is not equal to the infimum of \(I_2\). Non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) is the subclass of 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\), where every interval used in clock constraints within the same freeze quantifier is non-adjacent to itself and to every other timing interval that appears within the same scope. (W.l.o.g., we consider formulae in negation normal form only.) The non-adjacency restriction disallows punctual timing intervals: Every punctual timing interval is adjacent to itself. It can be shown (Theorem 3.2) that non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\), while seemingly very restrictive, is strictly more expressive than \(\text{MITL}\) and it can also express the counting and the Pnueli modalities [25]. Thus, the logic is of considerable interest in practical real-time specification (see Example 3.1).

Our second contribution is to give a decision procedure for the satisfiability checking of non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\). We do this in two steps. (1) We introduce a logic \(\text{PnEMTL}\) that combines and generalizes the automata modalities of References [27, 43, 46] and the Pnueli modalities of References [25, 26, 42] and has not been studied before to the best of our knowledge. We show that a formula in non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) can be reduced to an equivalent formula of non-adjacent \(\text{PnEMTL}\) (Theorem 5.16). (2) We prove that the satisfiability of non-adjacent \(\text{PnEMTL}\) is decidable with EXPSPACE complete complexity (Theorem 8.1) by reducing it to an equisatisfiable \(\text{EMITL}_{0,\infty }\) formulae (subclass of \(\text{EMTL}\) where the timing constraints are restricted to be of the form \(\langle 0, u\rangle\) or \(\langle l, \infty \rangle\) where \(l,u \in \mathbb {N}\cup \lbrace 0\rbrace\)).

As our third and final contribution, we show that the logic PnEMTL is expressively equivalent to an extension of MSO[\(\lt\)] (Monadic Second Order Logic of Orders) with Guarded Metric Quantifiers (GQMSO). The latter is a versatile and expressive logic, allowing properties of real-time systems to be defined conveniently. The use of Guarded Metric Quantifiers appeared in the pioneering formulations of logics QMLO and Q2MLO (with non-punctual guards) by Hirshfeld and Rabinovich [25] and it was further explored by Hunter (with punctual guards) [28]. We have generalized these to an anchored block of guarded quantifiers with arbitrary depth. This provides the required power to obtain expressive completeness. We show this by providing effective reductions from \(\text{PnEMTL}\) to GQMSO and vice versa. Unfortunately, the full \(\text{PnEMTL}\), being a syntactic extension of \(\text{MTL}[\mathsf {U},\mathsf {S}]\), is clearly undecidable. As our final main result, we define the non-adjacency condition, suitably applied to the logic GQMSO. We observe that the effective reductions between GQMSO and PnEMTL preserve non-adjacency. From the previously established EXPSPACE-complete decidability of non-adjacent \(\text{PnEMTL}\), it follows that the satisfiability checking for non-adjacent GQMSO is decidable.

The article is organized as follows: Section 2 introduces the models and logics \(\text{LTL}\), \(\text{MTL},\) and \(\text{TPTL}\). Section 3 introduces MTL extended with Pnueli automata modalities (\(\text{PnEMTL}\)) and non-adjacent fragments of \(1-\text{TPTL}\) and \(\text{PnEMTL}\). Section 4 introduces a novel notion of anchored interval word abstractions that we use to abstract timed languages. Its theory is central in the reduction of any (non-adjacent) \(\text{TPTL}\) formula to an equivalent (non-adjacent) \(\text{PnEMTL}\) formula presented in Section 5. Section 6 introduces a new extension of MSO[\(\lt\)] with Guarded Metric Quantifiers (GQMSO) and its non-adjacent fragment (NA-GQMSO). As mentioned, this logic is a natural syntactic generalization of QMLO and Q2MLO of Reference [26] and QkMSO of Reference [33]. In Section 7, we show that \(\text{PnEMTL}\) (and non-adjacent \(\text{PnEMTL}\)) is equivalent to GQMSO (and non-adjacent GQMSO, respectively) by giving effective reductions in both directions. Finally, Section 8 shows that satisfiability checking for non-adjacent \(\text{PnEMTL}\) and \(1-\text{TPTL}\) is decidable with EXSPACE complete complexity. This, along with the reduction from non-adjacent GQMSO to non-adjacent \(\text{PnEMTL},\) implies that the satisfiability checking problem for non-adjacent GQMSO is decidable. Finally, in Section 10, we conclude our article and discuss its place in the existing literature. We finish by proposing a fundamental open question in timed logics.

Discussion and related work. Much of the related work has already been discussed. \(\text{MITL}\) with counting and Pnueli modalities has been shown to have EXPSPACE-complete satisfiability [41, 42]. Here, we tackle more expressive logics, namely, non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) and non-adjacent \(\text{PnEMTL}\). We show that the EXPSPACE-complete satisfiability checking is retained in spite of the additional expressive power. These decidability results are proved by equisatisfiable reductions to logic \(\text{EMITL}_{0,\infty }\) of Ho [27]. As argued by Ho, it is quite practicable to extend the existing model checking tools like UPPAAL to logic \(\text{EMITL}_{0,\infty }\) and hence to our logics, too.

Addition of regular expression-based modalities to untimed logics like LTL has been found to be quite useful for practical specification; even the IEEE standard temporal logic PSL has this feature [15, 18, 29]. With a similar motivation, there has been considerable recent work on adding regular expression/automata-based modalities to \(\text{MTL}\) and \(\text{MITL}\). Raskin as well as Wilke added automata modalities to \(\text{MITL}\) as well as an Event-Clock logic ECL [43, 46] and showed its satisfiability checking problem to be decidable. Krishna et al. showed that \(\text{MTL}[\mathsf {U},\mathsf {S}_{NP}]\) (where \(\mathsf {U}\) can use punctual intervals but \(\mathsf {S}\) is restricted to non-punctual intervals), when extended with counting as well as regular expression modalities preserves decidability of satisfaction [31, 32, 33, 35]. Recently, Ferrère in Reference [17] proposed a very neat extension of MITL, called Metric Interval Dynamic Logic (MIDL), where the timing constraints appear within regular expressions as opposed to modalities (\(\text{LTL}[\mathsf {U}]\) extended with a fragment of timed regular (MIRE) expression modality). He showed that satisfiability checking for MIDL is decidable with EXPSPACE complete complexity. Moreover, Ho has investigated a PSPACE-complete fragment \(\text{EMITL}_{0,\infty }\) and showed that this fragment is surprisingly as expressive as the full logic \(\text{EMITL}\) [27]. Our non-adjacent \(\text{PnEMTL}\) is a novel extension of MITL with modalities that combine the features of EMITL [27, 43, 46] and the Pnueli modalities [25, 26, 42]. In terms of expressiveness, MIDL is also known to be strictly more expressive than EMITL. However, the relation between non-adjacent PnEMTL and MIDL remains open.

In terms of expressive completeness, Hirshfeld and Rabinovich [26] showed that MITL is expressively complete to an extension of FO[\(\lt\)] with metric quantifiers (quantifiers guarded with non-punctual timing constraints) where the subformulae within the scope of this metric quantifier is restricted to have only one free variable. Moreover, its extension, Q2MLO (where the subformulae within the scope of the metric quantifier can have no more than 2 free variables), is expressively equivalent to MITL extended with Pnueli Modalities. Hunter [28] showed that when one allows punctual guards in Q2MLO, one gets the complete first-order logic with distance operator FO[\(\lt ,+1\)] in continuous semantics. Inspired by these logics, Reference [33] proposed its extensions with restricted form of second-order quantification giving Q2MSO or QkMSO and allows punctuality. But these logics were restricted to reason about future time properties only, to preserve the decidability. Our proposed logic GQMSO is a syntactic generalization of all these logics.

Skip 2PRELIMINARIES Section

2 PRELIMINARIES

Let \(\Sigma\) be a finite set of propositions, and let \(\Gamma = 2^{\Sigma } \setminus \emptyset .\)5 A word over \(\Sigma\) is a finite sequence \(\sigma = \sigma _1 \sigma _2 \ldots \sigma _n\), where \(\sigma _i \in \Gamma\). A timed word \(\rho\) over \(\Sigma\) is a non-empty finite sequence \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\) of pairs \((\sigma _i, \tau _i) \in (\Gamma \times \mathbb {R}_{\ge 0})\); where \(\tau _1=0\) and \(\tau _i \le \tau _j\) for all \(1 \le i \le j \le n\) and n is the length of \(\rho\) (also denoted by \(|\rho |\)). The \(\tau _i\) are called timestamps. For a timed or untimed word \(\rho\), let \(dom(\rho) = \lbrace i | 1 \le i \le |\rho |\rbrace\), and \(\sigma [i]\) denotes the symbol at position \(i \in dom(\rho)\). The set of timed words over \(\Sigma\) is denoted \(T\Sigma ^*\). Given a (timed) word \(\rho\) and \(i \in dom(\rho)\), a pointed (timed) word is the pair \(\rho , i\). Let \(\mathcal {I}_\mathsf {int}\) (\(\mathcal {I}_\mathsf {nat}\)) be the set of open, half-open, or closed time intervals, such that the end points of these intervals are in \(\mathbb {Z}\cup \lbrace -\infty ,\infty \rbrace\) (\(\mathbb {N} \cup \lbrace 0,\infty \rbrace\), respectively).

2.1 Linear Temporal Logic

Formulae of \(\text{LTL}\) are built over a finite set of propositions \(\Sigma\) using Boolean connectives and temporal modalities (\(\mathsf {U}\) and \(\mathsf {S}\)) as follows: \(\varphi ::=a~|\top ~|~\varphi \wedge \varphi ~|~\lnot \varphi ~| ~\varphi \mathsf {U}\varphi ~|~ \varphi \mathsf {S}\varphi\), where \(a \in \Sigma\). The satisfaction of an \(\text{LTL}\) formula is evaluated over pointed words. For a word \(\sigma = \sigma _1 \sigma _2 \ldots \sigma _n \in \Sigma ^*\) and a point \(i \in dom(\sigma)\), the satisfaction of an \(\text{LTL}\) formula \(\varphi\) at point i in \(\sigma\) is defined, recursively, as follows:

(i)

\(\sigma , i \models a\) iff \(a \in \sigma _{i}\),

(ii)

\(\sigma ,i \models \top\) iff \(i \in dom (\rho)\),

(iii)

\(\sigma ,i \models \lnot \varphi\) iff \(\sigma ,i \nvDash \varphi\)

(iv)

\(\rho ,i \models \varphi _{1} \wedge \varphi _{2}\) iff \(\sigma ,i \models \varphi _{1}\) and \(\sigma ,i\ \models \ \varphi _{2}\),

(v)

\(\rho ,i \models \varphi _{1} \vee \varphi _{2}\) iff \(\sigma ,i \models \varphi _{1}\) or \(\sigma ,i\ \models \ \varphi _{2}\),

(vi)

\(\sigma ,i\ \models \ \varphi _{1} \mathsf {U}\varphi _{2}\) iff \(\exists j \gt i\), \(\sigma ,j\ \models \ \varphi _{2}\), and \(\sigma ,k\ \models \ \varphi _{1}\) \(\forall\) \(i\lt k \lt j\),

(vii)

\(\sigma ,i\ \models \ \varphi _{1} \mathsf {S}\varphi _{2}\) iff \(\exists j \lt i\), \(\sigma ,j\ \models \ \varphi _{2}\), and \(\sigma ,k\ \models \ \varphi _{1}\) \(\forall\) \(j\lt k \lt i\).

Derived operators can be defined as follows: \(\mathsf {F}\varphi = \top \mathsf {U}\varphi\), and \(\mathcal {G}\varphi = \lnot \mathsf {F}\lnot \varphi\). Symmetrically, \(\mathsf {P}\varphi = \top \mathsf {S}\varphi\), and \(\mathcal {H}\varphi = \lnot \mathsf {P}\lnot \varphi\). An LTL formula is said to be in negation normal form if it is constructed out of basic and derived operators above, but where negation appears only in front of propositional letters. It is well known that every LTL formula can be converted to an equivalent formula that is in negation normal form.

2.2 Metric Temporal Logic (MTL)

\(\text{MTL}\) is a real-time extension of \(\text{LTL}\) where the modalities (\(\mathsf {U}\) and \(\mathsf {S}\)) are guarded with intervals. Formulae of \(\text{MTL}\) are built from \(\Sigma\) using Boolean connectives and time-constrained versions \(\mathsf {U}_I\) and \(\mathsf {S}_I\) of the standard \(\mathsf {U},\mathsf {S}\) modalities, where \(I \in \mathcal {I}_\mathsf {nat}\). Intervals of the form \([x,x]\) are called punctual; a non-punctual interval is one that is not punctual. Formulae in \(\text{MTL}\) are defined as follows: \(\varphi ::=a ~|\top ~|\varphi \wedge \varphi ~|~\lnot \varphi ~| ~\varphi \mathsf {U}_I \varphi ~|~ \varphi \mathsf {S}_I \varphi\), where \(a \in \Sigma\) and \(I \in \mathcal {I}_\mathsf {nat}\). For a timed word \(\rho = (\sigma _1, \tau _1) (\sigma _2, \tau _2) \ldots (\sigma _n, \tau _n) \in T\Sigma ^*\), a position \(i \in dom(\rho)\), an \(\text{MTL}\) formula \(\varphi\), the satisfaction of \(\varphi\) at a position i of \(\rho\), denoted \(\rho , i \models \varphi\), is defined below. We discuss the time-constrained modalities.

\(\rho ,i\ \models \ \varphi _{1} \mathsf {U}_{I} \varphi _{2}\) iff \(\exists j \gt i\), \(\rho ,j\ \models \ \varphi _{2}, \tau _{j} - \tau _{i} \in I\), and \(\rho ,k\ \models \ \varphi _{1}\) \(\forall\) \(i\lt k \lt j\).

\(\rho ,i\ \models \ \varphi _{1} \mathsf {S}_{I} \varphi _{2}\) iff \(\exists j \lt i\), \(\rho ,j\ \models \ \varphi _{2}, \tau _{i} - \tau _{j} \in I\), and \(\rho ,k\ \models \ \varphi _{1}\) \(\forall\) \(j\lt k \lt i\).

The language of an \(\text{MTL}\) formula \(\varphi\) is defined as \(L(\varphi) = \lbrace \rho | \rho , 1 \models \varphi \rbrace\). Using the above, we obtain some derived formulae: the constrained eventual operator \(\mathsf {F}_I \varphi \equiv \top \mathsf {U}_I \varphi\) and its dual is \(\mathcal {G}_I \varphi \equiv \lnot \mathsf {F}_I \lnot \varphi\). Similarly \(\mathcal {H}_I \varphi \equiv \top \mathsf {S}_I \varphi\). The next operator is defined as \(\oplus _I \varphi \equiv \bot \mathsf {U}_I \varphi\). The non-strict versions of \(\mathsf {F}, \mathcal {G}\) are, respectively, defined as \(\mathsf {F}^w \varphi \equiv \varphi ~\vee ~ \mathsf {F}\varphi\) and \(\mathcal {G}^w\varphi \equiv \varphi \wedge \mathcal {G}\varphi\) include the present point. Symmetric non-strict versions for past operators are also allowed. The subclass of \(\text{MTL}\) obtained by restricting the intervals I in the until and since modalities to non-punctual intervals is known as Metric Interval Temporal logic and denoted by \(\text{MITL}[\mathsf {U}, \mathsf {S}]\). We say that a formula \(\varphi\) is satisfiable iff \(L(\varphi)\ne \emptyset\).

Theorem 2.1.

Satisfiability checking for \(\text{MTL}[\mathsf {U},\mathsf {S}]\) is undecidable [4]. Satisfiability Checking for \(\text{MITL}[\mathsf {U}, \mathsf {S}]\) is EXPSPACE-complete [1, 2, 3].

2.3 Timed Propositional Temporal Logic (TPTL)

The logic \(\text{TPTL}\) also extends \(\text{LTL}\) using freeze quantifiers. Like \(\text{MTL}\), \(\text{TPTL}\) is also evaluated on timed words. Formulae of \(\text{TPTL}\) are built from \(\Sigma\) using Boolean connectives, modalities \(\mathsf {U}\) and \(\mathsf {S}\) of \(\text{LTL}\). In addition, \(\text{TPTL}\) uses a finite set of real valued clock variables \(X = \lbrace x_1,\ldots ,x_n\rbrace\). Let \(\nu : X \rightarrow \mathbb {R}_{\ge 0}\) represent a valuation assigning a non-negative real value to each clock variable. The formulae of \(\text{TPTL}\) are defined as follows: Without loss of generality, we work with \(\text{TPTL}\) in the negation normal form. \(\varphi ::=a~|~\lnot a ~|\top ~|~\bot ~|~ x.\varphi ~|~ T-x \in I ~|~x-T \in I~|~\varphi \wedge \varphi ~|~ \varphi \vee \varphi ~| ~\varphi \mathsf {U}\varphi ~|~ \varphi \mathsf {S}\varphi ~|~ \mathcal {G}\varphi ~|~ \mathcal {H}\varphi\), where \(x \in X\), \(a \in \Sigma\), \(I \in \mathcal {I}_\mathsf {int}\). Here, T denotes the timestamp of the point where the formula is being evaluated. \(x. \varphi\) is the freeze quantification construct that remembers the timestamp of the current point in variable x and evaluates \(\varphi\).

For a timed word \(\rho =(\sigma _1,\tau _1)\ldots (\sigma _n,\tau _n)\), \(i \in dom(\rho)\) and a \(\text{TPTL}\) formula \(\varphi\), we define the satisfiability relation, \(\rho , i, \nu \models \varphi\) with valuation \(\nu\) of all the clock variables. We omit the semantics of Boolean, \(\mathsf {U}\) and \(\mathsf {S}\) operators as they are similar to those of \(\text{LTL}\).

\(\rho , i, \nu \models a\) iff \(a \in \sigma _{i}\), and \(\rho ,i,\nu \models x.\varphi\) iff \(\rho ,i,\nu [x \leftarrow \tau _i] \models \varphi\),

\(\rho ,i,\nu \models T-x\ \in I\) iff \(\tau _i - \nu (x) \in I\), and \(\rho ,i,\nu \models x-T \in I\) iff \(\nu (x) - \tau _i \in I\),

\(\rho ,i,\nu \ \models \ \mathcal {G}\varphi\) iff \(\forall j \gt i\), \(\rho ,j,\nu \ \models \ \varphi\), and

\(\rho ,i,\nu \ \models \ \mathcal {H}\varphi\) iff \(\forall j \lt i\), \(\rho ,j,\nu \ \models \ \varphi\).

Let \(\overline{0}=(0,0,\ldots ,0)\) represent the initial valuation of all clock variables. For a timed word \(\rho\) and \(i \in dom(\rho)\), we say that \(\rho , i\) satisfies \(\varphi\) denoted \(\rho , i \models \varphi\) iff \(\rho ,i,\overline{0}\models \varphi\). The language of \(\varphi\), \(L(\varphi) = \lbrace \rho | \rho , 1 \models \varphi \rbrace\). The Pointed Language of \(\varphi\) is defined as \(L_{pt}(\varphi) = \lbrace \rho ,i ~\mid ~ \rho , i \models \varphi \rbrace\). A \(\text{TPTL}\) formula is said to be closed if every variable is quantified using freeze quantifier before it appears in a clock constraint. For example, \(x.y.(a \mathsf {U}(b \wedge x \in (1,2) \wedge y\in (2,3)))\) is a closed formula while \(x.(a \wedge y\in (2,3))\mathsf {U}y.(b \wedge x \in (1,2))\) is not closed (or open), as y is used in a clock constraint before it is frozen. Note that for a closed formula, the satisfaction of the model is independent of the clock valuation. In other words, if \(\psi\) is a closed formula, then either for every valuation \(\nu\), \(\rho , i, \nu \models \psi\); or for every valuation \(\nu\), \(\rho ,i, \nu \nVdash \psi\). Hence, for a closed formula \(\psi\), we drop the valuation tuple while evaluating for satisfaction as \(\rho ,i,\nu \models \psi\) for any valuation \(\nu\), iff \(\rho ,i,\overline{0} \models \psi\).

Logic 1-TPTL: The subclass of \(\text{TPTL}\) that uses only 1 clock variable (i.e., \(|X| = 1\)) is known as 1-\(\text{TPTL}\). As an example, the closed formula \(\varphi =x.(a \mathsf {U}(b \mathsf {U}(c \wedge T-x \in [1,2])))\) is satisfied by the timed word \(\rho =(a,0)(a,0.2)(b,1.1)(b,1.9)(c,1.91)(c,2.1)\), since \(\rho , 1 \models \varphi\). The word \(\rho ^{\prime }=(a,0)(a,0.3)(b,1.4) (c,2.1)(c,2.5)\) does not satisfy \(\varphi\). However, \(\rho ^{\prime },2 \models \varphi\): If we start from the second position of \(\rho ^{\prime }\), then we assign \(\nu (x)=0.3\), and when we reach the position 4 of \(\rho ^{\prime }\) with \(\tau _4=2.1,\) we obtain \(T-x=2.1-0.3 \in [1,2]\). Note that an \(\text{MTL}[\mathsf {U},\mathsf {S}]\) formula can straightforwardly be translated to an equivalent 1-\(\text{TPTL}[\mathsf {U}, \mathsf {S}]\) (closed) formula. Hence, by Theorem 2.1, we get that the satisfiability checking for 1-\(\text{TPTL}[\mathsf {U}, \mathsf {S}]\) is undecidable.

Notation: Let x denote the unique freeze variable we use in 1-\(\text{TPTL}\). All constraints in 1-TPTL have the form \(T-x \in I\). (Note that \(x-T \in I\) is equivalent to \(T-x \in -I\).) Thus, for 1-TPTL, let \(\widehat{I}\) abbreviate \(T-x \in I\).

2.4 Expressive Completeness and Strong Equivalence

Given any specification (formula or automaton) X and Y, X is equivalent to Y when for any pointed timed word \(\rho ,i\), \(\rho ,i \models X \iff \rho ,i \models Y\). We say that a formalism \(\mathcal {X}\) (logic or machine) is expressively complete to \(\mathcal {Y}\), denoted by \(\mathcal {Y} \subseteq \mathcal {X}\), if and only if, for any formulae/automata \(Y \in \mathcal {Y}\) there exists an equivalent \(X \in \mathcal {X}\). \(\mathcal {X}\) is said to be expressively equivalent to \(\mathcal {Y}\), denoted by \(\mathcal {X} \cong \mathcal {Y}\), when \(\mathcal {X} \subseteq \mathcal {Y}\) and \(\mathcal {Y} \subseteq \mathcal {X}\).

Skip 3INTRODUCING NON-ADJACENT 1-TPTL AND PNUELI EMTL Section

3 INTRODUCING NON-ADJACENT 1-TPTL AND PNUELI EMTL

In this section, we define non-adjacent 1-\(\text{TPTL}\). We also give a generalization of \(\text{MTL}\) called \(\text{PnEMTL}\) and define its non-adjacent fragment. Let x denote the unique freeze variable we use in 1-\(\text{TPTL}\).

3.1 Non-adjacent 1-TPTL

Non-Adjacent 1-\(\text{TPTL}\) (NA-1-TPTL) is defined as a subclass of 1-\(\text{TPTL}\) where adjacent intervals within the scope of any freeze quantifier is disallowed. Two intervals \(I_1, I_2 \in \mathcal {I}_\mathsf {int}\) are non-adjacent iff \(\sup (I_1) {\ne } \inf (I_2)~~ \vee ~~ \sup (I_1) = 0\). A set \(\mathcal {I}_{\nu }\) of intervals is non-adjacent iff any two intervals in \(\mathcal {I}_{\nu }\) are non-adjacent. It does not contain punctual intervals other than \([0,0],\) as every punctual interval is adjacent to itself. For example, the set \(\lbrace [1,2), (2,3], [5,6)\rbrace\) is not a non-adjacent set, while \(\lbrace [0,0], [0,1), (3,4], [5,6)\rbrace\) is. Let \(\mathcal {I}_{na}\) denote a set of non-adjacent intervals with end points in \(\mathbb {Z} \cup \lbrace -\infty ,\infty \rbrace\). Consider the following example of a formula in non-adjacent 1-\(\text{TPTL}\):

Example 3.1

(Non-adjacent 1-TPTL)

An indoor cycling exercise regime may be specified as follows: One must slow-pedal (prop. sp) for at least 60 seconds but until the odometer reads 1 km (prop. od1). From then onwards one must fast-pedal (prop fp) to a time point in the interval [600, 900] from the start of the exercise such that pulse rate is sufficiently high (prop ph) for the last 60 seconds of the exercise. This can be given by the following formula: \(\begin{equation*} x. sp ~~\mathsf {U}~~ \left[ \begin{array}[t]{l} ~\widehat{[60,\infty)} ~~\wedge od1 ~~\wedge \\ (~~fp ~~\mathsf {U}~~ (\widehat{[600, 900]} \wedge x. H(\widehat{[-60,0]} \Rightarrow ph))~~) \end{array} \right] . \end{equation*}\)

It can be shown that this formula cannot be expressed in logic \(\text{MITL}\).

The freeze depth of a \(\text{TPTL}\) formula \(\varphi\), \(\mathsf {fd}(\varphi)\) is defined inductively. For a propositional formula prop, \(\mathsf {fd}(prop)=0\). Also, \(\mathsf {fd}(x.\varphi)=\mathsf {fd}(\varphi)+1\), and \(\mathsf {fd}(\varphi _1 \mathsf {U}\varphi _2)=\mathsf {fd}(\varphi _1 \mathsf {S}\varphi _2) =\mathsf {fd}(\varphi _1 \wedge \varphi _2)=\mathsf {fd}(\varphi _1 \vee \varphi _2)=\mathsf {Max(\mathsf {fd}(\varphi _1),\mathsf {fd}(\varphi _2))}, \mathsf {fd}(\mathcal {G}(\varphi))=\mathsf {fd}(\mathcal {H}(\varphi))=\mathsf {fd}(\varphi)\).

Theorem 3.2.

Non-adjacent 1-\(\text{TPTL}[\mathsf {U},\mathsf {S}]\) is more expressive than \(\text{MITL}[\mathsf {U}, \mathsf {S}]\). It can also express the Counting and the Pnueli modalities of References [25, 26].

Proof.

The straightforward translation of MITL into TPTL in fact gives rise to non-adjacent 1-TPTL formula, e.g., MITL formula \(a \mathsf {U}_{[2,3]} (b \mathsf {U}_{[3,4]} c)\) translates to \(x.(a \mathsf {U}(\widehat{[2,3]} \wedge x.(b \mathsf {U}(\widehat{[3,4]} \wedge c)))\). It has been previously shown that \(\mathsf {F}[x.(a \wedge \mathsf {F}(b \wedge \widehat{(1,2)} \wedge \mathsf {F}(c \wedge \widehat{(1,2)})))]\), which is in fact a formula of non-adjacent 1-TPTL, is inexpressible in \(\text{MTL}[\mathsf {U},\mathsf {S}]\) (see Reference [39]). The Pnueli modality \(\mathsf {Pn}_I(\phi _1, \ldots , \phi _k)\) expresses that there exist positions \(i_1 \le \cdots \le i_k\) within (relative) interval I where each \(i_j\) satisfies \(\phi _j\). This is equivalent to the non-adjacent 1-TPTL formula \(x. (\mathsf {F}(\hat{I} \wedge \phi _1 \wedge \mathsf {F}(\hat{I} \wedge \phi _2 \wedge \mathsf {F}(\ldots))))\). Similarly, the (simpler) counting modality can also be expressed.□

3.2 Pnueli Automata Modalities

There have been several attempts to extend logic \(\text{MTL}\) with regular expression/automaton modalities [17, 27, 32, 46]. One of the most general amongst these is Automata Modalities, proposed by Wilke [46]. MITL (or MTL) extended with these automata modalities was called \(\text{EMITL}\) (or EMTL, respectively). We further generalize these automata modalities to give automata modalities of arbitrary arity. We call these modalities as Pnueli Automata Modalities. The extension is in the same spirit as the extension of future and past modalities to Pnueli future and Pnueli Past modalities in Reference [26]. We call MTL extended with these Pnueli Automata Modalities as \(\text{PnEMTL}\). We now first introduce EMTL before introducing \(\text{PnEMTL}\) for the sake of readability. For any finite automaton A, let \(L(A)\) denote the language of A.

3.2.1 MTL Extended with Automata Modalities, EMTL.

Given a finite alphabet \(\Sigma\), formulae of \(\text{EMTL}\) have the following syntax:

\(\varphi ::=a~|~\varphi \wedge \varphi ~|~\lnot \varphi ~|~ \mathcal {F}_{I} (\mathsf {A})(S)~|~ \mathcal {P}_{I} (\mathsf {A})(S)\) where \(a \in \Sigma\), \(I \in \mathcal {I}_\mathsf {nat}\) and \(\mathsf {A}\) is an automata over \(2^S\) where S is a set of formulae from \(\text{EMTL}\). \(\mathcal {F}_{I}\) and \(\mathcal {P}_{I}\) are future and past Automata Modalities, respectively.

Let \(\rho = (\sigma _1, \tau _1),\ldots (\sigma _n, \tau _n) \in T\Sigma ^*\), \(x,y \in dom(\rho)\), \(x\le y\) and \(S = \lbrace \varphi _1,\ldots , \varphi _n\rbrace\) be a given set of \(\text{EMTL}\) subformulae. Let \(S_i\) be the exact subset of formulae from S evaluating to true at \(\rho , i\), and let \(\mathsf {Seg^+}({\rho },{x},{y},S)\) and \(\mathsf {Seg^{-}}({\rho },{y},{x},S)\) be the untimed words \(S_x S_{x+1} \ldots S_y\) and \(S_y S_{y-1} \ldots S_x\), respectively. Then, the satisfaction relation for \(\rho ,i_0\) satisfying a \(\text{EMTL}\) formula \(\varphi\) is defined recursively as follows:

\(\rho ,i_0 \models \mathcal {F}_{I}(\mathsf {A})(S)\) iff \({\exists } {i_0 \le i_1\le n}\) s.t. \({[(\tau _{i_1} - \tau _{i_0} \in I_1)} \wedge \mathsf {Seg^+}(\rho , i_{0}, i_1,S) \in L({\mathsf {A}})]\),

\(\rho ,i_0 \models \mathcal {P}_{I} (\mathsf {A})(S)\) iff \({\exists } i_0 \ge i_1 \ge 1\) s.t. \([(\tau _{i_0} - \tau _{i_1} \in I_1) \wedge \mathsf {Seg^{-}}(\rho , i_{0}, i_1,S) \in L({\mathsf {A}})]\).

Language of any \(\text{EMTL}\) formula \(\varphi\), \(L(\varphi) = \lbrace \rho ~\mid ~ \rho ,1 \models \varphi \rbrace\). The Pointed Language of \(\varphi\) is defined as \(L_{pt}(\varphi) = \lbrace \rho ,i ~\mid \rho , i \models \varphi \rbrace\). Logic \(\text{EMITL}\) is a sublogic of \(\text{EMTL}\) where only non-punctual intervals are allowed along with the modalities \(\mathcal {F}\) a and \(\mathcal {P}\). Similarly, \(\text{EMITL}_{0,\infty }\) is defined as a sublogic of \(\text{EMITL}\) where the timing intervals associated with both the modalities is restricted to be either of the form \(\langle 0, u \rangle\) or of the form \(\langle l, \infty)\) where l and u are any non-negative integers.

Theorem 3.3.

Satisfiability Checking for \(\text{EMITL}\) is decidable [46] with EXPSPACE complete [17, 27]. Moreover, satisfiability checking for \(\text{EMITL}_{0,\infty }\) is PSPACE complete [27].

3.2.2 MTL Extended with Pnueli Automata Modalities, PnEMTL.

\(\text{PnEMTL}\) is defined similarly as \(\text{EMTL}\). Given a finite alphabet \(\Sigma\), formulae of \(\text{PnEMTL}\) have the following syntax: \(\varphi ::=a~|\varphi \wedge \varphi ~|~\lnot \varphi ~| \mathcal {F}^k_{I_1,\ldots ,I_k} (\mathsf {A}_1,\ldots , \mathsf {A}_{k+1})(S)~|~ \mathcal {P}^k_{I_1,\ldots ,I_k} (\mathsf {A}_1,\ldots ,\mathsf {A}_{k+1})(S)\) where \(a \in \Sigma\), \(I_1, I_2, \ldots I_k \in \mathcal {I}_\mathsf {nat}\) and \(\mathsf {A}_1, \ldots \mathsf {A}_{k+1}\) are automata over \(2^S\) where S is a set of formulae from \(\text{PnEMTL}\). \(\mathcal {F}^k\) and \(\mathcal {P}^k\) are the new modalities called future and past Pnueli Automata Modalities, respectively, where k is the arity of these modalities.

Let \(\rho = (\sigma _1, \tau _1),\ldots (\sigma _n, \tau _n) \in T\Sigma ^*\), \(x,y \in dom(\rho)\), \(x\le y\) and \(S = \lbrace \varphi _1,\ldots , \varphi _n\rbrace\) be a given set of \(\text{PnEMTL}\) formulae. Let \(\mathsf {Seg^+}({\rho },{x},{y},S)\) and \(\mathsf {Seg^{-}}({\rho },{y},{x},S)\) be as defined previously. Then, the satisfaction relation for \(\rho ,i_0\) satisfying a \(\text{PnEMTL}\) formula \(\varphi\) is defined recursively as follows:

\(\rho ,i_0 \models \mathcal {F}^k_{I_1,\ldots ,I_k}(\mathsf {A}_1,\ldots ,\mathsf {A}_{k+1})(S)\) iff \({\exists } {i_0 \le i_1\le i_2 \ldots \le i_k \le n}\) s.t. \(\bigwedge \nolimits _{w=1}^{k}{[(\tau _{i_w} - \tau _{i_0} \in I_w)} \wedge \mathsf {Seg^+}(\rho , i_{w-1}, i_w,S) \in L({\mathsf {A}_w})] \wedge \mathsf {Seg^+}(\rho , i_{k}, n,S) \in L({\mathsf {A}_{k+1}})\),

\(\rho ,i_0 \models \mathcal {P}^k_{I_1,I_2,\ldots ,I_k} (\mathsf {A}_1,\ldots ,\mathsf {A}_k, \mathsf {A}_{k+1})(S)\) iff \({\exists } i_0 \ge i_1 \ge i_2 \ldots \ge i_k \ge 1\) s.t. \(\bigwedge \nolimits _{w=1}^{k}[(\tau _{i_0} - \tau _{i_w} \in I_w) \wedge \mathsf {Seg^{-}}(\rho , i_{w-1}, i_w,S) \in L({\mathsf {A}_{w}})] \wedge \mathsf {Seg^{-}}(\rho , i_{k}, 1,S) \in L({\mathsf {A}_{k+1}})\).

Refer to Figure 1 for semantics of \(\mathcal {F}^{k}\).

Fig. 1.

Fig. 1. Figure showing semantics of \(\mathcal {F}^k_{I_1,\ldots ,I_k}(\mathsf {A}_1, \mathsf {A}_2, \ldots , \mathsf {A}_{k},\mathsf {A}_{k+1})(S)\) .

Language of any \(\text{PnEMTL}\) formula \(\varphi\), as \(L(\varphi) = \lbrace \rho ~\mid ~ \rho ,1 \models \varphi \rbrace\). The Pointed Language of \(\varphi\) is defined as \(L_{pt}(\varphi) = \lbrace \rho ,i ~\mid ~ \rho , i \models \varphi \rbrace\). Given a \(\text{PnEMTL}\) formula \(\varphi\), its arity is the maximum number of intervals appearing in any \(\mathcal {F}, \mathcal {P}\) modality of \(\varphi\). For example, the arity of \(\varphi =\mathcal {F}^2_{I_1,I_2}(\mathsf {A}_1,\mathsf {A}_2, \mathsf {A}_3)(S_1) \wedge \mathcal {P}^1_{I_1} (\mathsf {A}_1,\mathsf {A}_2)(S_2)\) for some sets of formulae \(S_1, S_2\) is 2. For the sake of brevity, \(\mathcal {F}^k_{I_1,\ldots ,I_k}(\mathsf {A}_1,\ldots ,\mathsf {A}_{k})(S)\) denotes \(\mathcal {F}^k_{I_1,\ldots ,I_k}(\mathsf {A}_1,\ldots ,\mathsf {A}_{k}, \mathsf {A}_{k+1})(S)\) where automata \(\mathsf {A}_{k+1}\) accepts all the strings over S. We define non-adjacent \(\text{PnEMTL}\) (NA-\(\text{PnEMTL}\)) as a subclass where every modality \(\mathcal {F}^{k}_{\mathsf {I_1,\ldots , I_k}}\) and \(\mathcal {P}^{k}_{\mathsf {I_1,\ldots , I_k}}\) is such that \(\lbrace \mathsf {I_1, \ldots , I_k}\rbrace\) is a non-adjacent set of intervals.

Note that \(\text{EMITL}\) of Reference [46] (and variants of it studied in References [27, 32, 33]) are special cases of the non-adjacent \(\text{PnEMTL}\) modality where the arity is restricted to 1 and the second automata in the argument accepts all the strings. Hence, automaton modality of Reference [46] is of the form \(\mathsf {\mathcal {F}_I}(A)(S)\) and \(\mathcal {P}_I(A)(S)\). Following is an example of a specification that could be naturally written as non-adjacent \(\text{PnEMTL}\) formula.

Example 3.4

(Non-adjacent PnEMTL)

A sugar-level test involves the following: A patient visits the lab and is given a sugar measurement test (prop sm) to get fasting sugar level. After this, she is given glucose (prop gl) and this must be within 5 min of coming to the lab. After this, the patient rests between 120 and 150 minutes and she is administered sugar measurement again to check the sugar clearance level. Following this, the result (prop rez) is given out between 23 to 25 hours (1,380, 1,500 min) of coming to the lab. We assume that these propositions are mutually exclusive and prop idle denotes negation of all of them. This protocol is specified by the following non-adjacent \(\text{PnEMTL}\) formula. For convenience, we specify the automata by their regular expressions. We follow the convention where the tail automaton \(A_{k+1}\) can be omitted in \(\mathcal {F}^k\). \(\begin{equation*} \mathcal {F}^2_{[0,5],~ [1,380,1,500]} \left[ \begin{array}{l} sm \cdot (idle^*) \cdot (gl \wedge \mathcal {F}^1_{[120,150]}(~gl \cdot (idle^*) \cdot sm~), \\ gl \cdot ((\lnot rez)^*) \cdot rez \end{array} \right] \end{equation*}\) For readability, the two regular expressions of the top \(\mathcal {F}^2\) are given in two separate lines. It states that the first regular expression must end at time within \([0,5]\) of starting and the second regular expression must end at a time within [1,380, 1,500] of starting. Note the nested use of \(\mathcal {F}\) to anchor the duration between glucose and the second sugar measurement.

3.3 Size of Formulae

The size of a temporal logic formula can be measured as usual, using the parse tree of the formula, or using the parse DAG (Directed Acyclic Graph) of the formula, where a syntactically unique subformula occurs only once. The latter representation is more succinct and is used widely starting from the classical LTL formula-to-automaton construction [14, 45]. For our results also, we will use the notion of DAG-size of a formula.

The (DAG) size of a formula \(\varphi\) denoted by \(|\varphi |\) is a measure of how many bits are required to store it in the DAG representation. The size of a \(\text{TPTL}\) formula is defined as the sum of the number of \(\mathsf {U}\), \(\mathsf {S}\) and Boolean operators and freeze quantifiers in it. For \(\text{PnEMTL}\) formulae, \(|op|\) is defined as the number of Boolean operators and variables used in it. \(|(\mathcal {F}^{k}_{I_1,\ldots ,I_k}(\mathsf {A}_1, \ldots , \mathsf {A}_{k+1})(S)|=\sum \nolimits _{\varphi \in S}(|\varphi |)+ |\mathsf {A}_1| + \cdots + |\mathsf {A}_{k+1}|+ 2k\times log(\mathsf {cmax})\) where \(|\mathsf {A}|\) denotes the size of the automaton \(\mathsf {A}\) given by the sum of number of its states and transitions and \(\mathsf {cmax}\) denotes the maximum allowable value of the constant used in the intervals \(I_1, \ldots , I_k.\)6

Skip 4ANCHORED INTERVAL WORD ABSTRACTION Section

4 ANCHORED INTERVAL WORD ABSTRACTION

All the logics considered here have the feature that a sub-formula asserts timing constraints on various positions relative to an anchor position; e.g., the position of freezing the clock in TPTL. Such constraints can be symbolically represented as an interval word with a unique anchor position and all other positions carry a set of time intervals constraining the timestamp of the position relative to the timestamp of the anchor. See interval word \(\kappa\) in Figure 2. We now define these interval words formally. Let \(I_{\nu }\subseteq \mathcal {I}_\mathsf {int}\). An \(I_{\nu }\)-interval word over \(\Sigma\) is a word \(\kappa\) of the form \(\sigma _1 \sigma _2 \dots \sigma _n \in (2^{\Sigma \cup \lbrace \mathsf {anch}\rbrace \cup I_\nu })^*\) such that:

Fig. 2.

Fig. 2. Point within the triangle has more than one interval. The encircled points are intermediate points and carry redundant information. The required timing constraint is encoded by first and last time-restricted points of all the intervals (within boxes).

(1)

There is a unique \(i \in dom(\kappa)\) such that \(\mathsf {anch}\in \sigma _i\). Such a position is called the anchor of \(\kappa\) and denoted by \(\mathsf {anch}(\kappa)\).

(2)

At all the points in \(\kappa\), at least one of the propositions from \(\Sigma\) holds. That is, for all \(i \in dom(\kappa)\), \(\sigma _i \cap \Sigma\) is a non-empty set.

Let J be any interval in \(I_\nu\). We say that a point \(i \in dom(\kappa)\) is a J-time-restricted point if and only if, \(J \in \sigma _i\). i is called time-restricted point if and only if either i is J-time restricted for some interval J in \(I_\nu\) or \(\mathsf {anch}\in a_i\).

From \(I_\nu\)-interval word to Timed Words: Given a \(I_\nu\)-interval word \(\kappa =\sigma _1 \dots \sigma _n\) over \(\Sigma\) and a timed word \(\rho =(\sigma ^{\prime }_1, \tau _1)\dots (\sigma ^{\prime }_m, \tau _m)\), the pointed timed word \(\rho , i\) is consistent with \(\kappa\) iff \(dom(\rho)=dom(\kappa)\) (i.e., \(m=n\)), \(i=\mathsf {anch}(\kappa)\), for all \(j\in dom(\kappa)\), \(\sigma ^{\prime }_j=\sigma _j\cap \Sigma\) and, \(I \in \sigma _j \cap I_\nu\) implies \(\tau _j - \tau _i \in I\). Thus, \(\kappa\) and \(\rho ,i\) agree on propositions from \(\Sigma\) at all positions, and the timestamp of any position j in \(\rho\) satisfies every interval constraint in \(\sigma _j\) relative to \(\tau _i\), the timestamp of anchor position. \(\mathsf {Time(\kappa)}\) denotes the set of all the pointed timed words consistent with a given interval word \(\kappa\), and \(\mathsf {Time}(\Omega)=\bigcup \nolimits _{\kappa \in \Omega } (\mathsf {Time(\kappa)})\) for a set of interval words \(\Omega\). Note that the “consistency relation” is a many-to-many relation.

Example 4.1.

Let \(\kappa =\lbrace a, b, (-1,0)\rbrace \lbrace b, (-1,0)\rbrace \lbrace a, \mathsf {anch}\rbrace \lbrace b, [2, 3]\rbrace\) be an interval word over the set of intervals \(\lbrace (-1,0), [2,3]\rbrace\). Consider timed words \(\rho\) and \(\rho ^{\prime }\) s.t. \(\rho = {(\lbrace a,b\rbrace , 0)(\lbrace b\rbrace , .5), (\lbrace a\rbrace , .95) (\lbrace b\rbrace , 3)}\), \(\rho ^{\prime } = {(\lbrace a,b\rbrace ,0)(\lbrace b\rbrace ,0.8)(\lbrace a\rbrace ,0.9)(\lbrace b\rbrace ,2.9)}\). Then, \(\rho , 3\) as well as \(\rho ^{\prime }, 3\) are consistent with \(\kappa\) while \(\rho , 2\) is not. Likewise, for the timed word \(\rho ^{\prime \prime } = (\lbrace a,b\rbrace , 0), (\lbrace b\rbrace , 0.5), (\lbrace a\rbrace , 1.1) (\lbrace b\rbrace , 3)\), \(\rho ^{\prime \prime }, 3\) is not consistent with \(\kappa\) as \(\tau _1 - \tau _3 \notin (-1,0)\), as also \(\tau _4 - \tau _3 \notin [2,3]\).

Let \(I_\nu , I_\nu ^{\prime } \subseteq \mathcal {I}_\mathsf {int}\). Let \(\kappa =\sigma _1\dots \sigma _n\) and \(\kappa ^{\prime }=\sigma ^{\prime }_1 \dots \sigma ^{\prime }_m\) be \(I_\nu\) and \(I_\nu ^{\prime }\)-interval words, respectively. \(\kappa\) is similar to \(\kappa ^{\prime }\), denoted by \(\kappa \sim \kappa ^{\prime }\) if and only if, (i) \(dom(\kappa)=dom(\kappa ^{\prime })\), (ii) for all \(i \in dom(\kappa)\), \(a_i \cap \Sigma =b_i\cap \Sigma\), and (iii) \(\mathsf {anch}(\kappa)=\mathsf {anch}(\kappa ^{\prime })\). Additionally, \(\kappa\) is congruent to \(\kappa ^{\prime }\), denoted by \(\kappa \cong \kappa ^{\prime }\), iff \(\mathsf {Time}(\kappa)=\mathsf {Time}(\kappa ^{\prime })\). That is, \(\kappa\) and \(\kappa ^{\prime }\) abstract the same set of pointed timed words.

Collapsed Interval Words: The set of interval constraints at a position can be collapsed into a single interval by taking the intersection of all the intervals at that position giving a Collapsed Interval Word. Given an \(I_{\nu }\)-interval word \(\kappa =\sigma _1 \dots \sigma _n\), let \(I_j = \sigma _j \cap I_{\nu }\). Let \(\kappa ^{\prime }=\mathsf {Col}(\kappa)\) be the word obtained by replacing \(I_j\) with \(\bigcap _{I \in I_j} I\) in \(\sigma _j\), for all \(j \in dom(\kappa)\). Note that \(\kappa ^{\prime }\) is an interval word over \(\mathsf {CL}(I_\nu)=\lbrace I | I=\bigcap I^{\prime }, I^{\prime } \subseteq I_\nu \rbrace\). Note that, if for any j, the set \(I_j\) contains two disjoint intervals (like \([1,2]\) and \([3,4]\)), then \(\mathsf {Col}(\kappa)\) is undefined. It is clear that \(\mathsf {Time}(\kappa)=\mathsf {Time}(\kappa ^{\prime })\). An interval word \(\kappa\) is called collapsed iff \(\kappa =\mathsf {Col}(\kappa)\). Normalization of Interval Words: An interval I may repeat many times in a collapsed interval word \(\kappa\). Some of these occurrences are redundant, and we can keep only the first and last occurrence of the interval without changing the set of pointed timed words consistent with it hence giving a normalized form of \(\kappa\). See Figure 2. For a collapsed interval word \(\kappa\) and any \(I \in I_\nu\), let \(\mathsf {first}(\kappa , I)\) and \(\mathsf {last}(\kappa ,I)\) denote the positions of first and last occurrence of I in \(\kappa\). If \(\kappa\) does not contain any occurrence of I, then both \(\mathsf {first}(\kappa , I)=\mathsf {last}(\kappa ,I)=\bot\). We define, \(\mathsf {Boundary}(\kappa)=\lbrace i | i \in dom(\kappa) \wedge \exists I \in I_\nu\) s.t. \((i=\mathsf {first}(\kappa ,I) {\vee }i=\mathsf {last}(\kappa ,I){\vee }i=\mathsf {anch}(\kappa)) \rbrace\).

The normalized interval word corresponding to \(\kappa\), denoted \(\mathsf {Norm}(\kappa)\), is defined as \(\kappa _{nor}=\sigma ^{\prime }_1 \dots \sigma ^{\prime }_m,\) such that (i) \(\kappa _{nor} \sim \mathsf {Col}(\kappa)\), (ii) for all \(I \in \mathsf {CL}(I_\nu)\), \(\mathsf {first}(\kappa , I)=\mathsf {first}(\kappa _{nor}, I)\), \(\mathsf {last}(\kappa , I)=\mathsf {last}(\kappa _{nor}, I)\), and for all points \(j \in dom(\kappa _{nor})\) with \(\mathsf {first}(\kappa , I) \lt j \lt \mathsf {last}(\kappa ,I)\), j is not a I-time-constrained point. See Figure 2. Hence, a normalized word is a collapsed word where for any \(J\in \mathsf {CL}(I_\nu)\) there are at most two J-time-restricted points. This is the key property that will be used to reduce 1-\(\text{TPTL}\) to a (bounded arity) \(\text{PnEMTL}\) formulae. In what follows, for any interval work \(\kappa = \sigma _1 \ldots \sigma _n\), for any point \(j \in dom(\kappa)\), \(\kappa [j] = \sigma _j\). Similarly, for any timed word \(\rho = (\sigma ^{\prime }_1, \tau _1) \ldots (\sigma ^{\prime }_m, \tau _m)\), for any \(j \in dom(\rho)\), \(\rho [j] = (\sigma _j,\tau _j)\), \(\rho [j](1) = \sigma _j\) and \(\rho [j](2) = \tau _j\).

The proof follows from the fact that \(\kappa \cong \mathsf {Col}(\kappa)\) and, since \(\mathsf {Col}(\kappa) \sim \mathsf {Norm}(\kappa)\), the set of timed words consistent with any of them will have identical untimed behavior. For the timed part, the key observation is as follows: For some interval \(I \in I_{\nu }\), let \(i^{\prime }=\mathsf {first}(\kappa ,I), j^{\prime }=\mathsf {last}(\kappa , I)\). Then, for any \(\rho ,i\) in \(\mathsf {Time}(\kappa)\), points \(i^{\prime }\) and \(j^{\prime }\) are within the interval I from point i. Hence, any point \(i^{\prime } \le i^{\prime \prime } \le j^{\prime }\) is also within interval I from i. Thus, the interval I need not be explicitly mentioned at intermediate points. Formally, the following two lemmas Lemma 4.2 and Lemma 4.3 imply Lemma 4.4. Lemma 4.2 shows \(\kappa \cong \mathsf {Col}(\kappa)\). Lemma 4.3 implies that \(\mathsf {Col}(\kappa) \cong \mathsf {Norm}(\kappa)\).

Lemma 4.2.

Let \(\kappa\) be an \(I_\nu\)-interval word. Then, \(\kappa \cong \mathsf {Col}(\kappa)\).

Proof.

A pointed word \(\rho ,i\) is consistent with \(\kappa\) iff

(i)

\(dom(\rho)=dom(\kappa)\),

(ii)

\(i=\mathsf {anch}(\kappa)\),

(iii)

for all \(j\in dom(\kappa)\), \(\rho [j](1)=\kappa [j]\cap \Sigma\) and

(iv)

for all \(j \ne i\), \(I \in a_j \cap I_\nu\) implies \(\rho [j](2) - \rho [i](2)\in I\).

(v)

\(\kappa \sim \mathsf {Col}(\kappa)\), by definition of \(\mathsf {Col}\).

Hence, given (v), (i) iff (a) (ii) iff (b) (iii) iff (c) where: (a) \(dom(\rho)=dom(\kappa)=dom(\mathsf {Col}(\kappa))\), (b) \(i=\mathsf {anch}(\kappa) = \mathsf {anch}(\mathsf {Col}(\kappa))\), (c) for all \(j\in dom(\kappa)\), \(\rho [j](1)=\kappa [j]\cap \Sigma = \mathsf {Col}(\kappa)[j] \cap \Sigma\). (iv) is equivalent to \(\rho [j](2) - \rho [i](2) \in \bigcap (\kappa [j] \cap I_\nu)\), but \(\bigcap (\kappa [j] \cap I_\nu) = \mathsf {Col}(\kappa)[j]\). Hence, (iv) iff (d) \(\rho [j](2) - \rho [i](2) \in \mathsf {Col}(\kappa)[j]\). Hence, (i), (ii), (iii), and (iv) iff (a), (b), (c), and (d). Hence, \(\rho ,i\) is consistent with \(\kappa\) iff it is consistent with \(\mathsf {Col}(\kappa)\).□

Lemma 4.3.

Let \(\kappa\) and \(\kappa ^{\prime }\) be \(I_\nu\)-interval words such that \(\kappa \sim \kappa ^{\prime }\). If for all \(I \in I_\nu\), \(\mathsf {first}(\kappa , I) = \mathsf {first}(\kappa ^{\prime }, I)\) and \(\mathsf {last}(\kappa , I) = \mathsf {last}(\kappa ^{\prime }, I)\), then \(\kappa \cong \kappa ^{\prime }\).

Proof.

The proof idea is the following:

As \(\kappa \sim \kappa ^{\prime }\), the set of timed words consistent with any of them will have identical untimed behavior.

For the timed part, the key observation is as follows: For some interval \(I \in I_{\nu }\), let \(i^{\prime }=\mathsf {first}(\kappa ,I), j^{\prime }=\mathsf {last}(\kappa , I)\). Then, for any \(\rho ,i\) in \(\mathsf {Time}(\kappa)\), points \(i^{\prime }\) and \(j^{\prime }\) are within the interval I from point i. Hence, any point \(i^{\prime } \le i^{\prime \prime } \le j^{\prime }\) is also within interval I from i. Thus, the intermediate I-time-restricted points (I-time-restricted points other than the first and the last) do not offer any extra information regarding the timing behavior. In other words, the restriction from the first and last I restricted points will imply the restrictions offered by intermediate I restricted points. Hence, their presence or absence makes no difference.

Both Lemmas 4.2 and 4.3 imply the following lemma, which will be used in the reduction of 1-TPTL to PnEMTL:

Lemma 4.4.

\(\kappa \cong \mathsf {Norm}(\kappa)\). Note, \(\mathsf {Norm}(\kappa)\) has at most \(2{\times }|I_\nu |^2+1\) restricted points.□

Let \(\rho = (a_1, \tau _1),\ldots (a_n,\tau _n)\) be any timed word. \(\rho , i\) is consistent with \(\kappa\) iff

(1)

(i)

\(dom(\rho) = dom(\kappa)\),

(ii)

\(i = \mathsf {anch}(\kappa)\),

(iii)

for all \(j \in dom(\rho)\), \(\kappa [j]\cap \Sigma = a_j\) and

(iv)

for all \(j\ne i \in dom(\rho)\), \(\tau _j -\tau _i \in \bigcap (I_\nu \cap \kappa [j])\).

Similarly, \(\rho , i\) is consistent with \(\kappa ^{\prime }\) if and only if

(2)

(a)

\(dom(\rho) = dom(\kappa ^{\prime })\),

(b)

\(i = \mathsf {anch}(\kappa ^{\prime })\),

(c)

for all \(j \in dom(\rho)\), if \(\kappa ^{\prime }[j] \cap \Sigma = a_j\) and

(d)

for all \(j\ne i \in dom(\rho)\), \(\tau _j -\tau _i \in \bigcap (I_\nu \cap \kappa ^{\prime }[j])\).

Note that, as \(\kappa \sim \kappa ^{\prime }\), we have, \(dom(\kappa) = dom(\kappa ^{\prime })\), \(\mathsf {anch}(\kappa) = \mathsf {anch}(\kappa ^{\prime })\), for all \(j \in dom (\kappa)\), \(\kappa [j]\cap \Sigma = \kappa ^{\prime }[j]\cap \Sigma\). Thus, 2(a) \(\equiv\) 1(i), 2(b) \(\equiv\) 1(ii), and 2(c) \(\equiv\) 1(iii).

Suppose there exists a \(\rho , i\) consistent with \(\kappa\) but there exists \(j^{\prime } \ne i \in dom(\rho)\), \(\tau _j^{\prime } -\tau _i \notin I^{\prime }\) for some \(I^{\prime } \in \kappa ^{\prime }[j^{\prime }]\). By definition, \(\mathsf {first}(\kappa ^{\prime },I^{\prime }) \le j^{\prime } \le \mathsf {last}(\kappa ^{\prime },I^{\prime })\). But \(\mathsf {first}(\kappa ^{\prime },I^{\prime }) = \mathsf {first}(\kappa ,I^{\prime })\), \(\mathsf {last}(\kappa ^{\prime },I^{\prime }) = \mathsf {last}(\kappa ,I^{\prime })\). Hence, \(\mathsf {first}(\kappa ,I^{\prime }) \le j^{\prime } \le \mathsf {last}(\kappa , I^{\prime })\). As the timestamps of the timed word increases monotonically, \(x \le y \le z\) implies that \(\tau _x \le \tau _y \le \tau _z\), which implies that \(\tau _x - \tau _i \le \tau _y - \tau _i \le \tau _z -\tau _i\). Hence, \(\tau _{\mathsf {first}(\kappa ,I^{\prime })} - \tau _i \le \tau _{j^{\prime }} - \tau _i \le \tau _{\mathsf {last}(\kappa ,I^{\prime })}- \tau _i\). But \(\tau _{\mathsf {first}(\kappa ,I^{\prime })} - \tau _i \in I^{\prime }\) and \(\tau _{\mathsf {last}(\kappa ,I^{\prime })}- \tau _i \in I^{\prime }\), because \(\rho\) is consistent with \(\kappa\). This implies that \(\tau _{j^{\prime }} - \tau _i \in I^{\prime }\) (as \(I^{\prime }\) is a convex set), which is a contradiction. Hence, if \(\rho ,i\) is consistent with \(\kappa\), then it is consistent with \(\kappa ^{\prime }\), too. By symmetry, if \(\rho , i\) is consistent with \(\kappa ^{\prime }\), then it is also consistent with \(\kappa\). Hence, \(\kappa \cong \kappa ^{\prime }\).

We give a road map to the proofs of results in Figure 3. In summary, \(\begin{equation*} \begin{array}{ll} \text{1-TPTL} ~~\lt ~~ \text{GQMSO} ~~\equiv ~~ \text{PnEMTL}, \ & (1) \\ \text{NA-1-TPTL} ~~\lt ~~ \text{NA-GQMSO} ~~\equiv ~~ \text{NA-}\text{PnEMTL}. & (2) \end{array} \end{equation*}\) The logics in (1) all have undecidable satisfiability, whereas logics in (2) all have decidable satisfiability. Specifically, NA-1-TPTL and NA-PnEMTLhave EXPSPACE-complete satisfiability checking.

Fig. 3.

Fig. 3. The above figure is a road map of all the main and intermediate results. An arrow from result A to result B indicates that the proof of result B uses the result A. For definitions of GQMSO and NA-GQMSO, please refer Section 6.

Skip 51-TPTL TO PNEMTL Section

5 1-TPTL TO PNEMTL

In this section, we reduce a 1-\(\text{TPTL}\) formula into an equivalent \(\text{PnEMTL}\) formula. First, we consider 1-\(\text{TPTL}\) formula in negation normal form with a single outermost freeze quantifier (call these simple \(\text{TPTL}\) formulae) and give the reduction. More complex formulae can be handled by applying the same reduction (shown below) recursively. For any set of formulae S, let \(\bigvee S\) denote \(\bigvee \nolimits _{s\in S}s\). This notation will be extensively used from this point onwards in all the succeeding sections, too.

A \(\text{TPTL}\) formula is said to be simple if it is in the negation normal form and of the form \(x.\varphi\) where, \(\varphi\) is a 1-\(\text{TPTL}\) formula with no freeze quantifiers. Let \(\psi =x.\varphi\) be a simple \(\text{TPTL}\) formula. Let \(I_\nu = \lbrace I ~|~ T-x \in I\) or appears in \(\varphi \rbrace \cup \lbrace -I ~|~ x-T \in I\) or appears in \(\varphi \rbrace\) and let \(\mathsf {CL}(I_\nu)=I_\nu\). We construct a \(\text{PnEMTL}\) formula \(\phi\), such that \(\rho ,i \models \psi {\iff }\rho , i \models \phi\). We break this construction into the following steps:

(1)

We construct an \(\text{LTL}\) formula \(\alpha\) s.t. \(L(\alpha)\) contains only \(I_\nu\)-interval words and \(\rho ,i \models \psi\) iff \(\rho , i \in \mathsf {Time}(L(\alpha))\). Let \(A_\alpha\) be the NFA s.t. \(L(A_{\alpha })=L(\alpha)\) (constructed using Reference [20]). We then construct NFA, A, over \(I_\nu =\mathsf {CL}(I_\nu)\) interval words from \(A_\alpha\) such that \(L(A) = \mathsf {Col}(L(A_\alpha))\). Note that \(|I_\nu | \le |I_\nu |^2\). Hence, \(\rho ,i \models \psi\) iff \(\rho , i \in \mathsf {Time}(L(A))\).

(2)

Let W be the set of all \(I_\nu\)-interval words. We can partition W into finitely many types, each type capturing a certain relative ordering between first and last occurrences of intervals from \(I_\nu\) as well as \(\mathsf {anch}\). Let \(\mathcal {T}(I_\nu)\) be the finite set of all such types. For each type \(\mathsf {seq}\in \mathcal {T}(I_\nu)\), we construct an NFA, \(A_{\mathsf {seq}}\), such that \(L(A_\mathsf {seq})=\mathsf {Norm}(L(A) \cap W_\mathsf {seq})\), where \(W_\mathsf {seq}\) is the set of all the \(I_\nu\)-interval words of type \(\mathsf {seq}\). Hence, \(A_\mathsf {seq}\) accepts only normalized interval words of type \(\mathsf {seq}\).

(3)

For every type \(\mathsf {seq}\), using the \(A_{\mathsf {seq}}\) above, we construct a \(\text{PnEMTL}\) formula \(\phi _\mathsf {seq}\) such that, \(\rho , i \models \phi _\mathsf {seq}\) if and only if \(\rho , i \in \mathsf {Time}(L(A_\mathsf {seq}))\). The desired \(\phi = \bigvee \nolimits _{\mathsf {seq}\in \mathcal {T}(I_\nu)} ~\phi _\mathsf {seq}\). Hence, \(L_{pt}(\phi)=\bigcup \nolimits _{\mathsf {seq}\in \mathcal {T}(I_\nu)} \mathsf {Time}(L(A_\mathsf {seq})) =\mathsf {Time}(L(A)) = L_{pt}(\psi)\).

We suggest the reader to refer to our running example (Examples 5.1, 5.7, 5.8, 5.11, 5.14) for step-by-step reduction of simple 1-TPTL formula to PnEMTL formula, Section 5.1. Example 5.1 gives reduction from simple 1-TPTL formula, \(\psi\), to an LTL formula, \(\alpha\), over interval words. Example 5.7 (Figure 4) gives the automaton, \(A_\alpha\), over interval words equivalent to \(\alpha\) constructed in Example 5.1. Example 5.8 (Figure 5) gives the construction of automaton, A, over collapsed interval words from \(A_\alpha\) constructed in Example 5.7. Example 5.11 (Figure 6) gives the construction of a normalized automaton, \(A_\mathsf {seq}\), for type one particular type \(\mathsf {seq}\) from automaton A. Finally, Example 5.14 gives a construction of PnEMTL formula \(\phi _\mathsf {seq}\) equivalent to timed behaviors encoded by automata, \(A_\mathsf {seq}\). Disjunctions over all possible types \(\mathsf {seq}\) is the required PnEMTL formula \(\phi\) equivalent to given 1-TPTL formula \(\psi\).

Fig. 4.

Fig. 4. The automaton, \(A_\alpha\) , above is equivalent to LTL formula \(\alpha\) from Example 5.1. For the sake of succinctness, \(q_1 \stackrel{a,anch}{\leftarrow }q_2\) denotes set of all transitions from \(q_1\) to \(q_2\) labelled by some subset of \(S=\lbrace \mathsf {anch}, a, b, c, (-3,0), (1,2), (0,3)\rbrace\) containing a and \(\mathsf {anch}\) , and not containing b and c. Similarly, transition labelled \(\lnot \mathsf {anch}\) denotes set of all the transitions labelled by some subset of S contains either a or b or c exclusively.

Fig. 5.

Fig. 5. The automaton, A, is collapsed version of \(A_\alpha\) in Example 5.7. Intuitively, we take intersection of all the intervals appearing in a label of a particular transition. In case the intersection is empty, we delete the transition. Following this, if \(\mathsf {anch}\) appears along with an interval that does not contain \([0,0]\) , then that transition is deleted, as the constraint enforced by that interval is in contradiction with that enforced by \(\mathsf {anch}\) . Otherwise, the transition is retained by removing the interval from the label, as the constraint enforced by the interval is already enforced by \(\mathsf {anch}\) . Note that now \(\lnot \mathsf {anch}^{\prime }\) denotes set of all the transitions labelled by some subset of S containing either a or b or c exclusively. Moreover, the labels of these transitions contain at most 1 interval.

Fig. 6.

Fig. 6. The automaton \(A_\mathsf {seq}\) depicts the construction of \(A_\mathsf {seq}\) from A (Example 5.8, Figure 5) for \(\mathsf {seq}= (-3,0)(-3,0)\mathsf {anch}(1,2)(0,3)\) . Note that the transition from \(q_0, 1\) to \(q_0,2\) is a progress transition. The behavior of \(q_0,2\) is identical to \(q_0\) of A on reading {a} and {b}. On reading {c}, it either behaves like \(q_0\) on c or like \(q_0\) on \(\lbrace c,(-3,0)\rbrace\) as the \(Status(1,(-3,0)) = mid\) .

We give a running example (Examples 5.1, 5.7, 5.8, 5.11, 5.14) along with the construction to facilitate readers in understanding the steps of the construction.

5.1 Simple TPTL to NFA over Interval Words

In this section, we elaborate the first step of the reduction.

5.1.1 Simple TPTL to LTL over Interval Words.

Let \(\gamma\) be any 1-TPTL formula without any freeze quantifier. We define \(\text{LTL}(\gamma)\) as an \(\text{LTL}\) formula obtained from \(\gamma\) by replacing clock constraints \(T-x \in I\) with I and \(x-T \in I\) with \(-I\).7 As above, \(\psi =x.\varphi\). Consider an \(\text{LTL}\) formula \(\alpha =\mathsf {F}[\text{LTL}(\varphi) \wedge \mathsf {anch}\wedge \lnot (\mathsf {F}(\mathsf {anch}) \vee \mathsf {P}(\mathsf {anch}))] \wedge \mathcal {G}(\bigvee \Sigma)\) over \(\Sigma ^{\prime }=\Sigma \cup I_\nu \cup \lbrace \mathsf {anch}\rbrace\), (\(\text{LTL}(\varphi)\) is well defined, as \(\varphi\) has no freeze quantifier). Note that all the words in \(L(\text{LTL}(\varphi))\) are \(I_\nu\)-interval words, as subformula \(\mathsf {anch}\wedge \lnot (\mathsf {F}(\mathsf {anch}) \vee \mathsf {P}(\mathsf {anch}))\) makes sure \(\mathsf {anch}\) is true at exactly one point, i.e., the point where \(\text{LTL}(\varphi)\) is asserted (condition (1) in definition of \(I_\nu\) interval word) and the conjunct \(\mathcal {G}(\bigvee \Sigma)\) makes sure that there is no such point where only propositions from \(I_\nu \cup \lbrace \mathsf {anch}\rbrace\) hold (condition (2) in definition of \(I_\nu\) interval word).

Example 5.1.

Let \(\psi = x.\varphi\) where \(\varphi = [\varphi _a \wedge \mathsf {F}\lbrace b \wedge x\in (1,2) \wedge \mathsf {F}(c \wedge x\in (0,3))\rbrace \wedge \lbrace (a \wedge x \in (-3,0) \mathsf {S}(c \wedge x \in (-3,0))\rbrace \wedge \mathcal {G}(\varphi _a \vee \varphi _b \vee \varphi _c) \wedge \mathcal {H}(\varphi _a \vee \varphi _b \vee \varphi _c)]\), where \(\varphi _a = a \wedge \lnot b \wedge \lnot c, \varphi _b = \lnot a \wedge b \wedge \lnot c, \varphi _c = \lnot a \wedge \lnot b \wedge c\). Then, \(\text{LTL}(\varphi) = [\mathsf {F}\lbrace \varphi _a \wedge \mathsf {F}(\varphi _b \wedge (1,2) \wedge \mathsf {F}(\varphi _c \wedge (0,3)))\rbrace \wedge \lbrace \varphi _a \wedge (-3,0) \mathsf {S}(\varphi _c \wedge (-3,0))\rbrace \wedge \mathsf {anch}\wedge \lnot (\mathsf {F}(\mathsf {anch}) \vee \mathsf {P}(\mathsf {anch}))]\) and \(\alpha = [\mathsf {F}\lbrace \text{LTL}(\varphi) \wedge {\mathcal {G}(\bigvee \Sigma)\rbrace }]\).

Theorem 5.2.

For any timed word \(\rho\), \(i\in dom(\rho)\), \(\rho , i \models \psi {\iff }\rho , i \in \mathsf {Time}(L(\text{LTL}(\alpha)))\).

Proof.

Note that for any timed word \(\rho = (\sigma _1,\tau _1) \ldots (\sigma _n,\tau _n)\) and \(i \in dom(\rho)\), \(\rho ,i, [x=:\tau _i] \models \varphi\) is equivalent to \(\rho , i \models \psi\). Moreover, it is straightforward that \(\alpha\) accepts all (and only) those words that are valid \(I_\nu\) interval words where the anchor point satisfies \(\text{LTL}(\varphi)\). Let \(\kappa\) be any \(I_\nu\)-interval word over \(\Sigma\) with \(\mathsf {anch}(\kappa) = i\). It suffices to prove the following:

(i)

If \(\kappa , i \models \text{LTL}(\varphi),\) then for all \(\rho \in \mathsf {Time}({\kappa })\) \(\rho ,i \models \psi .\)

(ii)

If for any timed word \(\rho\), \(\rho ,i \models \psi\), then there exists some \(I_\nu\)-interval word over \(\Sigma\) such that \(\rho , i \in \mathsf {Time}(\kappa)\) and \(\kappa , i \models \text{LTL}(\varphi)\).

Intuitively, this is because \(\text{LTL}(\varphi)\) is asserting similar timing constraints via interval words that is asserted by \(\varphi\) on the timed words directly. Note, (i) and (ii) is implied by Lemma 5.6. Substitute \(j=i\) and \(\gamma = \varphi\) in Lemma 5.6. Hence, the above theorem can be seen as a corollary of Lemma 5.6 (below).□

We give some interesting properties of interval words in the next two propositions before giving 5.6.

Proposition 5.3.

Let \(\gamma\) be any subformulae of \(\varphi\). Let \(\kappa , \kappa ^{\prime }\) be any \(I_\nu\)-interval words such that \(\kappa ^{\prime }\sim \kappa\) and for any \(i \in dom(\kappa)\) \(\kappa [i]\subseteq \kappa ^{\prime }[i]\). For any \(j \in dom(\kappa)\), if \(\kappa , j \models \text{LTL}(\gamma)\) then \(\kappa ^{\prime }, j \models \text{LTL}(\gamma)\).

Proof.

Note that \(\gamma\) is in negation normal form. Hence, any subformulae of the form \(x \in I\) will never be within the scope of a negation. Hence, \(\gamma\) can never have a subformulae of the form \(\lnot (x \in I)\). This implies that \(\text{LTL}(\gamma)\) can never have a subformulae of the form \(\lnot I\) for any \(I \in I_\nu\). We apply structural induction on depth of \(\gamma\). For base case, \(\gamma\) is a propositional logic formula and \(\text{LTL}(\gamma)\) is also a propositional logic formula over \(\Sigma\) and the statement holds trivially for any pair of similar \(I_\nu\)-interval words.

If \(\gamma = x \in I\), then \(\text{LTL}(\gamma) = I\). If \(\kappa , j \models I\), then \(I \in \kappa [j]\). This implies that \(I \in \kappa ^{\prime } [j]\) (as \(\kappa [j] \subseteq \kappa ^{\prime }[j]\)). Hence, \(\kappa ^{\prime }, j\models \text{LTL}(\gamma)\). Let \(\gamma\) be any formula such that the proposition is true for every subformula of \(\gamma\) (induction hypothesis). If \(\gamma\) is of the form \(\gamma _1 \vee \gamma _2\), and if \(\kappa , j \models \gamma\), then \(\kappa , j\models \gamma _1\) and \(\kappa , j \models \gamma _2\). By induction hypothesis, \(\kappa ^{\prime }, j \models \gamma _1\) and \(\kappa ^{\prime }, j \models \gamma _2\). Hence, \(\kappa ^{\prime }, j \models \gamma\). Similar argument holds if \(\gamma\) is of the form \(\gamma _1 \vee \gamma _2\).

If \(\gamma\) is of the form \(\gamma _1 \mathsf {U}\gamma _2\). If \(\kappa , j \models \gamma\), then (a)\(\exists j^{\prime } \gt j\) such that \(\kappa , j^{\prime } \models \gamma _2\) and \(\forall j\lt j^{\prime \prime }\lt j^{\prime }\) \(\kappa , j^{\prime \prime } \models \gamma _1\). (a) along with the induction hypothesis implies, (b)\(\exists j^{\prime } \gt j\) such that \(\kappa ^{\prime }, j^{\prime } \models \gamma _2\) and \(\forall j\lt j^{\prime \prime }\lt j^{\prime }\) \(\kappa ^{\prime }, j^{\prime \prime } \models \gamma _1\). (b) implies \(\kappa ^{\prime }, j \models \gamma\). For the case where \(\gamma\) is of the form \(\gamma _1 \mathsf {S}\gamma _2\), \(\mathcal {G}\gamma ^{\prime }\) or \(\mathcal {H}\gamma ^{\prime }\) similar argument holds.□

Proposition 5.4.

Let \(\kappa , \kappa ^{\prime }, \kappa ^{\prime \prime }\) be \(I_\nu\)-interval words such that \(\kappa \sim \kappa ^{\prime } \sim \kappa ^{\prime \prime }\) and \(\kappa [j] = \kappa ^{\prime }[j] \cup \kappa ^{\prime \prime }[j]\) for any \(j \in dom(\kappa)\). Then, \(\mathsf {Time}(\kappa) = \mathsf {Time}(\kappa ^{\prime }) \cap \mathsf {Time}(\kappa ^{\prime \prime })\).

Proof.

We need to prove that \(\rho ,i \in \mathsf {Time}(\kappa)\) iff \(\rho , i \in \mathsf {Time}(\kappa ^{\prime })\) and \(\rho , i \in \mathsf {Time}(\kappa ^{\prime \prime })\). For any \(\rho = (\sigma _1,\tau _1) \ldots (\sigma _n,\tau _n)\) and \(i \in dom(\rho)\), \(\rho ,i \in \mathsf {Time}(\kappa ^{\prime }) \cap \mathsf {Time}(\kappa ^{\prime \prime })\) \(\iff\) \(\forall j \in dom(\rho), \sigma _j = \kappa ^{\prime }[j]\cap \Sigma = \kappa ^{\prime \prime }[j] \cap \Sigma\) (as \(\kappa ^{\prime } \sim \kappa ^{\prime \prime }\)) and \(\tau _j - \tau _i \in I\) for all \(I \in (\kappa ^{\prime }[j]\cap I_\nu) \cup (\kappa ^{\prime \prime }[j]\cap I_\nu)\) \(\iff\) \(\forall j \in dom(\rho), \sigma _j = \kappa [j]\cap \Sigma\)(as \(\kappa \sim \kappa ^{\prime \prime } \sim \kappa ^{\prime }\)) and \(\tau _j - \tau _i \in I\) for all \(I \in (\kappa [j]\cap I_\nu)\)(as \(\kappa [j] = \kappa ^{\prime }[j] \cup \kappa ^{\prime \prime }[j]\)) \(\iff\) \(\rho ,i \in \mathsf {Time}(\kappa)\).□

Before giving Proposition 5.6, We need to define the notion of canonical \(I_\nu\) interval word abstraction for a given pointed timed word \(\rho , i\). Let \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\) and \(i\in dom (\rho)\).

Definition 5.5

(Canonical Abstraction).

An \(I_\nu\) interval word \(\kappa\) is a canonical \(I_\nu\) interval word abstraction of \(\rho\), denoted by \(\mathsf {Can}(I_\nu , \rho , i)\), iff \(\rho , i \in \mathsf {Time}(\kappa)\) and for any \(j \in dom(\rho)\) and \(I \in I_\nu\), \(I \in \kappa [j]\) iff \(\tau _{j} - \tau _i \in I\).

Hence, \(\kappa\) is the tightest abstraction of \(\rho , i\) with respect to the set of intervals \(I_\nu\). It is trivial to observe that \(\mathsf {Can}\) is a well defined function. We now present the main lemma, which implies Theorem 5.2.

Lemma 5.6.

Let \(\gamma\) be any subformula of \(\varphi\).

(i)

For any \(I_\nu\)-interval word \(\kappa\) and \(j \in dom(\kappa)\), \(\kappa ,j \models \text{LTL}(\gamma)\) implies for all \(\rho , i \in \mathsf {Time}(\kappa)\), \(\rho ,j, [x =: \tau _i] \models \gamma\).

(ii)

For every timed word \(\rho = (a_1,\tau _1) \ldots (a_n,\tau _n)\) and \(j \in dom(\rho)\), \(\rho ,j, [x =: \tau _i] \models \gamma\) implies \(\kappa , j \models \text{LTL}(\gamma)\) where \(\kappa = \mathsf {Can}(I_\nu , \rho ,i)\).

Proof.

We apply structural induction on \(\gamma\).

Base Case: For \(\gamma = a\) or \(\gamma = \lnot a\) where \(a \in \Sigma\), (i) and (ii) trivially holds as for every interval word \(\kappa = \sigma _1^{\prime } \ldots \sigma _n^{\prime }\) and timed word \(\rho = (\sigma _1, \tau _1) \ldots \sigma _n, \tau _n)\), \(\rho , i \in \mathsf {Time}(\kappa)\) implies \(\rho\) and \(\kappa\) agree on the set of propositions from \(\Sigma\). That is, \(\sigma ^{\prime }_j \cap \Sigma = \sigma _j\). Moreover, for any propositional formulae \(\gamma\), \(\text{LTL}(\gamma) = \gamma\) and the satisfaction of \(\gamma\) only depends on the present point. For \(\gamma = T-x \in I\), \(\text{LTL}(\gamma) = I\).

Proving (i): \(\kappa , j \models \text{LTL}(\gamma)\) would imply \(I \in \kappa [j]\). Then, for any \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\), \(\rho , i \in \mathsf {Time}(\kappa)\) only if \(\tau _j - \tau _i \in I\), which implies that \(\rho , j, [x =:\tau _i] \models \gamma\).

Proving (ii): Consider any timed word \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\) such that \(\rho ,j,[x=:\tau _i] \models \gamma\). Then, by semantics, \(\tau _j - \tau _i \in I\). By definition of canonical abstraction if \(\kappa = \mathsf {Can}(I_\nu , \rho ,i)\), then \(I \in \kappa [j]\). Hence, \(\kappa , j \models \text{LTL}(\gamma)\). Similar argument holds for \(\gamma = x-T \in I\). Note that we do not have to deal with the case \(\gamma = \lnot (T-x \in I)\) (or \(\gamma = \lnot ((x-T) \in I)\)), as the given \(\psi\) and hence (all its subformula \(\varphi\) and \(\gamma\)) are in negation normal form. This is an important observation, as the above lemma will fail to hold for \(\gamma = \lnot (T-x \in I)\). In this case, \(\text{LTL}(\gamma) = \lnot I\). Hence, all the interval words \(\kappa = \sigma ^{\prime }_1 \ldots \sigma ^{\prime }_n\) will satisfy \(\text{LTL}(\gamma)\) if \(I \notin \sigma ^{\prime }_j\). Note that this would not disallow \(\mathsf {Time}(\kappa)\) to contain a timed word \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\) such that \(\tau _j - \tau _i \in I\) (where i is the anchor point of \(\kappa\)). Just consider an example where both \(\sigma ^{\prime }_{j-1}\) and \(\sigma ^{\prime }_{j+1}\) contain I but \(\sigma ^{\prime }_j\) does not. Hence, (i) fails to hold. Intuitively, this is because the intervals in Interval words are only positive witnesses for their timing constraints. That is, presence of an interval I implies the timing constraint corresponding to I but absence of it does not imply negation of the timing constraint.

Induction: The induction case is trivial, as both the modalities of \(\text{TPTL}\) and \(\text{LTL}\) are identical with exactly the same semantics. For the sake of completeness, we enumerate this trivial argument. Let \(\gamma\) be any arbitrary formulae such that lemma holds for every subformulae of \(\gamma\) [Induction Hypothesis]. We now show that the above lemma holds \(\gamma\), too.

Case 1: Suppose \(\gamma = \gamma _1 \wedge \gamma _2\).

Proving (i): For any \(\kappa , j \models \text{LTL}(\gamma) \Rightarrow \kappa , j \models \text{LTL}(\gamma _1) \wedge \kappa , j\models \text{LTL}(\gamma _2)\). This along with the induction hypothesis (i.e., the lemma holds for \(\gamma _1\) and \(\gamma _2\)) implies, \(\forall \rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n). \rho ,i \in \mathsf {Time}(\kappa) \rightarrow \rho , j, [x=:\tau _i] \models \gamma _1\) and \(\forall \rho ^{\prime } = (\sigma ^{\prime }_1, \tau ^{\prime }_1) \ldots (\sigma ^{\prime }_n,\tau ^{\prime }_n). i \in \mathsf {Time}(\kappa) \rightarrow \rho ^{\prime }, j, [x=:\tau ^{\prime }_i] \models \gamma _2\). Which is equivalent to \(\forall \rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n). \rho ,i \in \mathsf {Time}(\kappa) \rightarrow \rho , j, [x=:\tau _i] \models \gamma\). Hence, (i) holds for \(\gamma = \gamma _1 \wedge \gamma _2\).

Proving (ii): Let \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n)\) be any arbitrary timed word and let \(i,j \in dom(\rho)\) be some arbitrary pair of points in \(\rho\). Then, \(\rho , j, [x=:\tau _i] \models \gamma \Rightarrow \rho , j, [x=:\tau _i] \models \gamma _1 \wedge \rho , j, [x=:\tau _i] \models \gamma _2\). This along with the induction hypothesis (i.e., (ii) holds for \(\gamma _1\) and \(\gamma _2\)) implies for \(\kappa = \mathsf {Can}(I_\nu , \rho ,i)\), \(\kappa , j \models \text{LTL}(\gamma _1) \wedge \kappa , j \models \text{LTL}(\gamma _2)\). Hence, \(\kappa ,j \models \text{LTL}(\gamma)\).

Case 2: Suppose \(\gamma = \gamma _1 \vee \gamma _2\).

Proving (i): For any \(\kappa , j \models \text{LTL}(\gamma)\), \(\kappa , j \models \text{LTL}(\gamma _1),\) or \(\kappa , j \models \text{LTL}(\gamma _2)\). If \(\kappa \models \text{LTL}(\gamma _1)\). Then, every timed \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n)\) where \(\rho , i \in \mathsf {Time}(\kappa)\) is s.t. \(\rho , j, [x=:\tau _i] \models \gamma _1\) (and hence \(\gamma\)) because (i) holds for \(\gamma _1\) by induction hypothesis. Similarly, if \(\kappa \models \text{LTL}(\gamma _2)\). Then, every timed \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n)\) where \(\rho , i \in \mathsf {Time}(\kappa)\) is s.t. \(\rho , j, [x=:\tau _i] \models \gamma _2\) (and hence \(\gamma\)) because (i) holds for \(\gamma _2\) (again by induction hypothesis). Hence, (i) holds for \(\gamma\), too.

Proving (ii): Suppose (ii) does not hold for \(\gamma\). This implies there exists a timed word \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n)\) such that \(\rho , j, [x =:\tau _i] \models \gamma _1 \vee \gamma _2\) for some \(i \in dom(\rho)\) but for \(\kappa = \mathsf {Can}(I_\nu , \rho ,i)\), \(\kappa , j \not\models \text{LTL}(\gamma _1)\) and \(\kappa , j \not\models \text{LTL}(\gamma _2)\). This contradicts the induction hypothesis.

Case 3: Suppose \(\gamma = \gamma _1 \mathsf {U}\gamma _2\).

Proving (i): By semantics of \(\mathsf {U}\), for any \(\kappa , j \models \text{LTL}(\gamma)\) implies (a) \(\exists j^{\prime } \gt j\) such that \(\kappa , j^{\prime } \models \text{LTL}(\gamma _2)\) and \(\forall j\lt j^{\prime \prime }\lt j^{\prime }.\) \(\kappa ,j^{\prime \prime } \models \text{LTL}(\gamma _1).\)8 As (i) holds for \(\gamma _1\) and \(\gamma _2\) by induction hypothesis, (a) implies that for any word \(\rho , i \in \mathsf {Time}(\kappa)\), (b)\(\exists j^{\prime }\gt j\) such that \(\rho , j^{\prime }, [x=:\tau _i] \models \gamma _2\) and \(\forall j\lt j^{\prime \prime }\lt j. \rho ,j^{\prime \prime }, [x=:\tau _i] \models \gamma _1\). Note that (b) iff \(\rho , j, [x=: \tau _i] \models \gamma\). Hence, (i) holds for \(\gamma\).

Proving (ii): Let \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n,\tau _n)\) be any arbitrary timed word and let \(i,j \in dom(\rho)\) be some arbitrary time points of \(\rho\). Then, \(\rho , j, [x=:\tau _i] \models \gamma\), implies that there exists a point \(j^{\prime } \gt j\) such that (c) \(\rho , j^{\prime }, [x=:\tau _i] \models \gamma _2\) and (d) for all \(j\lt j^{\prime \prime }\lt j^{\prime }\) \(\rho , j^{\prime \prime }, [x=:\tau _i] \models \gamma _1\). Let \(\kappa = \mathsf {Can}(I_\nu , \rho ,i)\). By induction hypothesis, (c) implies there exists a point \(j^{\prime } \gt j\) such that \(\kappa , j^{\prime } \models \text{LTL}(\gamma _2)\) and (d) implies for all \(j\lt j^{\prime \prime }\lt j^{\prime }\) \(\kappa , j^{\prime \prime } \models \text{LTL}(\gamma _1)\). Hence, \(\kappa , j \models \text{LTL}(\gamma)\).

Case 4: \(\gamma = \gamma _1 \mathsf {S}\gamma _2\). This case is symmetric to Case 3 and can be argued similarly.

Case 5: \(\gamma = \mathcal {G}(\gamma ^{\prime })\).

Proving (i): \(\kappa , j \models \text{LTL}(\gamma)\) iff \(\forall j^{\prime } \gt j. \kappa , j^{\prime } \models \text{LTL}(\gamma ^{\prime })\). By induction hypothesis, the lemma statement holds for \(\gamma ^{\prime }\). Hence, for every \(\rho , i \in \mathsf {Time}(\kappa)\), \(\forall j^{\prime } \gt j. \rho , j^{\prime }, [x=: \tau _i] \models \gamma ^{\prime }\). Hence, \(\rho , j, [x=: \tau _i] \models \gamma\).

Proving (ii): \(\rho , j, [x=: \tau _i] \models \gamma\). This implies \(\forall j^{\prime } \gt j. \rho , j^{\prime }, [x=: \tau _i] \models \gamma ^{\prime }\). By induction hypothesis, if \(\kappa = \mathsf {Can}(I_\nu , \rho , i)\), then \(\forall j^{\prime } \gt j. \kappa , j^{\prime } \models \text{LTL}(\gamma ^{\prime })\). Hence, \(\kappa , j \models \text{LTL}(\gamma)\).

Case 6: \(\gamma = \mathcal {H}(\gamma ^{\prime })\). This case is symmetric to Case 5 and can be argued similarly.

5.1.2 LTL to NFA over Collapsed Interval Words.

It is known that for any \(\text{LTL}[\mathsf {U},\mathsf {S}]\) formula, one can construct an equivalent NFA with at most exponential number of states [20]. We reduce the LTL formula \(\alpha\) to an equivalent NFA \(A_\alpha =(Q, \mathsf {init}, 2^{\Sigma ^{\prime }}, \delta ^{\prime }, F)\) over \(I_{\nu }\)-interval words, where \(\Sigma ^{\prime }=2^{\Sigma \cup I_{\nu } \cup \lbrace \mathsf {anch}\rbrace }\).

Example 5.7.

Consider the LTL formula \(\alpha\) from Example 5.1, Figure 4, is the automaton equivalent to \(\alpha\). Note that we constructed this automaton without using the procedure in Reference [20], as \(\alpha\) was not very complicated. But, in general, we need to rely on the procedure mentioned in Reference [20]. Moreover, Figure 5 is the collapsed automaton, A constructed from automaton \(A_\alpha\) in Figure 4.

From \(A_{\alpha }\), we construct an automaton \(A=(Q, \mathsf {init}, 2^{\Sigma ^{\prime }}, \delta , F)\) s.t. \(L(A)=\mathsf {Col}(L(A_\alpha))\). Automaton A is obtained from \(A_\alpha\) by replacing the set of intervals I on the transitions by the single interval \(\bigcap I\). In case \(\exists I_1, I_2 \in I\) s.t. \(I_1 \cap I_2 = \emptyset\) (i.e., with contradictory interval constraints), the transition is omitted in A. Also, note that \(\mathsf {anch}\) semantically implies interval \([0,0]\). Hence, all the intervals that contain \([0,0]\) along with the proposition \(\mathsf {anch}\) are omitted from the transition labels, as those intervals enforce redundant constraints (constraints that are already enforced by \(\mathsf {anch}\)). Moreover, if any transition label contains an interval I disjoint from \([0,0]\) appearing along with the proposition \(\mathsf {anch}\), then the transition is omitted, as the presence of \(\mathsf {anch}\) and I at any point j implies contradictory timing constraints on j. Note that each transition of the collapsed automaton is labelled by letters of the form S or \(S\cup \lbrace \mathsf {anch}\rbrace\) or \(S \cup \lbrace I\rbrace\) where \(S \subseteq \Sigma\) and \(I \in I_\nu = CL(I_\nu)\). This gives \(L(A)=\mathsf {Col}(L(A_\alpha))\). This implies \(\mathsf {Time}(L(A)) = \mathsf {Time}(L(A_{\alpha })) = \mathsf {Time}(L(\alpha)) = L_{pt}(\psi)\). Hence, from this point onwards, we have language of collapsed \(I_\nu\) (rather than \(I_\nu\)) interval words capturing the semantics of the given \(\text{TPTL}\) formula, \(\psi\).

Example 5.8.

Figure 5 is the collapsed automaton, A constructed from automaton \(A_\alpha\) in Figure 4, as mentioned above.

In the upcoming Sections 5.2 and 5.3, we show that we can construct a PnEMTL formula \(\phi\) using intervals in \(I_\nu\) such that it accepts all the pointed timed words in \(\mathsf {Time}(L(A))\). In general, the construction of PnEMTL formula from the NFA over collapsed interval words along with the construction of NFA over collapsed interval words from NFA over interval words (construction of A from \(A_\alpha\)) proves the following result:

Theorem 5.9.

Let \(L(A)\) be the language of any \(I_\nu\)-interval words definable by any NFA A. We can construct a \(\text{PnEMTL}\) formula \(\phi\) s.t. \(\rho ,i \models \phi\) iff \(\rho ,i \in \mathsf {Time}(L(A))\). Moreover, the number of distinct modalities is at most \(|A|\), Number of Boolean operators is in \(\mathcal {O}{2^{Poly(|A|)}}\) and arity of \(\phi\) is at most \(2|I_\nu |^2+1\).

5.2 Constructing Normalized Automata for Each Type Sequence

In this section, we elaborate on step 2 of the reduction. We discuss here how to partition W, the set of all collapsed \(I_\nu\)-interval words, into finitely many classes. Each class is characterized by its type given as a finite sequences \(\mathsf {seq}\) over \(I_{\nu } \cup \lbrace \mathsf {anch}\rbrace\). For any collapsed \(w \in W\), its type \(\mathsf {seq}\) gives an ordering between \(\mathsf {anch}(w)\), \(\mathsf {first}(w,I)\) and \(\mathsf {last}(w,I)\) for all \(I \in I_\nu\), such that, any \(I \in I_{\nu }\) appears at most twice and \(\mathsf {anch}\) appears exactly once in \(\mathsf {seq}\). For instance, \(\mathsf {seq}=I_1 I_1 \mathsf {anch}I_2 I_2\) is a sequence different from \(\mathsf {seq}^{\prime }=I_1I_2 \mathsf {anch}I_2I_1\), since the relative orderings between the first and last occurrences of \(I_1, I_2\) and \(\mathsf {anch}\) differ in both. Let the set of types \(\mathcal {T}(I_\nu)\) be the set of all such sequences; by definition, \(\mathcal {T}(I_\nu)\) is finite.

Intuition: For every type \(\mathsf {seq}\in \mathcal {T}(I_\nu)\), we construct an automaton \(A_{\mathsf {seq}}\) that accepts the normalization of all the words of type \(\mathsf {seq}\) accepted by A (i.e., \(L(A_\mathsf {seq}) = \lbrace Norm(w)| w\in L(A) \wedge w\) is of type \(\mathsf {seq}\rbrace\). Hence, \(\bigcup _{\mathsf {seq}\in \mathcal {T}(I_\nu)} \mathsf {Time}(L(A_\mathsf {seq})) = \mathsf {Time}(\mathsf {Norm}(L(A))) = \mathsf {Time}(L(A))\). Hence, the union of all these newly constructed automata encodes the required timed languages. The motivation behind construction of such an automaton is as follows: Each of the words accepted by \(A_\mathsf {seq}\) for any \(\mathsf {seq}\in \mathcal {T}(I_\nu)\) has bounded number of (as \(|\mathsf {seq}| \le 2 \times |I_\nu | + 1\)) time-restricted points. The main reason to do this is so we get automata (i.e., \(A_\mathsf {seq}\)) with structure similar to that shown in Figure 7. Such an automaton over \(I_\nu\)-interval words can be factored at time-restricted points and its corresponding timed language can then be expressed using a PnEMTL formula, \(\phi _\mathsf {seq}\), with arity bounded by the length of sequence \(\mathsf {seq}\). The construction of the required formula will be presented in Section 5.3. Hence, restricting to only normalized words makes it possible to construct a PnEMTL formula with bounded arity. Moreover, as \(\mathcal {T}(I_\nu)\) is bounded, we can get a bounded size formula, \(\phi = \mathsf {Time}(L(A))\), by disjuncting \(\phi _\mathsf {seq}\) over all possible values of \(\mathsf {seq}\in \mathcal {T}(I_\nu)\).

Fig. 7.

Fig. 7. Figure representing set of runs \(A_{\mathsf {I_1 \mathsf {anch}I_3 I_4}}\) of type Qseq where each \(S_i \subseteq \Sigma\) and each sub-automaton \(Q_i\) has only transitions without any intervals. Here, \(Qseq = T_1 T_2 T_3 T_4\) , for \(1\le i \le 4\) , \(T_i = (p_{i-1} \stackrel{S_i\cup \lbrace I_i\rbrace }{\rightarrow } q_i),\) \(I_2 = \lbrace \mathsf {anch}\rbrace\) .

Given \(w \in W\), let \(\mathsf {Boundary}(w)=\lbrace i_1, i_2, \ldots , i_k\rbrace\) be the positions of w that are either \(\mathsf {first}(w,I)\) or \(\mathsf {last}(w,I)\) for some \(I \in I_{\nu }\) or is \(\mathsf {anch}(w)\). Let \(w\downarrow _{\mathsf {Boundary}(w)}\) be the subword of w obtained by projecting w to the positions in \(\mathsf {Boundary}(w)\), restricted to the subalphabet \(2^{I_{\nu }}\cup \lbrace \mathsf {anch}\rbrace\). For example, \(w=\lbrace a,I_1\rbrace \lbrace b,I_1\rbrace \lbrace c,I_2\rbrace \lbrace \mathsf {anch},a\rbrace \lbrace b,I_1\rbrace \lbrace b,I_2\rbrace \lbrace c,I_2\rbrace\) gives \(w\downarrow _{\mathsf {Boundary}(w)}\) as \(I_1I_2\mathsf {anch}I_1I_2\). Then, w is in the partition \(W_{\mathsf {seq}}\) iff \(w\downarrow _{\mathsf {Boundary}(w)}=\mathsf {seq}\). Clearly, \(W=\bigcup _{\mathsf {seq}\in \mathcal {T}(I_\nu)}W_{\mathsf {seq}}\). Continuing with the example above, w is a collapsed \(\lbrace I_1, I_2\rbrace\)-interval word over \(\lbrace a,b,c\rbrace\), with \(\mathsf {Boundary}(w)=\lbrace 1,3,4,5,7\rbrace\), and \(w \in W_{\mathsf {seq}}\) for \(\mathsf {seq}=I_1I_2\mathsf {anch}I_1 I_2\), while \(w\notin W_{\mathsf {seq}^{\prime }}\) for \(\mathsf {seq}^{\prime }=I_1I_1\mathsf {anch}I_2 I_2\).

For type sequence \(\mathsf {seq}=I_1,I_2, \ldots , I_k\), let \(Support(\mathsf {seq})\) give the set of intervals (including \(\mathsf {anch}\)) occurring in \(\mathsf {seq}\). Each such interval occurs 1 or 2 times. Let \(Idx(\mathsf {seq})=\lbrace 1 \ldots k+1\rbrace\). We define function \(Status(\mathsf {seq}) ~:~ Idx(\mathsf {seq}) \rightarrow Support(\mathsf {seq}) \rightarrow \lbrace pre,mid,post\rbrace\) as follows: Let \(j \in Idx(\mathsf {seq})\) and \(I \in Support(\mathsf {seq})\). Then, \(Status(j)(I)=pre\) if I does not occur in \(\mathsf {seq}\) strictly before index j. Also, \(Status(j)(I)=post\) if I does not occur in \(\mathsf {seq}\) at or after index j. Finally, \(Status(j)(I)=mid\) if I occurs in \(\mathsf {seq}\) both strictly before j and also at or after index j. For example, for \(\mathsf {seq}= \mathsf {anch}~I_1 I_2 I_1\), we have \(Status(2)(I_1)=pre\) and \(Status(2)(I_2)=pre\) and \(Status(2)(\mathsf {anch})=post\). Also, \(Status(4)(I_1)=mid\) and \(Status(4)(I_2)=post\).

Let \(\mathsf {seq}\) be any sequence in \(\mathcal {T}(I_\nu)\). We construct an NFA, \(Aut_\mathsf {seq}\), which recognizes exactly the collapsed interval words, \(W_\mathsf {seq}\), of type seq. Automaton \(Aut_\mathsf {seq}=(Idx(\mathsf {seq})), 1, 2^{\Sigma ^{\prime }}, \delta _2,\lbrace |\mathsf {seq}|+1\rbrace)\). Its transitions are as follows: Let \(S \subseteq \Sigma\), \(j \in Idx(seq)\), \(I_j\) be the jth element of \(\mathsf {seq}\) and \(I \in Support(seq)\). Then,

\(\delta _2(j,S) = j\). Call such transitions as unconstrained type transitions.

\(\delta _2(j,S \cup I_j) = j+1\) if \(Status(j)(I_j)=pre\). Note that if \(I_j\) occurs exactly once in \(\mathsf {seq}\), then the status changes from pre to post after the transition, and if it occurs twice, the status changes from pre to mid after the transition. Call such transitions as progress transitions (since j increments).

If \(Status(j)(I_j)=mid\), then we have a non-deterministic choice of two transitions.

Choice 1: \(j+1 \in \delta _2(j,S \cup I_j)\). In this case, the status of \(I_j\) changes from mid to post. Call this also as progress type transition. It corresponds to accepting the second occurrence of \(I_j\).

Choice 2: \(j \in \delta _2(j,S \cup I_j)\). Call this transition as middle type of transition. This corresponds to accepting a redundant middle occurrence of \(I_j\) between its first and last occurrence.

For \(I \not= I_j\) if \(Status(j)(I)=mid\), then \(\delta _2(j,S \cup I)=j\). This also represents a middle type of transition where redundant middle occurrence of I is accepted. The position j in \(\mathsf {seq}\) is unchanged and the status of I remains mid after the transition.

\(Aut_\mathsf {seq}\) has no transitions other than given above.

The following proposition follows directly from the construction:

Proposition 5.10.

\(L(Aut_\mathsf {seq}) ~=~ W_\mathsf {seq}\).

Given collapsed interval word automaton \(A=(Q, \mathsf {init}, 2^{\Sigma ^{\prime }}, \delta , F)\) for the LTL formula as constructed above, the product automaton \(A_{prod}= (A \times Aut_\mathsf {seq})\) has the property \(L(A_{prod})=(L(A) \cap W_\mathsf {seq})\). Thus, \(A_{prod}\) accepts the collapsed words belonging to the partition \(W_\mathsf {seq}\) and accepted by A. The automaton \(A_{prod}\) has the form \(((Q \times Idx(\mathsf {seq})), (\mathsf {init}, 1), 2^{\Sigma ^{\prime }}, \delta _1, F \times \lbrace |\mathsf {seq}|+1\rbrace),\) where \(\delta _1\) is obtained by synchronous composition of \(\delta\) and \(\delta _2\) as usual. Observe that in an accepting run of \(A_{prod}\) on a word w, the progress transitions increment the index component j of the product state \((q,j)\). These transitions occur exactly at \(Boundary(w)\) positions in the word and they represent intervals in w that are retained in the normalized version of w. The middle type transitions, which leave the index component j unchanged, correspond to redundant middle intervals that do not occur in the normalized version of w.

To obtain the automaton \(A_\mathsf {seq}\) accepting normalized words corresponding to words accepted by the product \(A_{prod}\), we project out the redundant intervals in middle type transition in the automaton \(A_{prod}\). Thus, \(A_\mathsf {seq}\) has same states (including initial and final states) but its transition function \(\delta _\mathsf {seq}\) differs from the transition function \(\delta _1\) of \(A_{prod}\). Let \(A_\mathsf {seq}=((Q \times Pos(\mathsf {seq})), (\mathsf {init}, 1), 2^{\Sigma ^{\prime }}, \delta _\mathsf {seq}, F \times \lbrace |\mathsf {seq}|+1\rbrace)\). Its transitions are:

\(\delta _\mathsf {seq}((q,j),S) = \delta _1((q,j),S\). Thus, unconstrained transitions are identical to \(A_{prod}\).

If \(\delta _1((q_1,j),S \cup \lbrace I\rbrace) = (q_2,j+1),\) then \(\delta _\mathsf {seq}((q_1,j),S \cup \lbrace I\rbrace) = (q_2,j+1)\). Thus, progress transitions are identical to \(A_{prod}\).

If \(\delta _1((q_1,j),S \cup \lbrace I\rbrace) = (q_2,j),\) then \(\delta _\mathsf {seq}((q_1,j),S) = (q_2,j)\). Thus, redundant interval in middle transitions of \(A_{prod}\) are projected out.

The reader may notice the following features of \(A_\mathsf {seq}\): Let I be any element of where \(I_\nu \cup \lbrace \mathsf {anch}\rbrace\). The only transitions with labels of the form \(S \cup \lbrace I\rbrace\) (these are called time-constrained transitions) are the progress transition, and they occur in order specified by \(\mathsf {seq}\) in any accepting run. All other transitions are labelled with \(S \subseteq \Sigma\). They are unconstrained. Hence, the automaton graph partitions into disjoint subgraphs with only unconstrained transitions. These subgraphs are connected by progress transitions. See Figure 7.

Example 5.11.

Given automata A from Figure 5 in Example 5.8, we construct \(\mathcal {A}_\mathsf {seq}\) for different type of sequences accepted by A. We illustrate the construction of \(A_\mathsf {seq}\) where \(\mathsf {seq}= (-3,0)(-3,0)\mathsf {anch}(1,2)(1,2)\), in Figure 6.

From the construction of \(A_\mathsf {seq}\), the following property clearly holds:

Proposition 5.12.

\(L(A_\mathsf {seq}) ~=~ Normalize(L(A) \cap W_\mathsf {seq})\).

From the above proposition, it follows that \(\bigcup _{\mathsf {seq}\in \mathcal {T}(I_\nu)}L(A_\mathsf {seq}) = \mathsf {Norm}(L(A))\). Hence, using Theorem 5.2, we get \(\begin{equation*} \bigcup _{\mathsf {seq}\in \mathcal {T}(I_\nu)} \mathsf {Time}(L(A_\mathsf {seq})) = \mathsf {Time}(\mathsf {Norm}(L(A))) = \mathsf {Time}(L(A)) = L_{pt}(\psi) . \end{equation*}\) Hence, \(A_\mathsf {seq}\) is the required Normalized Automata for type \(\mathsf {seq}\).

5.3 Reducing NFA of Each Type to PnEMTL

Our next step is to reduce the NFAs \(A_{\mathsf {seq}}\) corresponding to each type \(\mathsf {seq}\) as constructed in the previous step to a language equivalent formula of logic \(\text{PnEMTL}\). This is step 3 of the reduction. The words in \(L(A_{\mathsf {seq}})\) are all normalized and have at most \(2|I_\nu |+1\)-time-restricted points. Thanks to this, its corresponding timed language can be expressed using \(\text{PnEMTL}\) formulae with arity at most \(2|I_\nu |\).

For each \(A_\mathsf {seq}\), we construct \(\text{PnEMTL}\) formula \(\phi _\mathsf {seq}\) such that, for a timed word \(\rho\) with \(i \in dom(\rho), \rho , i \models \phi _\mathsf {seq}\) iff \(\rho , i \in \mathsf {Time}(L(A_\mathsf {seq}))\).

5.3.1 Important Notations.

For any NFA, \(N = (St,\Sigma , i,Fin,\Delta)\), \(q \in Q\)\(F^{\prime } \subseteq Q\), let \(N[q,F^{\prime }] = (St,\Sigma , q, F^{\prime }, \Delta)\). For brevity, we denote \(N[q,\lbrace q^{\prime }\rbrace ]\) as \(N[q,q^{\prime }]\). We denote by \((N)\), the NFA \(N^{\prime }\) that accepts the reverse of \(L(N)\). The right/left concatenation of \(a \in \Sigma\) with \(L(N)\) is denoted \(N \cdot a\) and \(a \cdot N\), respectively.

Lemma 5.13.

We can construct a \(\text{PnEMTL}\) formula \(\phi _\mathsf {seq}\) with arity \(\le\) \(2|I_\nu |+1\) and size \(\mathcal {O}(|A_\mathsf {seq}|^{|\mathsf {seq}|})\) containing intervals from \(I_\nu\) s.t. \(\rho , i \models \phi _\mathsf {seq}\) iff \(\rho , i \in \mathsf {Time}(L(A_\mathsf {seq}))\).

Proof.

Let \(\mathsf {seq}=I_1\ I_2\ \ldots \ I_n\), and \(I_j=\mathsf {anch}\) for some \(1\le j \le n\).

Intuition: Note that we know the sequence, \(\mathsf {seq}\), of intervals that we will read. Moreover, this sequence is of bounded size. Hence, any accepting run will pass through at most n transitions, \(T_1, T_2, \ldots , T_n\) labelled with some interval or \(\mathsf {anch}\). Thus, the part of accepting run between \(T_i\) and \(T_{i+1}\) for \(i \in \lbrace 1,\ldots ,n-1\rbrace\) contains transitions labelled only by some non-empty subset of \(\Sigma\). Hence, the set of words read by runs between \(T_{i}\) and \(T_{i+1}\) for the set of runs passing through transitions, \(T_1, T_2, \ldots , T_n\), can be expressed by an automaton, \(A_{i+1}\), over alphabets in \(2^{\Sigma } \setminus \emptyset\).

Proof: Before starting the proof, notice the structure of \(A_\mathsf {seq}\). The state space is partitioned in to sets \(Q_1, \ldots Q_{n+1}\). Transitions within any partition \(Q_i\) are unconstrained transitions. From any state in \(Q_i\) there are constrained transition on proposition containing interval \(I_i\) that leads to some state in \(Q_{i+1}\). Hence, set of states in \(Q_i\) are reachable exactly after \(i-1\) time-constrained transitions. Let \(\Gamma =2^{\Sigma }\) and \(\mathsf {Qseq}=T_1\ T_2\ \ldots T_n\) be a sequence of time-constrained transitions of \(A_{\mathsf {seq}}\) where for any \(1 \le i \le n\), \(T_i=p_{i-1} \stackrel{S^{\prime }_{i}}{\rightarrow } q_{i}\), \(S^{\prime }_i=S_i \cup \lbrace I_i\rbrace\), \(S_i \subseteq \Sigma\), we define \(\mathsf {R}_{\mathsf {Qseq}}\) as set of accepting runs containing transitions \(T_1\ T_2\ \ldots T_n\). Hence, the runs in \(\mathsf {R}_{\mathsf {Qseq}}\) are of the following form:

\(T_{0,1}~T_{0,2} \ldots T_{0,m_0}~T_1~~T_{1,1}~ \ldots T_{1,m_1}~T_{2}~~\cdots \cdots ~~T_{n-1,1}~T_{n-1,2} \ldots T_{n}~~T_{n,1} \ldots T_{n+1}\)

where the source of the transition \(T_{0,1}\) is \(q_0\) and the target of the transition \(T_{n+1}\) is any accepting state of \(A_\mathsf {seq}\). Moreover, all the transitions \(T_{i,j}\) for \(0\le i\le n\), \(1\le j \le n_i\) are unconstrained transitions of the form \((p^{\prime } \stackrel{S_{i,j}}{\rightarrow } q^{\prime })\) where \(S_{i,j} \subseteq \Sigma\) and \(p^{\prime },q^{\prime } \in Q_{i+1}\). Hence, only \(T_1, T_2, \ldots T_n\) are labelled by any interval from \(I_\nu\). Moreover, only on these transitions the position counter (i.e., second element of the state) increments. Let \(\mathsf {A}_i = (Q_{i}, 2^{\Sigma }, q_{i-1}, \lbrace p_{i-1}\rbrace , \delta _\mathsf {seq}) \equiv A_{\mathsf {seq}}[q_{i-1},p_{i-1}]\) for \(1 \le i \le n\) and \(\mathsf {A}_{n+1}= (Q_{n+1}, 2^{\Sigma }, q_{n}, F_{\mathsf {seq}},\delta _{\mathsf {seq}}) \equiv A[q_n,F]\). Let \(\mathcal {W}_{Qseq}\) be set of words associated with any run in \(\mathsf {R}_{Qseq}\). In other words, any word w in \(\mathcal {W}_{Qseq}\) admits an accepting run on \(A_\mathsf {seq}\) that starts from \(q_0\) reads letters without intervals (i.e., symbols of the form \(S \subseteq \Sigma\)) ends up at \(p_0\), reads \(S^{\prime }_1\), ends up at \(q_1\) reads letters without intervals, ends up at \(p_1\), reads \(S^{\prime }_2\), and so on. Refer to Figure 7 for illustration. Hence, \(w \in W_{Qseq}\) if and only if \(w \in L(\mathsf {A}_1).S^{\prime }_1. L(\mathsf {A}_2).S^{\prime }_2. \cdots . L(\mathsf {A}_n).S^{\prime }_n.L(\mathsf {A}_{n+1})\). Let \(\mathsf {A}^{\prime }_{k}=S_{k-1}\cdot {\mathsf {A}_{k}}\cdot S_k\) for \(1\le k\le n+1\), with \(S_0=S_{n+1}=\epsilon\).9 Let \(\rho =(b_1, \tau _1) \ldots (b_m, \tau _m)\) be a timed word over \(\Gamma\). Then \(\rho ,i_j \in \mathsf {Time}(W_{Qseq})\) iff \(\exists\) \(1\le i_1 \le i_2 \le \cdots \le i_{j-1} \le i_j \le i_{j+1} \le \cdots \le i_n \le m\) s.t. \(\bigwedge \nolimits _{k =1}^{j-1}[(\tau _{i_k}-\tau _{i_j} \in I_k) \wedge \mathsf {Seg^-}(\rho , i_{k+1}, i_{k},\Gamma) \in L(({\mathsf {A}^{\prime }_{k}}))] \wedge \bigwedge \nolimits _{k=j}^{n}[(\tau _{i_k}-\tau _{i_j} \in I_k) \wedge \mathsf {Seg^+}(\rho , i_{k}, i_{k+1},\Gamma) \in L(\mathsf {A}^{\prime }_k)]\), where \(i_0=1\) and \(i_{n+1}=m\). Hence, by semantics of \(\mathcal {F}^{k}\) and \(\mathcal {P}^{k}\) modalities, \(\rho ,i \in \mathsf {Time}(\mathcal {W}_{Qseq})\) if and only if \(\rho , i \models \phi _{\mathsf {qseq}}\) where \(\phi _{\mathsf {qseq}}=\mathcal {P}^{j}_{I_{j-1},\ldots ,I_{1}} ((\mathsf {A}^{\prime }_1),\ldots ,(\mathsf {A}^{\prime }_j))(\Gamma) \wedge \mathcal {F}^{n-j}_{I_{j+1},\ldots ,I_{n}} (\mathsf {A}^{\prime }_{j+1},\ldots ,\mathsf {A}^{\prime }_{n+1})(\Gamma)\). Let \(\mathsf {State{-}seq}\) be the set of all possible sequences of the form \(\mathsf {Qseq}\). As \(A_\mathsf {seq}\) accepts only words that have exactly n time-restricted points, the number of possible sequences of the form \(\mathsf {Qseq}\) is bounded by \(|Q|^{n}\). Hence, any word \(\rho , i \in \mathsf {Time}(L(A_\mathsf {seq}))\) iff \(\rho , i \models \phi _{\mathsf {seq}}\) where \(\phi _{\mathsf {seq}} = \bigvee \nolimits _{\mathsf {qseq} \in \mathsf {State{-}seq}} \phi _{\mathsf {qseq}}\). Disjuncting over all possible sequences \(\mathsf {seq}\in \mathcal {T}(I_\nu)\), we get the required formula \(\phi = \bigvee \nolimits _{\mathsf {seq}\in \mathcal {T}(I_\nu)} \phi _\mathsf {seq}\).

Example 5.14.

As a continuation of our running example, we give a construction of PnEMTL formula \(\phi _\mathsf {seq}\) for automaton \(A_\mathsf {seq}\) from Example 5.11, Figure 6. Note that the accepting runs of the automaton \(A_\mathsf {seq}\) can either contain transition \(q_0,1 \rightarrow q_2,2\) bypassing \(q_1,2\) or pass via \(q_1,2\). The timed behaviors for the former (and latter) case can be captured by formula \(\phi _1\) (and \(\phi _2\), respectively), where \(\phi _1 = {\color {blue}a} \wedge \phi _{fut} \wedge \phi _{past,1}\) and \(\phi _2 = {\color {blue}a} \wedge \phi _{fut} \wedge \phi _{past,1}\) where,

\(\phi _{fut} = \mathcal {F}^2_{(1,2),(0,3)}({\color {blue}a}.\Sigma ^*.b,b.\Sigma ^*.c,c.\Sigma ^*)(\Sigma),\)

\(\phi _{past,1} = \mathcal {P}^2_{(0,3),(0,3)}({\color {blue}a}.a,a.a^*.c,c.\Sigma ^*,\Sigma ^*)(\Sigma),\)

\(\phi _{past_2} = \bigvee \nolimits _{x \in \lbrace a,b,c\rbrace }\mathcal {P}^2_{(0,3),(0,3)}({\color {red}a}.a,a.a^*.c.\Sigma ^*.x,x.\Sigma ^*)(\Sigma).\)

The \({\color {blue}a}\) in blue is the a occurring at the present position (i.e., a that occurred along with the anchor point in the interval word automata \(\mathcal {A}_\mathsf {seq}\)). Moreover, \(\phi _\mathsf {seq}= \phi _1 \vee \phi _2\).□

The construction in Section 5.2, Proposition 5.12, and proof of Lemma 5.13 imply Theorem 5.9. Note that, if \(\psi\) is a simple 1-\(\text{TPTL}\) formula with intervals in \(I_\nu\), then the equivalent \(\text{PnEMTL}\) formula, \(\phi\), constructed above contains only interval in \(\mathsf {CL}(I_\nu)\). Hence, we have the following theorem:

Theorem 5.15.

For a simple non-adjacent 1-\(\text{TPTL}\) formula \(\psi\) containing intervals from \(I_\nu\), we can construct a non-adjacent \(\text{PnEMTL}\) formula \(\phi\), s.t. for any valuation v, \(\rho ,i,v \models \psi\) iff \(\rho , i \models \phi\) where, \(|\phi |=O(2^{Poly(|\psi |)})\) and arity of \(\phi\) is at most \(2|I_\nu |^2 + 1\).

Proof.

Let \(|\psi | = m, |I_\nu | = n\).

Construct an \(\text{LTL}\) formula \(\alpha\) over interval words such that \(\rho ,i \models \varphi\) if and only if \(\rho ,i \models \mathsf {Time}(L(\alpha))\) as in Section 5.1.1 such that \(|\alpha | = O(n)\).

Reduce the \(\text{LTL}\) formula \(\alpha\) to language equivalent NFA \(A^{\prime }\) using Reference [20]. This has the complexity \(O(2^{n})\). This step is followed by reducing \(A^{\prime }\) to A over interval words over \(I_\nu\) such that \(L(A) = \mathsf {Col}(L(A^{\prime }))\). Note that \(|I_\nu | = |I_\nu |^2 = O(n^2)\) Section 5.1.2.

As shown in Section 5.2, for any type \(\mathsf {seq}\), we can construct \(A_\mathsf {seq}\) from A such that \(L(A_\mathsf {seq}) = \mathsf {Norm}(L(A_\mathsf {seq}) \cup W_\mathsf {seq})\) with number of states \(k = O(2^{Poly(m)})\).

As shown in Section 5.3, for any \(\mathsf {seq}\), we can construct \(\phi _\mathsf {seq}\) using intervals from \(I_\nu\) such that \(\rho ,i \models \phi _\mathsf {seq}\) iff \(\rho , i \in L(A_\mathsf {seq})\). Note that \(Time(L(\varphi)) = Time(L(A)) = \bigcup \nolimits _{\mathsf {seq}\in \mathcal {T}(I_\nu)} Time(L(A_\mathsf {seq}))\). Note that \(|\mathcal {T}(I_\nu)| \le (n)^{2n^2} = O(2^{Poly(n)})\). Size of formula \(\phi _\mathsf {seq}\) is \((2^{n*m}) \le 2^{m^2}\). Moreover, the arity of the formula \(\phi _{\mathsf {seq}} = 2 \times |\mathsf {seq}| = O(2 \times |I_\nu | + 1)\) (as each interval from \(I_\nu\) appears at most twice in \(\mathsf {seq}\), and \(\mathsf {anch}\) appears exactly once) = \(O(n^2)\). Hence, \(\rho , i \models \psi\) if and only if \(\rho ,i \models \phi\) where \(\phi = \bigvee \nolimits _{\mathsf {seq}\in \mathcal {T}(I_\nu)} \phi _\mathsf {seq}\) and the timing intervals used in \(\phi\) comes from \(I_\nu\). Note that if I is non-adjacent, then \(I_\nu\) is non-adjacent, too. Hence, we get a non-adjacent \(\text{PnEMTL}\) formula \(\phi\) the size of which is \(O(2^{Poly(m)})\) and arity is \(O(n^2)\).

The above theorem (Theorem 5.15) is lifted to a general (non-simple) 1-\(\text{TPTL}\) formula \(\psi\) as follows: Given a 1-\(\text{TPTL}\) formula \(\psi\) in DAG form, we first convert innermost simple sub-formulae \(\zeta ^1_i\) to their equivalent \(\text{PnEMTL}\) formulae \(\hat{\zeta ^1_i}\). We substitute a fresh witness proposition \(a^1_i\) in place of \(\zeta ^1_i\) giving formula \(\psi ^1 = \psi [a^1_i/\zeta ^1_i]\). Superscript 1 states that we have eliminated depth 1 simple subformulae. We repeat the procedure for \(\psi ^1\) giving \(\psi ^2\) where we introduce depth 2 witness propositions \(a^2_i\) for depth 1 simple sub-formula \(\zeta ^2_i\) in \(\psi ^1\). We recursively apply this procedure till a purely propositional formula \(\psi ^k\) is obtained having \(\Sigma\) as well as witness variables. We substitute top depth witness variable \(a^k_i\) by equivalent \(\text{PnEMTL}\) formulae \(\hat{\zeta ^k_i}\). This formula refers to lower-level witness variables of the form \(a^l_j\) with \(l\lt k\). We recursively substitute these witness variables by their equivalent \(\text{PnEMTL}\) formulae \(\hat{\zeta ^l_j}\), keeping the formula in DAG form. This process is repeated till we obtain a pure \(\text{PnEMTL}\) formula \(\hat{\psi }\) without witness proposition, which is equivalent to \(\psi\). Thus, we have the following result:

Theorem 5.16.

Any (non-adjacent) 1-\(\text{TPTL}\) formula \(\psi\) with intervals in \(I_\nu\) can be reduced to an equivalent (non-adjacent) \(\text{PnEMTL}\), \(\phi\), with \(|\phi |= 2^{Poly(|\psi |)}\) and arity of \(\phi =O(|I_\nu |^2)\) such that \(\rho ,i \models \psi\) iff \(\rho ,i \models \phi\).

Skip 6MSO WITH GUARDED METRIC QUANTIFIERS, GQMSO Section

6 MSO WITH GUARDED METRIC QUANTIFIERS, GQMSO

In this section, we define an extension of \(\text{MSO}[\lt ]\) with Guarded Metric Quantifiers (GQMSO). The logic is a natural extension of QMLO and Q2MLO of Hirshfeld and Rabinovich where a single metric quantifier is generalized to an anchored block of metric quantifiers of arbitrary depth. We show that \(\text{PnEMTL}\) is expressively complete for this logic. We define non-adjacency restriction in context of GQMSO and show that the non-adjacency is preserved while translating from \(\text{PnEMTL}\) to GQMSO and vice versa. Hence, the reduction (from GQMSO to \(\text{PnEMTL}\)) also serves as a proof of decidability for satisfiability checking of Non-adjacent GQMSO. This is by far the most general fragment of \(\text{MSO}[\lt ,+ \mathbb {N}]\) (syntactically) for which satisfiability checking is decidable. As a corollary, we get that (non-adjacent) 1-TPTL is expressively complete for (non-adjacent) GQFO, the first-order fragment of (non-adjacent) GQMSO.

6.1 GQMSO: Syntax and Semantics

We define a real-time logic GQMSO that is interpreted over timed words. It includes \(\text{MSO}[\lt ]\) over words with respect to some alphabet \(\Sigma\). This is extended with a notion of time-constraint formula \(\psi (t)\), where t is a free first-order variable. All variables in our logic range over positions in the timed word and not over timestamps (unlike continuous interpretation of these logics). There are two sorts of formulae in GQMSO that are mutually recursively defined: \(\text{MSO}^{\mathsf {UT}}\) and \(\text{MSO}^{\mathsf {T}}\) (where \(\mathsf {UT}\) stands for untimed and \(\mathsf {T}\) for timed). An \(\text{MSO}^{\mathsf {UT}}\) formula \(\phi\) has no real-time constraints except for the time-constraint subformula \(\psi (t) \in \text{MSO}^{\mathsf {T}}\). A formula \(\psi (t)\) has only one free variable t (called anchor), which is a first-order variable. \(\psi (t)\) is defined as a block of real-time-constrained quantification applied to a GQMSO formula with no free second-order variables; it has the form \(\mathcal {Q}_1t_1. \mathcal {Q}_2t_2. \dots \mathcal {Q}_jt_j. ~ \phi (t,t_1,\ldots t_j)\) where \(\phi \in \text{MSO}^{\mathsf {UT}}\). All the metric quantifiers in the quantifier block constrain their variable relative only to the anchor t. The precise syntax follows below.10

Remark: This form of real time constraints in first-order logic were pioneered by Hirshfeld and Rabinovich [25] in their logic Q2MLO (with only non-punctual guards) and its punctual extension was later shown to be expressively complete to FO[\(\lt ,+1\)] by Hunter [28] over signals. Here, we extend the quantification to an anchored block of quantifiers of arbitrary depth.

We have a two sorted logic consisting of \(\text{MSO}^{UT}\) formulae \(\phi\) and time-constrained formulae \(\psi\). Let \(a \in \Sigma\), and let \(t,t^{\prime }\) range over first-order variables, while T range over second-order variables. The syntax of \(\phi \in \text{MSO}^{\mathsf {UT}}\) is given by:

\(t=t^{\prime }~|~t\lt t^{\prime }~|~Q_a(t)~|~T(t) \mid \phi \wedge \phi ~|~{\lnot } \phi ~|~ {\exists } t. \phi ~|~{\exists } T \phi ~ |~\psi (t)\).

Here, \(\psi (t) \in \text{MSO}^{\mathsf {T}}\) is a time-constraint formula whose syntax and semantics are given a little later. A formula in \(\text{MSO}^{\mathsf {UT}}\) with first-order free variables \(t_0,t_1, \ldots t_k\) and second-order free variables \(T_1, \ldots , T_m\) is denoted \(\phi (t_0,\ldots t_k,T_1, \ldots , T_m)\). The semantics of such formulae is as usual. Let \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _n, \tau _n)\) be a timed word over \(\Sigma\). Given \(\rho\), positions \(i_0, \ldots , i_k\) in \(dom(\rho)\), and sets of positions \(A_1, \ldots , A_m\) with \(A_i \subseteq dom(\rho)\), we define \(\rho ,(i_0,i_1,\ldots ,i_k,A_1,\ldots ,A_m) \models \phi (t_0,t_1, \ldots t_k,T_1, \ldots , T_m)\) inductively in \(\text{MSO}[\lt ]\).

\((\rho ,i_0,\ldots ,i_k,A_1,\ldots ,A_m) \models\) \(t_x \lt t_y\) iff \(i_x \lt i_y\),

\((\rho ,i_0,\ldots ,i_k,A_1,\ldots ,A_m) \models\) \(Q_a(t_x)\) iff \(a \in \sigma _{i_x}\),

\((\rho ,i_0,\ldots ,i_k,A_1,\ldots ,A_m) \models\) \(T_j(t_x)\) iff \(i_x \in A_j\),

\((\rho ,i_0,\ldots ,i_k,A_1,\ldots ,A_m) \models\) \(\exists t^{\prime }.\phi (t_0, \ldots t_k,t^{\prime },T_1, \ldots , T_m)\) iff \((\rho ,i_0,\ldots ,i_k,i^{\prime },A_1,\ldots ,A_m) \models \phi (t_0, \ldots t_k,t^{\prime },T_1, \ldots , T_m)\) for some \(i^{\prime } \in dom(\rho)\).

The time-constraint formula \(\psi (t) \in \text{MSO}^{\mathsf {T}}\) has the form: \(\mathcal {Q}_1t_1. \mathcal {Q}_2t_2. \dots \mathcal {Q}_jt_j. ~ \phi (t,t_1,\ldots t_j)\) where \(t_1, \ldots , t_j\) are first-order variables and \(\phi \in \text{MSO}^{\mathsf {UT}}\). Each quantifier \(\mathcal {Q}_x t_x\) has the form \(\overline{\exists }t_x \in t+ I_x\) or \(\overline{\forall }t_x \in t+ I_x\) for a time interval \(I_x \in \mathcal {I}_\mathsf {int}\). \(\mathcal {Q}_x\) is called a metric quantifier. Note that each metric quantifier constrains its variable only relative to the anchor variable t. Moreover, \(\psi (t)\) has no free second-order variables. The semantics of such an anchored metric quantifier is obtained recursively as follows: Let

\((\rho ,i_0,i_1,\ldots ,i_{j-1}) \models \overline{\exists }t_j \in t+I_j. \phi (t,t_1,\ldots ,t_j)\) iff \(\begin{Bmatrix} \text{there exists } i_j \text{ such that }\tau _{i_j} \in \tau _{i_0}+I_j \text{ and,}\\ (\rho ,i_0,i_1 \ldots i_j) \models \phi (t,t_1,\ldots , t_j) \end{Bmatrix}\),

\((\rho ,i_0,i_1,\ldots ,i_{j-1}) \models \overline{\forall }t_j \in t+I_j. \phi (t,t_1,\ldots t_j)\) iff \(\begin{Bmatrix} \text{for all } i_j \text{ such that }\tau _{i_j} \in \tau _{i_0}+I_j \text{ implies,}\\ (\rho ,i_0,i_1 \ldots i_j) \models \phi (t,t_1,\ldots , t_j) \end{Bmatrix}\).

Note that metric quantifiers quantify over positions of the timed word, and the metric constraint is applied on the timestamp of the corresponding positions. Each time-constraint formula in GQMSO has exactly one free variable; variables \(t_1,\ldots ,t_j\) are called time-constrained in \(\psi (t)\). If we restrict the grammar of a time-constrained formula \(\psi (t) \in \text{MSO}^{\mathsf {T}}\) to contain only a single metric quantifier (i.e., \(\mathcal {Q}_1t_1. \phi (t,t_1)\)) and disallow the usage of second-order quantification, then we get the logic Q2MLO of Reference [26].

Example 6.1.

Consider sequences over \(\Sigma =\lbrace a,b\rbrace\) such that the event a is the last event in the first unit interval. \(\phi = \exists t. [\lbrace \forall t^{\prime }. t\le t^{\prime }\rbrace \wedge \lbrace \exists s \in t + (0,1). \forall s^{\prime } \in t+(0,1).(s\ge s^{\prime } \wedge Q_a(s))\rbrace ]\).

Example 6.2.

Consider sequences over events \(\Sigma =\lbrace a,b\rbrace\) such that from every a there was a positive even number of b’s in the previous unit interval. \(\phi = \forall t. Q_a(t) \rightarrow \psi (t)\) where \(\psi (t) = [ \overline{\exists }t_{f} \in t+ [-1,0]. \overline{\exists }t_{l} \in t+[-1,0] \overline{\forall }t^{\prime } \in t+[-1, 0]. \gamma (t,t_f,t_l,t^{\prime })\) where \(\gamma (t,t_f,t_l,t^{\prime }) = t_f \le t^{\prime } \le t_l \wedge \exists X_{o}. \exists X_e. X_o(t_f) \wedge X_e(t_l) \wedge \forall t_1. \forall t_2.\) \([\lbrace Q_b(t_1) \wedge Q_b(t_2) \wedge \forall t_3. (t_1\lt t_3\lt t_2 \rightarrow \lnot Q_b(t_3))\rbrace \rightarrow\) \(\lbrace (X_o(t_1) \wedge \lnot X_e(t_1) \wedge X_e(t_2) \wedge \lnot X_o(t_2)) \vee (X_e(t_1) \wedge \lnot X_e(t_1) \wedge X_o(t_2) \wedge \lnot X_o(t_2))\rbrace ]\). Here, \(\phi\) is a formula of type \(\text{MSO}^{\mathsf {UT}}\) containing the subformula \(\psi (t)\) of type \(\text{MSO}^T\), which in turn contains the formula \(\gamma (t,t_f,t_l,t^{\prime })\) of type \(\text{MSO}^{\mathsf {UT}}\).

Note that, while GQMSO extends classical MSO\([\lt ]\), it is not closed under second-order quantification: Arbitrary use of second-order quantification is not allowed, and its syntactic usage, as explained above, is restricted to prevent a second-order free variable from occurring in the scope of the real-time constraint (similar to References [23, 43, 46]). For example, \(\exists X. \exists t. [X(t) \wedge \overline{\exists }t^{\prime } \in t+(1,2) Q_a(t^{\prime })]\) is a well-formed GQMSO formula, while \(\exists X. \exists t. \overline{\exists }t^{\prime } \in t+(1,2)[Q_a(t^{\prime })\wedge X(t)]\) is not, since X occurs freely within the scope of the metric quantifier.

Example 6.3.

We define a language \(\mathsf {L_{inst err}}\) over the singleton alphabet \(\Sigma = \lbrace b\rbrace\) accepting words satisfying the following conditions:

(1)

One b with timestamp 0 at the first position. (Positions are counted \(1,2,3,\ldots\)).

(2)

Exactly two points in the interval \((0,1)\) at positions 2 and 3 with timestamps called \(\tau _2\) and \(\tau _3\), respectively.

(3)

Exactly one b in \([\tau _2+1, \tau _3 +1]\) at some position p. Other b’s can occur freely elsewhere.

The above language was proposed by Lasota and Walukiewicz [34] (Theorem 2.8) as an example of language not recognizable by 1 clock Alternating Timed Automata but expressible by a Deterministic Timed Automata with 2 clocks. Let \(S(u,v)\) be the FO[\(\lt\)] formula specifying the successor relation (i.e., \(u = v+1\)). This can be specified as the GQMSO formula \(\psi = \psi _1 \wedge \psi _3\), where

(1)

Let \(Pos_1(t) = \lnot \exists w. S(t, w)\), \(Pos_i(t) = \exists t^{\prime }. S(t,t^{\prime }) \wedge Pos_{i-1}(t^{\prime })\). Hence, \(Pos_i(t)\) holds only when \(t=i\), where \(i \in \lbrace 1,2,3,4\rbrace\).

(2)

Let \(\psi _1=\exists t_1. ~Pos_1(t_1) {\wedge } (\overline{\exists }t_2 \in t_1+(0,1). \overline{\exists }t_3 \in t_1 +(0,1). [Pos_2(t_2) \wedge Pos_3(t_3) \wedge \lnot \overline{\exists }t \in t_1+(0,1). Pos_4(t)]\). This states that exactly two positions exist in the initial unit time interval \((0,1)\). Let their timestamps be \(\tau _2\) and \(\tau _3\).

(3)

Let \(\psi _2(p) = [~\overline{\exists }t \in p+[-1,0). Pos_3(t) ~~\wedge ~~ \lnot \overline{\exists }t \in p+(-1,0). Pos_2(t)~]\). This states that position p lies within \([\tau _2+1, \tau _3 +1]\).

(4)

\(\psi _3 ~=~ \exists p. ~[\psi _2(p) ~\wedge (\forall q. \psi _2 (q) ~\rightarrow (p=q))]\) states that there is exactly one position satisfying property \(\psi _2\).

Metric Depth. The metric depth of a formula \(\varphi\) denoted (\(\mathsf {MtD}(\varphi)\)) gives the nesting depth of time constraint constructs and is defined inductively: For atomic formulae \(\varphi\), \(\mathsf {MtD}(\varphi)=0\). \(\mathsf {MtD}[\varphi _1 \wedge \varphi _2]= \mathsf {MtD}[\varphi _1 \vee \varphi _2] = max(\mathsf {MtD}[\varphi _1],\mathsf {MtD}[\varphi _2])\) and \(\mathsf {MtD}[\exists t. \varphi (t)]=\mathsf {MtD}[\lnot \varphi ]=\mathsf {MtD}(\varphi (t))\). \(\mathsf {MtD}[\mathcal {Q}_1t_1 \ldots \mathcal {Q}_j t_j \phi ]\) \(=\) \(\mathsf {MtD}[\phi ] + 1\). For example, the sentence \(\forall t_3~\overline{\forall }t_1 \in t_3 + (1,2)~ \lbrace Q_a(t_1) {\rightarrow } (\overline{\exists }t_0 \in t_1 + [1,1]~ Q_b(t_0))\rbrace\) accepts all timed words such that for each a that is at distance \((1,2)\) from some timestamp t, there is a b at distance 1 from it. This sentence has metric depth two with time-constrained variables \(t_0,t_1\).

6.2 GQMSO with Alternation Free Metric Quantifiers (AF-GQMSO)

We define a syntactic fragment of GQMSO, called AF-GQMSO, where all the metric quantifiers in any anchored metric quantifier block only consist existential metric quantifiers. More precisely, AF-GQMSO is a syntactic fragment of GQMSO where the time constraint \(\psi (t_0)\) has the form \(\overline{\exists }t_1 \in t_0+I_1. \overline{\exists }t_2 \in t_0 + I_2. \dots \overline{\exists }t_j \in t_0 +I_j. ~ \phi (t_0,t_1,\ldots t_j)\) with \(\phi \in \text{MSO}^{\mathsf {UT} }\). Hence, there is no alternation of metric quantifiers within a block of the metric quantifier. Note that the negation of the timed subformula is allowed in the syntax of GQMSO (and hence AF-GQMSO). Hence, alternation free \(\overline{\forall }^* \phi\) formulae can also be expressed as equivalent \(\lnot \overline{\exists }^* \lnot \phi\) using AF-GQMSO. We now show that, surprisingly, AF-GQMSO is as expressive as GQMSO.

Theorem 6.4.

The subclass AF-GQMSO is expressively equivalent to GQMSO.

Proof.

Let \(\psi (t_0) = \mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_j t_j. \varphi (t_0, \ldots , t_j)\) be any GQMSO formula where every quantifier \(\mathcal {Q}_i t_i\) is of the form \(\overline{\exists }t_i\in t_0 + I_i\) or \(\overline{\forall }t_i \in t_0 + I_i\). Let \(I_j = [l,u)\) (similar construction can be given for all other type of intervals). We convert the innermost metric quantifier \(\mathcal {Q}_j t_j\) to a non-metric quantifier by adding (at most) four existential metric quantifiers at the top level of the form \(\overline{\exists }t^{\prime }_{j} \in t_0+(-\infty ,l). \overline{\exists }t^{\prime \prime }_{j} \in t_0+[u,\infty).\overline{\exists }t_{first,j} \in I_j. \overline{\exists }t_{last, j} \in I_j.\). Intuitively, the variables \(t^{\prime }_j\) will take the value of the last point within interval \((-\infty , l)\) from \(t_0\). Similarly, \(t^{\prime \prime }_j\) will take the value of the first point within interval \([u,\infty)\) from \(t_0\). Moreover, variable \(t_{first,j}\) (\(t_{last, j}\)) will take the value of the first (last, respectively) point within interval \(I_j\) from \(t_0\). Hence, we can replace the quantifier \(\mathcal {Q}_j t_j\) with \(\exists t_{first, j} \le t \le t_{last, j}\) if \(\mathcal {Q}_j\) is an existential metric quantifier and with \(\forall t_{first, j} \le t_j \le t_{last, j}\) if \(\mathcal {Q}_j\) is a universal metric quantifier. Hence, repeating the above steps for \(\mathcal {Q}_{j-1} \ldots \mathcal {Q}_1\), we get a AF-GQMSO with at most 4j existential metric quantifiers.

For \(1\le I \le j\), Let \(\varphi _i\) be the subformula \(\mathcal {Q}_i t_i. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_j t_j. \varphi (t_0, \ldots , t_j)\). Other types of intervals can be handled similarly. We eliminate \(\mathcal {Q}_{j} t_j\) as follows:

(1)

If there is no point within \([l,u)\) of \(t_0\), then the sub- formulae \(\varphi _j\) vacuously evaluates to true if \(\mathcal {Q}_j\) is a universal metric quantifier and evaluates to false if \(\mathcal {Q}_j\) is an existential metric quantifier. \(C_{1} = \lnot \overline{\exists }t \in I_j \rightarrow \mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_{j-1} t_{j-1}. \gamma\), where \(\gamma _1 = true\) in case \(\mathcal {Q}_j\) is a universal metric quantifier and \(\gamma _1 = false\) otherwise.

(2)

If there is a point in \([l,u)\) from \(t_0\), then we add existential metric quantifiers \(\overline{\exists }t_{first,j} \in I_j. \overline{\exists }t_{last, j} \in I_j.\) (along with some more existential metric quantifiers) at the top level and assert a formula that forces \(t_{first,j}\) to be the first point within interval \([l,u)\) from \(t_0\) and \(t_{last, j}\) to be the last. Then, the quantifier \(\overline{\exists }. t_{j} \in t_0 + I_j\) (\(\overline{\forall }. t_{j} \in t_0 + I_j\)) can be replaced by \(\exists . t_{first, j} \le t_j \le t_{last, j}\) (\(\forall . t_{first, j} \le t_j \le t_{last, j}\), respectively). Let \(\gamma _2 = \forall t_{first, j } \le t_j \le t_{last, j} \varphi (t_0, \ldots , t_j)\) if \(\mathcal {Q}_j\) is a universal metric quantifier else \(\gamma _2 = \exists t_{first, j } \le t_j \le t_{last, j} \varphi (t_0, \ldots , t_j)\) otherwise. Let \(S(t,t^{\prime })\) be the successor predicate that is true iff \(^{\prime } = t+1\). It is routine to express such a predicate in MSO[\(\lt\)]. Then, \(C_{2} = \overline{\exists }t \in t_0 + I_j \rightarrow C_{2,1} \vee C_{2,2} \vee C_{2,3} \vee C_{2,4}\) where:

\(C_{2,1}\) covers the possibility that there are points that occur within interval \((\infty , l)\) and \([u, \infty)\) from \(t_0\). Hence,

\(\begin{matrix} & \overline{\exists }t^{\prime }_{j} \in t_0+(-\infty ,l). \overline{\exists }t_{j}^{\prime \prime } \in t_0+[u,\infty). & \\ C_{2,1} \!\!&=\ \overline{\exists }t_{first, j} \in t_0+[l,u). \overline{\exists }t_{last,j} \in t_0+ [l,u).&\\ &\mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_{j-1} t_{j-1}. & \end{matrix} \begin{Bmatrix}S(t^{\prime }_j, t_{first, j}) \wedge \\ S(t_{last,j}, t^{\prime \prime }_j) \wedge \\ \gamma _2 \end{Bmatrix}\)

\(C_{2,2}\) covers the possibility where there is no point occurring within \((-\infty , l)\) but there are points occurring within \([u, \infty)\) from \(t_0\). Hence, \(\begin{matrix} & \overline{\exists }t^{\prime \prime }_{j} \in t_0+[u,\infty). & \\ C_{2,2} \!\!&=\ \overline{\exists }t_{first, j} \in t_0+[l,u). \overline{\exists }t_{last,j} \in t_0+ [l,u).&\\ &\mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_{j-1} t_{j-1}. & \end{matrix} \begin{Bmatrix}\forall t. t \ge t_{first, j} \wedge \\ S(t_{last,j}, t^{\prime \prime }_j) \wedge \\ \gamma _2 \end{Bmatrix}\)

\(C_{2,3}\) covers the possibility where there are points occurring within interval \((-\infty , l)\) but there is no point within \([u, \infty)\) from \(t_0\). Hence, \(\begin{matrix} & \overline{\exists }t^{\prime }_{j} \in t_0+(-\infty ,l). & \\ C_{2,3} \!\!&=\ \overline{\exists }t_{first, j} \in t_0+[l,u). \overline{\exists }t_{last,j} \in t_0+ [l,u).&\\ &\mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_{j-1} t_{j-1}. & \end{matrix} \begin{Bmatrix}S(t^{\prime }_j, t_{first, j}) \wedge \\ \forall t. t\le t_{last, j} \wedge \\ \gamma _2 \end{Bmatrix}\)

\(C_{2,4}\) covers the possibility where there is no point occurring within interval \((-\infty , l)\) and \([u, \infty)\) from \(t_0\). Hence, \(C_{2,4} = \overline{\exists }t_{first,j} \in t_0+ [l,u). \overline{\exists }t_{last, j} \in I_j. \mathcal {Q}_1 t_1. \mathcal {Q}_2 t_2. \ldots \mathcal {Q}_{j-1} t_{j-1}. (\forall t. t\ge t_{first, j} \wedge t\le t_{last, j} \wedge \gamma _2)\).

Hence, \(C_1 \wedge C_2\) is the required formula. Note that irrespective of \(\mathcal {Q}_j\) being a universal or existential quantifier, the new metric quantifiers that we add at the top level are only existential metric quantifiers. Hence, when we apply the above reduction for j steps, we will be able to get rid of all the \(\mathcal {Q}_1 \ldots \mathcal {Q}_j\) metric quantifiers (and hence the alternations within that block) and end up getting formulae where each time-constrained subformula contains a block of at most 4k existential metric quantifiers.

6.3 Non-Adjacent GQMSO (NA-GQMSO)

Any AF-GQMSO formula \(\varphi\) is said to be non-adjacent if and only if for every subformula \(\psi\) of \(\varphi\) of the form \(\overline{\exists }t_1 \in t + I_1 \ldots \overline{\exists }t_j \in t+I_j \Phi (t,t_1,\ldots ,t_j)\), the set of intervals \(\lbrace I_1, \ldots , I_j\rbrace\) is non-adjacent. Notice that NA-GQMSO is a syntactic subclass of AF-GQMSO. For example, \(\overline{\exists }t_1 \in t_0 +(2,3) \overline{\exists }t_2. \in t_0+ (3,4) [\exists t \lt t_0 \wedge \overline{\exists }t_3 \in t_0+(4,5)]\) is not non-adjacent, as intervals \((2,3)\) and \((3,4)\) appear within the same metric quantifier block and are adjacent. However, \(\overline{\exists }t_1 \in t_0 +(2,3) \overline{\exists }t_2. \in t_0+ (4,5) [\exists t \lt t_0 \wedge \overline{\exists }t_3 \in t_0+(3,4)]\) is non-adjacent, as \(\lbrace (1,2), (4,5)\rbrace\) and \((2,3)\) is non-punctual (and hence non-adjacent to itself). The formula in Example 6.3 is also an NA-GQMSO formula.

Skip 7CLASSICAL LOGIC CHARACTERIZATION OF PnEMTL Section

7 CLASSICAL LOGIC CHARACTERIZATION OF PnEMTL

In this section, we prove the following main theorem:

Theorem 7.1.

PnEMTL \(\cong\) GQMSO. Moreover, Non-adjacent PnEMTL \(\cong\) Non-adjacent GQMSO.

The theorem follows from Lemmas 7.2 and 7.3 given below.

Lemma 7.2.

PnEMTL \(\subseteq\) GQMSO.

Proof.

The key observation is that conditions of the form \(\mathsf {Seg}(i,j,\rho ,S) \in L(\mathsf {A})\) can be equivalently expressed as MSO[\(\lt\)] formulae \(\psi _{\mathsf {A}}(i,j)\) using Büchi Elgot Trakhtenbrot (BET) Theorem [16, 30, 44]. Replacing the former with latter, we get an equivalent AF-GQMSO formula (which is a syntactic subset of GQMSO), as shown below. We apply induction on modal depth of the given formula \(\varphi\). For modal depth 0, \(\varphi\) is a propositional formula and hence it is trivially an AF-GQMSO formula.

Let \(\varphi\) be a modal depth 1 formula of the form \(\mathcal {F}^{k}_{I_1, \ldots , I_k}(\mathsf {A}_1, \ldots , \mathsf {A}_{k+1})(\Sigma)\). We can easily translate the above to equivalent GQMSO formula \(\exists t_1 \in t+ I_1 \ldots \overline{\exists }t_j \in t+I_j~ \Phi (t,t_1,\ldots ,t _j),\) where \(\Phi (t,t_1,\ldots ,t_j)= \exists t_{k+1} ~ \psi _{A_1}(t_0,t_1) \wedge \cdots \wedge \psi _{A_{k}}(t_{k-1},t_k) \wedge \psi _{A_{k+1}}(t_k,t_{k+1}) \wedge EP(t_{k+1})\). Note that the GQMSO formula directly encodes the semantics of \(\mathcal {F}^{k}\) formula and hence their equivalence is clear by construction. The \(\mathcal {P}^{k}\) modality is handled similarly. Also note that this reduction preserves the non-adjacency. Dealing with Boolean operators is trivial, as the AF-GQMSO is closed under Boolean operations.

For the induction step, we assume that the lemma holds for all the PnEMTL formulae of modal depth \(\lt n\). Let \(\varphi = \mathcal {F}^{k}_{I_1, \ldots , I_k}(\mathsf {A}_1, \ldots , \mathsf {A}_{k+1})(\Sigma \cup S)\) of modal depth n. Therefore, S is a set of PnEMTL formula with modal depth \(\lt n\). We associate a unique new witness proposition with every subformulae in S and replace all the subformulae in S by their corresponding witness propositions getting a formula \(\varphi ^{\prime }\) of modal depth 1. As with the base case, we can construct an AF-GQMSO formula \(\psi ^{\prime }\) equivalent to \(\varphi ^{\prime }\). By inductive hypothesis, every subformulae \(\varphi _i\) in S can be reduced to an equivalent AF-GQMSO formula \(\psi _i\). We replace all the witnesses of \(\varphi _i\) by \(\psi _i\) getting an equivalent formulae \(\psi\) over \(\Sigma\). Note that if formula \(\varphi _i\) in S are non- adjacent, then, by induction hypothesis, equivalent \(\psi _i\) are in NA-GQMSO formula. Similarly, if \(\varphi ^{\prime }\) is NA-PnEMTL formula, then \(\psi ^{\prime }\) is NA-GQMSO formula. Hence, if \(\varphi\) in non-adjacent, then equivalent formula \(\psi\) is non-adjacent, too.□

Lemma 7.3.

GQMSO \(\subseteq\) PnEMTL.

Proof.

It suffices to show AF-GQMSO \(\subseteq\) PnEMTL (thanks to Theorem 6.4). The proof is done via induction on metric depth of the AF-GQMSO formulae.

(Base case) Let \(\psi (t_0) = \overline{\exists }t_1 \in t_0 + I_1 \ldots \overline{\exists }t_j \in t_0 + I_j. \varphi (t_0,t_1, \ldots , t_j)\) be any AF-GQMSO formula of metric depth 1. Then, \(\varphi (t_0,t_1, \ldots , t_j)\) is an untimed MSO formula over \(\Sigma _1=\Sigma \cup \lbrace t_0,\ldots ,t_j\rbrace\). By Büchi Elgot Trakhtenbrot Theorem [16, 30, 44], we can construct a finite state automaton \(A_1\) accepting same models as \(\varphi\). Note that the alphabet of \(A_1\) is \(2^{\Sigma _1}\). Every word \(\alpha\) accepted by \(A_1\) has exactly one position \(i_k\) where \(t_k \in \alpha [j]\). Hence, with some abuse of notation, we can write \(\alpha =\sigma \oplus (t_0\partial i_0,t_1\partial i_1, \ldots , t_j\partial i_j)\) and \(\sigma \in 2^\Sigma\). By the semantics of GQMSO, any pointed word \(\rho ,i \models \psi (t_0)\) iff \(\exists i_1, i_2, \ldots , i_j\) such that \(\tau _i - \tau _{i_1} \in I_1 \wedge \cdots \wedge \tau _i - \tau _{i_j} \in I_j\) and word \(untime(\rho)\oplus (t_0\partial i_0,t_1\partial i_1, \ldots , t_j\partial i_j)\) is accepted by \(A_1\). We modify \(A_1\) to give an I Interval word automaton \(A_2\) as follows: If label of an edge is \(S \subseteq \Sigma ^{\prime }\), then we relabel it with \(S^{\prime } \subseteq \Sigma \cup \lbrace anch,I_1,\ldots ,I_j\rbrace ,\) where anch replaces \(t_0\) and \(I_i\) replaces \(t_i\) in S. There is one-to-one correspondence between transitions of \(A_1\) and \(A_2\) where presence of interval \(I_i\) symbolically enforces the timing constraint. Hence, it is easy to see that \(\rho ,i \models \psi (t_0)\) iff \(\rho ,i \in \mathsf {Time}(L(A_2))\).

By the construction given in Section 5, for any NFA A over I interval words, we can construct a PnEMTL formulae \(\varphi (A)\) such that for any pointed timed word \(\rho ,i\), we have \(\rho ,i \in Time (L(A))\) iff \(\rho ,i \models \varphi\). Hence, \(\rho ,i \models \psi (t_0)\) iff \(\rho ,i { \in } Time(L(A_2))\) iff \(\rho , i \models \varphi (A_2)\). Moreover, if \(\psi\) is non-adjacent, then I is non-adjacent and thus \(\varphi\) is in NA-PnEMTL.

(Induction step) Assume that the lemma holds for all formulas of depth less than n. Let \(\psi (t_0)\) be any time-constraint formula of AF-GQMSO having metric depth n. With every timed subformulae \(\psi _i(t)\) of \(\psi\), we associate a witness proposition \(b_i\) such that \(b_i\) holds iff \(\psi _i\) holds. Let W be the set of witnesses. We replace each subformula \(\psi _i(t)\) of type \(MSO^T\) with its corresponding witness getting a formula \(\psi ^{\prime }(t_0)\) of metric depth 1. As shown in the base case, we can construct a PnEMTL formula \(\varphi ^{\prime }\) equivalent to \(\psi ^{\prime }(t_0)\) containing symbols from \(\Sigma \cup W\). Note that all subformulae \(\psi _i(t_0)\) of \(\psi\) are of metric depth less than n. Hence, by the induction hypothesis, we can construct a PnEMTL formula \(\varphi _i\) equivalent to \(\psi _i(t_0)\). Hence, the witnesses for \(\psi _i\) are also that for \(\varphi _i\). Replacing the witnesses \(b_i\) with its corresponding PnEMTL formulae \(\varphi _i\), we get the required PnEMTL formulae \(\varphi\). Also note that if \(\psi\) is non-adjacent, then all its subformulae \(\psi _i\) and formula \(\psi ^{\prime }\) are non-adjacent, too. This implies that formulae \(\varphi _i\), \(\varphi ^{\prime }\) and, hence \(\varphi\) are NA-PnEMTL formulae. We give a small toy example as follows: In this example, we write a regular expression, in place of NFA wherever required, for the sake of succinctness and readability.

Example 7.4.

Consider a GQMSO formulae \(\psi (t) = \overline{\exists }t_1 \in t+(0,1) \overline{\exists }t_2 \in t+(-1,0) \psi _{even,b}(t,t_1) \wedge \psi _{odd,a}(t,t_2)\), where \(\psi _{even,b}(x,y)(\psi _{odd,a}(x,y))\) is an MSO[\(\lt\)] formula that is true iff the number of b’s (a’s, respectively) between x and y (including x and y) is even (odd, respectively). The regular expression of the behavior starting from the beginning would be of the form: \(\mathbf {(a+b)^* \cdot \lbrace (a+b), x \in (-1,0)\rbrace \cdot (b^*.a.b^*.a.b^*)^*\cdot a\cdot b^*)\cdot } \mathbf {\mathsf {anch}\cdot (a^*\cdot b\cdot a^*\cdot b\cdot a^*)\cdot \lbrace (a+b)}\), \(\mathbf {x \in (0,1))\rbrace \cdot (a_b)^*}\). By PnEMTL semantics, \(\varphi = \mathcal {F}^1_{(0,1)}[(a^*.b.a^*.b.a^*),(a+b)^+](\lbrace a,b\rbrace)\wedge \mathcal {P}^1_{(0,1)}[(b^*.a.b^*.a.b^*)^*.a.b^*),(a+b)^+](\lbrace a,b\rbrace)\) when asserted on a point t will accept the same set of behaviors.□

Skip 8SATISFIABILITY CHECKING FOR NON-ADJACENT PNEMTL Section

8 SATISFIABILITY CHECKING FOR NON-ADJACENT PNEMTL

The main result of the section is as follows:

Theorem 8.1.

Satisfiability Checking for non-adjacent \(\text{PnEMTL}\) and non-adjacent 1-\(\text{TPTL}\) are decidable with EXPSPACE complete complexity. Satisfiability checking for NA-GQMSO is decidable.

The proof is via a satisfiability-preserving reduction to logic \(\text{EMITL}_{0,\infty }\) resulting in a formula whose size is at most exponential in the size of the input non-adjacent \(\text{PnEMTL}\) formula. Satisfiability checking for \(\text{EMITL}_{0,\infty }\) is PSPACE complete [27]. This, along with our construction, implies an EXPSPACE decision procedure for satisfiability checking of non-adjacent \(\text{PnEMTL}\). The EXPSPACE lower bound follows from the EXPSPACE hardness of the sublogic \(\text{MITL}\). The same complexity also applies to non-adjacent 1-\(\text{TPTL}\), using the reduction in the Section 5. This also implies decidability for satisfiability checking of NA-GQMSO formulae, as they can be reduced to equivalent non-adjacent \(\text{PnEMTL}\) formulae (see Lemma 7.3) for which satisfiability could be checked using the following algorithm (reduction to equisatisfiable \(\text{EMITL}_{0,\infty }\) formulae). But the reduction in Lemma 7.3 incurs non-elementary blow-up. Hence, this results in a non-elementary decision procedure. This is to be expected, as lower-bound complexity for satisfiability checking for sublogic FO[\(\lt\)] is non-elementary.

We now describe the technicalities associated with our reduction. We use the technique of equisatisfiability modulo oversampling [31, 35]. Let \(\Sigma\) and \(\mathsf {OVS}\) be disjoint set of propositions. Given any timed word \(\rho\) over \(\Sigma\), we say that a word \(\rho ^{\prime }\) over \(\Sigma \cup \mathsf {OVS}\) is an oversampling of \(\rho\) if \(|\rho | \le |\rho ^{\prime }|\) and when we delete the symbols in \(\mathsf {OVS}\) from \(\rho ^{\prime }\), we get back \(\rho\). Intuitively, \(\mathsf {OVS}\) contains propositions that are used to label oversampling points only. Informally, a formulae \(\alpha\) is equisatisfiable modulo oversampling to formulae \(\beta\) if and only if for every timed word \(\rho\) accepted by \(\beta\) there exists an oversampling of \(\rho\) accepted by \(\alpha\) and, for every timed word \(\rho ^{\prime }\) accepted by \(\alpha\) its projection is accepted by \(\beta\). Note that when \(|\rho ^{\prime }| \gt |\rho |\), \(\rho ^{\prime }\) will have some time points where no proposition from \(\Sigma\) is true. These new points are called oversampling points. Moreover, we say that any point \(i^{\prime } \in dom(\rho ^{\prime })\) is an old point of \(\rho ^{\prime }\) corresponding to i iff \(i^{\prime }\) is the ith point of \(\rho ^{\prime }\) when we remove all the oversampling points. For the rest of this section, let \(\phi\) be a non-adjacent \(\text{PnEMTL}\) formula over \(\Sigma\). We break down the construction of an \(\text{EMITL}_{0,\infty }\) formula \(\psi\) as follows:

(1)

Add oversampling points at every integer timestamp using \(\varphi _{\mathsf {ovs}}\) below.

(2)

Flatten the \(\text{PnEMTL}\) modalities to get rid of nested automata modalities, obtaining an equisatisfiable formula \(\phi _{flat}\).

(3)

With the help of oversampling points, assert the properties expressed by \(\text{PnEMTL}\) subformulae \(\phi _f\) of \(\phi _{flat}\) using only \(\text{EMITL}_{0,\infty } + \mathsf {F}_{np}\) (\(\mathsf {F}_I\) where I is restricted to be non-punctual) modalities getting formula \(\psi _f\). This is done recursively as follows: Using the oversampling points:

(a)

For every \(k\gt 1\) arity \(\text{PnEMTL}\) formula, construct an equivalent formula (for oversampled models) \(\psi ^{k-1}_f\) with arity at most \(k-1\).

(b)

For \(k=1\) arity formula construct an equivalent \(\text{EMITL}_0,\infty + \mathsf {F}_{np}\) modality.

(4)

Finally, in \(\psi _f\), only the \(\mathsf {F}\) operators are timed with intervals of the form \(\langle l, u \rangle\) where \(0\lt l\lt u\lt \infty\). We can reduce these time intervals into purely lower bound (\(\langle l, \infty)\)) or upper bound (\(\langle 0, u \rangle\)) constraints using these oversampling points, by reduction similar to that appearing in Reference [35], Chapter 5, Lemma 5.5.2, pp. 90–91, getting formula of size \(\mathcal {O}(\mathsf {cmax}\times |\psi _{f}|)\).

Let \(\mathsf {Last}=\mathcal {G}\bot\) and \(\mathsf {LastTS}=\mathcal {G}\bot \vee (\bot \mathsf {U}_{(0,\infty)} \top)\). \(\mathsf {Last}\) is true only at the last point of any timed word. Similarly, \(\mathsf {LastTS}\), is true at a point i if there is no next point \(i+1\) with the same timestamp \(\tau _i\). Let max be the maximum constant used in the intervals appearing in \(\phi\). Let \(\mathsf {cmax}= max+1\).

8.1 Behavior of Oversampling Points

We oversample timed words over \(\Sigma\) by adding new points where only propositions from \(\mathsf {Int}\) holds, where \(\mathsf {Int}\cap \Sigma = \emptyset\). Given a timed word \(\rho\) over \(\Sigma\), consider an extension of \(\rho\) called \(\rho ^{\prime }\), by extending the alphabet \(\Sigma\) of \(\rho\) to \(\Sigma ^{\prime } = \Sigma \cup \mathsf {Int}\). Compared to \(\rho\), \(\rho ^{\prime }\) has extra points called oversampling points, where \(\lnot \bigvee \Sigma\) (and \(\bigvee \mathsf {Int}\)) hold. These extra points are added at all integer timestamps in such a way that, if \(\rho\) already has points with integer timestamps, then the oversampled point with the same timestamp appears last among all points with the same timestamp in \(\rho ^{\prime }\). We will make use of these oversampling points to reduce the \(\text{PnEMTL}\) modalities into \(\text{EMITL}_{0,\infty }\). These oversampling points are labelled with a modulo counter \(\mathsf {Int}=\lbrace \mathsf {int}_0,\mathsf {int}_1,\ldots , \mathsf {int}_{\mathsf {cmax}-1}\rbrace\). The counter is initialized to be 0 at the first oversampled point with timestamp 0 and is incremented, modulo \(\mathsf {cmax}\), after exactly one time unit till the last point of \(\rho\). Let \(i \oplus j=(i+ j) \% \mathsf {cmax}\). The oversampled behaviors are expressed using the formula \(\varphi _{\mathsf {ovs}}\): \(\lbrace \lnot \mathsf {F}_{(0,1)} \bigvee \mathsf {Int}\wedge \mathsf {F}_{[0,1)} \mathsf {int}_{0}\rbrace\) \(\wedge\) \(\lbrace \bigwedge \nolimits _{i=0}^{\mathsf {cmax}-1}\mathcal {G}^w\lbrace (\mathsf {int}_i \wedge \mathsf {F}(\bigvee \Sigma))\rightarrow (\lnot \mathsf {F}_{(0,1)} (\bigvee \mathsf {Int}) \wedge \mathsf {F}_{(0,1]} (\mathsf {int}_{i \oplus 1} \wedge (\lnot \bigvee \Sigma) \wedge \mathsf {LastTS}))\rbrace\). to an extension \(\rho ^{\prime }\) given by \(\mathsf {ext}(\rho)=\rho ^{\prime }\) iff (i) \(\rho\) can be obtained from \(\rho ^{\prime }\) by deleting oversampling points and (ii) \(\rho ^{\prime } \models \varphi _{\mathsf {ovs}}\). Map \(\mathsf {ext}\) is well defined as for any \(\rho\), \(\rho ^{\prime }=\mathsf {ext}(\rho)\) if and only if \(\rho ^{\prime }\) can be constructed from \(\rho\) by appending oversampling points at integer timestamps and labelling kth such oversampling point (appearing at time \(k-1\)) with \(\mathsf {int}_{k{\%}\mathsf {cmax}}\).

8.2 Flattening

Next, we flatten \(\phi\) to eliminate the nested \(\mathcal {F}^{k}_{\mathsf {I_1,\ldots , I_k}}\) and \(\mathcal {P}^{k}_{\mathsf {I_1,\ldots , I_k}}\) modalities while preserving satisfiability. Flattening is well studied [27, 31, 35, 40]. The idea is to associate a fresh witness variable \(b_i\) to each subformula \(\phi _i\) that needs to be flattened. This is achieved using the temporal definition, \(T_i=\mathcal {G}^w((\bigvee \Sigma \wedge \phi _i) \leftrightarrow b_i)\), and replacing \(\phi _i\) with \(b_i\) in \(\phi\), \(\phi ^{\prime \prime }_i=\phi [b_i /\phi _i]\), where \(\mathcal {G}^w\) is the weaker form of \(\mathcal {G}\) asserting formula (within its scope) at the current point and all the strict future points. Then, \(\phi ^{\prime }_i=\phi ^{\prime \prime }_i \wedge T_i \wedge \bigvee \Sigma\) is equisatisfiable to \(\phi\). Repeating this across all subformulae of \(\phi\), we obtain \(\phi _{flat}=\phi _t \wedge T\) over the alphabet \(\Sigma ^{\prime }=\Sigma \cup W\), where W is the set of the witness variables, \(T=\bigwedge _i T_i\), \(\phi _t\) is a propositional logic formula over W. Each \(T_i\) is of the form \(\mathcal {G}^w(b_i \leftrightarrow (\phi _f \wedge \bigvee \Sigma)),\) where \(\phi _f=\mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}_1, \ldots , \mathsf {A}_{n+1})(S)\) (or uses \(\mathcal {P}^{n}_{\mathsf {I_1,\ldots , I_n}}\)) and \(S \subseteq \Sigma ^{\prime }\). For example, consider the formula \(\phi =\mathcal {F}^2_{(0,1)(2,3)}(\mathcal {A}_1, \mathcal {A}_2,\mathcal {A}_3)(\lbrace \phi _1, \phi _2\rbrace)\), where \(\phi _1 = \mathcal {P}^2_{(0,2)(3,4)}(A_4,A_5, A_6)(\Sigma),\phi _2 = \mathcal {P}^2_{(1,2)(4,5)}(A_7,A_8, A_9)(\Sigma)\). Replacing the \(\phi _1, \phi _2\) modality with witness propositions \(b_1, b_2\), respectively, we get \(\phi _t=\mathcal {F}^2_{(0,1)(2,3)}(A_1,A_2,A_3)(\lbrace b_1, b_2\rbrace) \wedge T\), where \(T=\mathcal {G}^w(b_1 \leftrightarrow (\bigvee \Sigma \wedge \phi _1)) \wedge \mathcal {G}^w(b_2 \leftrightarrow (\bigvee \Sigma \wedge \phi _2))\), \(A_1,A_2,A_3\) are automata constructed from \(\mathcal {A}_1, \mathcal {A}_2, \mathcal {A}_3\), respectively, by replacing \(\phi _1\) by \(b_1\) and \(\phi _2\) by \(b_2\) in the labels of their transitions. Hence, \(\phi _{flat}=\phi _t \wedge T\) is obtained by flattening the \(\mathcal {F}^{k}_{\mathsf {I_1,\ldots , I_k}},\mathcal {P}^{k}_{\mathsf {I_1,\ldots , I_k}}\) modalities.

8.3 Constructing Equisatisfiable EMITL0,∞ Formula

In this step, for every \(\text{PnEMTL}\) formula \(\phi _f\) appearing in each \(T_i=\mathcal {G}^w(b_i \leftrightarrow (\phi _f \wedge \bigvee \Sigma))\), we will obtain an equisatisfiable \(\text{EMITL}_{0,\infty }\) formula \(\psi _f\). We use oversampling to construct the formula \(\psi _f\) such that for any timed word \(\rho\) over \(\Sigma\), \(i \in dom(\rho)\), there is an extension \(\rho ^{\prime }=\mathsf {ext}(\rho)\) over an extended alphabet \(\Sigma ^{\prime }\), and a point \(i^{\prime } \in dom(\rho ^{\prime })\) that is an old point corresponding to i such that \(\rho ^{\prime }, i^{\prime } \models \psi _f\) iff \(\rho , i \models \phi _f\).

Consider \(\phi _f=\mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}_1, \ldots , \mathsf {A}_{n+1})(S)\) where \(S \subseteq \Sigma ^{\prime }\). Without loss of generality, we assume:

[Assumption 1]: \(\inf (\mathsf {I_1}) \le \inf (\mathsf {I_2}) \le \cdots \le \inf (\mathsf {I_n})\) and \(\sup (\mathsf {I_1}) \le \cdots \le \sup (\mathsf {I_n})\). This is w.l.o.g., since the check for \(\mathsf {A}_{j+1}\) cannot start before the check of \(\mathsf {A}_{j}\) in case of \(\mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}\) modality (and vice versa for \(\mathcal {P}^{n}_{\mathsf {I_1,\ldots , I_n}}\) modality) for any \(1\le j \le n\).

[Assumption 2]: Intervals \(\mathsf {I_1, \ldots I_{n-1}}\) are bounded intervals. Interval \(\mathsf {I_{n}}\) may or may not be bounded. This is also w.l.o.g.11

Let \(\rho =(\sigma _1, \tau _1) \dots (\sigma _n, \tau _n)\in T\Sigma ^*\), \(i \in dom(\rho)\). Let \(\rho ^{\prime }=\mathsf {ext}(\rho)\) be defined by \((\sigma ^{\prime }_1, \tau ^{\prime }_1)\dots (\sigma ^{\prime }_m, \tau ^{\prime }_m)\) with \(m \ge n\), and each \(\tau ^{\prime }_x\) is a either a new integer timestamp not among \(\lbrace \tau _1, \ldots , \tau _n\rbrace\) or is some \(\tau _y\) where x is an old action point corresponding to y. Let \(i^{\prime }\) be an old point in \(\rho ^{\prime }\) corresponding to i. Let \(i^{\prime }_0=i^{\prime }\) and \(i^{\prime }_{n+1} = |\rho ^{\prime }|\). As mentioned above, we make use of these extra action points in \(\rho ^{\prime }\) to assert specification same as \(\phi _f\) without using \(\text{EMITL}_{0,\infty }\) modalities (in case \(\phi _f\) is arity 1 formula) or using \(\text{PnEMTL}\) modality with strictly smaller arity. We first construct a formula \(\phi ^{\prime }_f\) in \(\text{PnEMTL}\) such that \(\rho , i \models \phi _f\) iff \(\rho ^{\prime },i^{\prime } \models \phi ^{\prime }_f\). Note that the satisfaction of \(\phi _f\) is sensitive to these extra action points of \(\rho ^{\prime }\). Hence, \(\rho , i \models \phi _f\) does not guarantee \(\rho ^{\prime }, i^{\prime } \models \phi _f\) unless \(\phi _f\) can be made to ignore oversampling points while checking for satisfaction. We do this as follows: For any \(1\le j \le n+1\), let \(\mathsf {A}^{\prime }_j\) be the automata built from \(\mathsf {A}_j\) by adding self loop on \(\lnot \bigvee \Sigma\) (oversampling points) and \(S^{\prime } = S \cup \lbrace \lnot \bigvee \Sigma \rbrace\). This self loop makes sure that \(\mathsf {A}^{\prime }_j\) ignores (or skips) all the oversampling points while checking for \(\mathsf {A}_j\). Hence, \(\mathsf {A}^{\prime }_j\) allows arbitrary interleaving of oversampling points while checking for \(\mathsf {A}_j\). We call such an NFA as NFA relativized w.r.t. \(\Sigma\). Thus, we have the following proposition:

Proposition 8.2 (Relativization of Automata Modalities).

For any \(g,h \in dom(\rho)\) with \(g^{\prime },h^{\prime }\) being old action points of \(\rho ^{\prime }\) corresponding to \(g,h\), respectively, \(\mathsf {Seg^{s}} (\rho , g, h, S) \in L(\mathsf {A}_i)\) iff \(\mathsf {Seg^{s}} (\rho ^{\prime }, g^{\prime }, h^{\prime }, S\cup \lbrace \lnot \bigvee \Sigma \rbrace) \in L(\mathsf {A}^{\prime }_i)\) for \(s \in \lbrace +,-\rbrace\). Hence, \(\rho , i \models \phi _f\) iff \(\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f\) where \(i^{\prime }\) is an old action point of \(\rho ^{\prime }\) corresponding to i and \(\phi ^{\prime }_f = \mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_{n+1})(S^{\prime })\).

From this point, we will work on eliminating \(\text{PnEMTL}\) modality from \(\phi ^{\prime }_f\) rather than \(\phi _f\), as they are both equisatisfiable (if the former is restricted to be evaluated on models satisfying \(\varphi _{ovs}\), i.e., oversampled models).

We present the reduction by applying induction on arity of the formula. That is, given a \(\text{PnEMTL}\) formula of arity k, we construct a formula of arity at most \(k-1\) such that, for all timed words \(\rho ^{\prime } \models \varphi _{ovs}\), for any old action point \(i^{\prime }\) of \(\rho ^{\prime }\), \(\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f\) iff \(\rho ^{\prime },i^{\prime }\models \phi ^{k-1}_f\) (Recursion Step). In other words, \(\phi ^{\prime }_f \wedge \varphi _{ovs}\) is equivalent to \(\phi ^{k-1}_f \wedge \varphi _{ovs}\). Similarly, if \(\phi ^{\prime }_f\) has arity 1, then we reduce it to an \(\text{EMITL}_{0,\infty }\) formula, \(\psi _f\), such that \(\phi ^{\prime }_f \wedge \varphi _{ovs}\) is equivalent to \(\psi _f \wedge \varphi _{ovs}\) (Base Step). We start with the latter (Base step). That is, we assume that \(\phi ^{\prime }_f = \mathcal {F}_{I_1}(\mathsf {A}^{\prime }_1, \mathsf {A}^{\prime }_2)(S^{\prime })\), we construct a formula \(\psi _f\) such that \(\psi _f\) only contains \(\text{EMITL}_{0,\infty }\) modalities. Before starting with the reduction, we state some useful notations and lemma. For the sake of readability, from this point onward, we do not explicitly mention set of formulae over which the automata modalities are being evaluated unless it is not clear from the context. For example, \(\phi ^{\prime }_f = \mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_{n+1})(S \cup \lnot (\bigvee \Sigma \rbrace)\) will be simply written as \(\mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_{n+1})\).

8.3.1 Notations.

Let \(\mathsf {A}= (Q,q_0, 2^\Sigma , \delta , F)\) be any NFA. For any \(q_1 \in Q\), \(Q_2 \subseteq Q\), \(\mathsf {A}[q_1, Q_2]\) denotes NFA \((Q,q_1, 2^\Sigma , \delta , Q_2)\). For the sake of readability, we abuse this notation by denoting \(\mathsf {A}[q_1, \lbrace q_2\rbrace ]\) as \(\mathsf {A}[q_1,q_2]\) for any \(q_2 \in Q\). \((\mathsf {A})\) denotes the NFA accepting the language that is reverse of \(\mathsf {A}\). Similarly, \(\mathsf {A}\cdot X\) for any set of propositions X denotes an NFA \((Q \cup f, q_0, 2^{\Sigma \cup X}, \delta ^{\prime }, \lbrace f\rbrace)\) where \(\delta ^{\prime } = \delta \cup \lbrace (q,X,f) | q \in F\rbrace\). In other words, \(\mathsf {A}\cdot X\) is an NFA that accepts all the words \(w\cdot X\) where w is accepted by \(\mathsf {A}\). Similarly, for any two automata \(\mathsf {A}\) and \(\mathsf {A}^{\prime }\), \(\mathsf {A}\cdot \mathsf {A}^{\prime }\) denotes NFA constructed by concatenating \(\mathsf {A}\) with \(\mathsf {A}^{\prime }\). Let \(a \notin \Sigma\). We define \(\mathsf {A}^{a} = (Q,q_0, 2^{\Sigma \cup \lbrace a\rbrace }, \delta ^a, F)\) where \(\delta ^a = \lbrace (q,W,q^{\prime }), (q,W\cup \lbrace a\rbrace ,q^{\prime }) | (q,W,q^{\prime }) \in \delta \rbrace\). Hence, for any \(g,h \in dom (\rho)\), \(\mathsf {Seg}^{+/-}(\rho ,g,h,\Sigma) \in L(\mathsf {A}) \iff \mathsf {Seg}^{+/-}(\rho ,g,h,\Sigma \cup \lbrace a\rbrace) \in L(\mathsf {A}^a)\). Hence, \(\mathsf {A}_a\) behaves exactly like \(\mathsf {A}\) irrespective of the occurrence or absence of a at any point. Similarly, we define \(\mathsf {A}^{last, a} = (Q \cup F^a, q_0, 2^{\Sigma \cup \lbrace a\rbrace }, \delta ^{last,a}, F^a)\) where \(F^a = \lbrace (q,1) | q \in F\rbrace\), \(\delta ^{last, a} = \delta ^a \cup \lbrace (q,W,(q^{\prime },1))| q^{\prime } \in F \wedge a \in W \wedge (q,W\setminus \lbrace a\rbrace ,q^{\prime }) \in \delta ^a\rbrace\). In other words, \(L(\mathsf {A}^{last,a}) = L(\mathsf {A}^a) \cap (\bigcup \nolimits _{W \subseteq \Sigma } L((2^{\Sigma })^*\cdot (W\cup \lbrace a\rbrace)))\). Hence, the \(\mathsf {A}^{last,a}\) accepts exactly those words w accepted by \(\mathsf {A}^a\) whose last letter contains proposition a. Note that all these operations result in an NFA that is linear in the size of the input NFA(s).

8.3.2 Lemma for Factoring Regular Languages.

As mentioned above, we fix \(\phi _f = \mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}_1, \ldots , \mathsf {A}_{n+1})\) and \(\phi ^{\prime }_f = \mathcal {F}^{n}_{\mathsf {I_1,\ldots , I_n}}(\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_{n+1}),\) where \(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{n+1}\) are NFA relativized w.r.t. \(\Sigma\). The case of \(\mathcal {P}^{k}_{\mathsf {I_1,\ldots , I_k}}\) modality can be handled symmetrically. We fix \(\rho = (\sigma _1, \tau _1) \ldots (\sigma _m, \tau _m)\) and \({\rho ^{\prime } = (\sigma ^{\prime }_1, \tau ^{\prime }_1) \ldots (\sigma ^{\prime }_{m^{\prime }}, \tau ^{\prime }_{m^{\prime }})=} \mathsf {ext}(\rho)\). Let i be any arbitrary point of \(\rho\) and \(i^{\prime }\) be an old action point of \(\rho ^{\prime }\) corresponding to i. \(i^{\prime }_0=i^{\prime }\). We first present a lemma that reduces the check for condition of the form \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\) by asserting some \(\text{EMITL}_{0,\infty }\) formulae at \(g^{\prime }\) and \(h^{\prime }\) for any \(i \in \lbrace 1,\ldots , n+1\rbrace\). For any \(g^{\prime },h^{\prime } \in dom (\rho)\), let us call segments of the form \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime })\) or \(\mathsf {Seg}^{-}(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime })\) as Segments of \(\rho ^{\prime }\) over \(S^{\prime }\). Let \(i \in \lbrace 1,\ldots ,n+1\rbrace\) be any integer.

Remark: As mentioned above, when \(\mathsf {A}^{\prime }_i\) is evaluated over segments of \(\rho ^{\prime }\) over \(S^{\prime }\), it skips all the oversampling points. But note that the same \(\mathsf {A}^{\prime }_i\) when evaluated over segments of \(S^{\prime } \cup \lbrace \mathsf {int}_j\rbrace\) for some \(0 \le j \lt \mathsf {cmax}\) skips all oversampling points except those labelled with \(\mathsf {int}_j\). This is because the transitions of \(\mathsf {A}^{\prime }_i\) are labelled using subformulae in \(S^{\prime }\) that do not contain any symbols from \(\mathsf {Int}\). Hence, \(\mathsf {A}^{\prime }_i\) has no transition on symbol \(\mathsf {int}_j\). Thus, \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i \cdot \lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)\) iff \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime }-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\), none of the points between \(g^{\prime }\) to \(h^{\prime }-1\) are labelled with \(\mathsf {int}_j\) and point \(h^{\prime }\) is labelled with proposition \(\mathsf {int}_j\). Hence, \(h^{\prime }\) is the first point after \(g^{\prime }\) where \(\mathsf {int}_j\) holds.

Let \(\mathsf {A}^{\prime }_i = (Q_i, init_i, S^{\prime }, \delta , F_i)\).

Lemma 8.3 (Factoring Check for Regular Language).

Let \(g^{\prime }, h^{\prime }\) be any two points of \(\rho ^{\prime }\) such that \(g^{\prime } \lt h^{\prime }\), \(\tau ^{\prime }_{h^{\prime }} - \tau ^{\prime }_{g^{\prime }} \le \mathsf {cmax}\) and \(\lceil \tau ^{\prime }_{g^{\prime }} \rceil \ne \lceil \tau ^{\prime }_{h^{\prime }} \rceil\). Then, \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\) iff \(\bigvee \nolimits _{j=0}^{\mathsf {cmax}-1} \bigvee \nolimits _{q \in Q_i}\bigvee \nolimits _{f \in F_i}[\rho ^{\prime }, g^{\prime } \models \psi ^+(i,init_i,q,j)\wedge \rho ^{\prime }, h^{\prime } \models \psi ^-(i,q,f,j),\) where \(\psi ^+(i,init_i,q,j) = \mathcal {F}(\mathsf {A}_i[init_i,q].\lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)(S^{\prime }\cup \lbrace \mathsf {int}_j\rbrace)\) and \(\psi ^-(i,q,f,j) = \mathcal {P}((\mathsf {A}^{\prime }_i[q,f]).\lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)(S^{\prime }\cup \lbrace \mathsf {int}_j\rbrace)]\).

Proof.

Intuition: We encourage readers to look at Figure 8. As mentioned above, the main purpose of this lemma is to reduce the checking of condition \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\) by asserting some \(\text{EMITL}_{0,\infty }\) formulae at \(g^{\prime }\) and \(h^{\prime }\). As \(\rho ^{\prime }\) satisfies \(\varphi _{\mathsf {ovs}}\) and \(\tau ^{\prime }_{h^{\prime }} - \tau ^{\prime }_{g^{\prime }} \le \mathsf {cmax}\), all the oversampling integer points between \(g^{\prime }\) and \(h^{\prime }\) are labelled with unique counters, as the counters increment at every oversampling point modulo \(\mathsf {cmax}\). Hence, if the oversampling integer time point immediately after \(g^{\prime }\) is labelled \(\mathsf {int}_j\), then no other point between \(g^{\prime }\) and \(h^{\prime }\) is labelled \(\mathsf {int}_j\). Moreover, the oversampling integer time point immediately after \(g^{\prime }\) (say, c) is labelled with a proposition \(\mathsf {int}_j\) iff \(\lceil \tau ^{\prime }_{g^{\prime }} \rceil \% \mathsf {cmax}= j\). For checking \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\), we make use of this oversampling point c to split the run(s) as follows:

Fig. 8.

Fig. 8. Figure for Lemma 8.3. For any word \(\rho ^{\prime }\) satisfying \(\varphi _{ovs}\) , checking whether pattern of satisfaction of subformulae in \(S^{\prime }\) between points \(g^{\prime }\) and \(h^{\prime }\) is accepted by \(\mathsf {A}^{\prime }_i\) can be reduced to asserting untimed \(\text{EMITL}\) formulae at \(g^{\prime }\) and \(h^{\prime }\) . The behavior from \(g^{\prime }\) to c is given by \(\mathcal {F}(\mathsf {A}_i[init,q].\lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)\) for some \(q \in Q_i\) and the corresponding behavior from c to \(h^{\prime }\) is given by \(\mathcal {P}((\mathsf {A}^{\prime }_k[q,f])\cdot \lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)\) for some final state \(f \in F_i\) . Disjuncting over all possible (but finitely many) \(j \in \lbrace 0, \mathsf {cmax}-1\rbrace\) , \(q \in Q_i\) and \(f \in F_i\) , we get the required formulae.

(1)

Checking the First Part: Concretely, checking for \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i)\), we start at \(g^{\prime }\) in \(\rho ^{\prime }\), from the initial state \(init_i\) of \(\mathsf {A}_i\), and move to the state (say, q) that is reached at the closest oversampling point c. Note that we use only \(\mathsf {A}_i\) (without any \(\lnot \bigvee \Sigma\) self loops) to disallow occurrence of any oversampling point except at the last point. This ensures that we end our run after reading the closest oversampling point c.

(2)

Checking the Latter Part: Reaching q from \(init_i\), we have read a partial behavior between \(g^{\prime }\) and c; this must be extended to check full behavior by starting from state q, continuing from point c, with transition rules of \(\mathsf {A}^{\prime }_i\) and assert that we end at an accepting state after reading the point \(h^{\prime }\). Note that we use \(\mathsf {A}^{\prime }_i\) instead of \(\mathsf {A}_i\) (used in the first part) to ignore the oversampling points that could be encountered while checking the latter part, i.e., from c to \(h^{\prime }\)). Hence, starting from \(g^{\prime }\) with initial state of \(\mathsf {A}^{\prime }_i\), we reach at the accepting state of \(\mathsf {A}^{\prime }_i\) after reading point \(h^{\prime }\) iff we end at some state q after the end of checking the first part while simulating \(\mathsf {A}_i\), after which on simulating \(\mathsf {A}^{\prime }_i\) and continuing from state q, we reach some accepting state of \(\mathsf {A}^{\prime }_i\) on reading till \(h^{\prime }\) and hence ending the check for the second part.

Note that check for the first part ending at some state q can be characterized by \(\rho ^{\prime }, g^{\prime } \models \mathcal {F}(\mathsf {A}_i[init,q].\lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)\). For reducing the check of latter part with a formula asserting at \(h^{\prime }\), we start the check for automata in reverse. That is, we assert that: Starting from some final state f from \(h^{\prime }\), if we simulate the \(\mathsf {A}^{\prime }_i\) in reverse direction till point c, then we should be able to reach q. Note that the end point of the segment in \(\text{EMITL}_{0,\infty }\) formula is within an existential quantifier. Then, how do we make sure that we end our check at c? This can be done by asserting that the check ends at the nearest point before \(h^{\prime }\) where \(\mathsf {int}_j\) holds first. As c is the only point between \(g^{\prime }\) and \(h^{\prime }\) where \(\mathsf {int}_j\) holds, we are sure to end at c. Hence, checking for latter part is equivalent to check \(\rho ^{\prime },h^{\prime } \models \mathcal {P}((\mathsf {A}^{\prime }_i[q,f]).\lbrace \mathsf {int}_j, \lnot \bigvee \Sigma \rbrace)(S^{\prime }\cup \lbrace \mathsf {int}_j\rbrace)\). Before starting the proof, we give a very simple example that gives some intuition about the construction.

Example 8.4.

Consider the formula \(\phi = \mathcal {F}^2_{(1,2),(3,4)}(Even_a, b^*,\Sigma ^*)\), where \(Even_a\) is an automaton accepting strings containing even number of \(a^{\prime }s\). \(\rho , i\) satisfies \(\phi\) if and only if there exist points \(i_1\) (within \((1,2)\) of i) and \(i_2\) (within \((3,4)\) of i) such that there is an even number of \(a^{\prime }s\) between i and \(i_1\) and only \(b^{\prime }s\) occur between \(i_1\) and \(i_2\). Observe the following Figure 9. Consider \(\rho ^{\prime } = \mathsf {ext}(\rho)\). Let \(i^{\prime }, i^{\prime }_1, i^{\prime }_2\) be the points of \(\rho ^{\prime }\) corresponding to old action points \(i, i_1, i_2\) of \(\rho\), respectively. Now, we can break the check between \(i^{\prime }\) and \(i_1^{\prime }\) at the smallest integer oversampling point occurring after \(i^{\prime }\) labelled \(\mathsf {int}_j\). The number of \(a^{\prime }s\) between \(i^{\prime }\) and \(i^{\prime }_1\) (and hence number of \(a^{\prime }s\) between i and \(i_1\)) is even iff the number of \(a^{\prime }s\) between \(i^{\prime }\) and next occurring \(\mathsf {int}_j\) and the number of \(a^{\prime }s\) between \(\mathsf {int}_j\) and \(i_1^{\prime }\) are either both odd or both even. Similarly, all points between \(i_1\) and \(i_2\) are labelled b iff at all old action points between \(i^{\prime }_1\) and nearest point x labelled \(\mathsf {int}_{j^{\prime }}\) (where \(j^{\prime } \in \lbrace 0,\ldots ,\mathsf {cmax}-1\rbrace)\), and continuing from x to \(i_2\) only b or \(\mathsf {int}_{j^{\prime \prime }}\) occurs where \(j^{\prime \prime } \ne j^{\prime }\). Hence, formula \(\phi\) is satisfiable iff the following formula \(\psi\) is satisfiable:

Fig. 9.

Fig. 9. Figure depicting the construction of equisatisfiable EMITL formula from non-adjacent PnEMTL formula. TS stands for timestamp. Hence, \(\tau\) is a timestamp of point \(i_0\) . At the top, we have timed word \(\rho\) and the bottom part of the figure denotes \(\rho ^{\prime } = \mathsf {ext}(\rho)\) .

\(\psi = \varphi _{ovs} \wedge \bigvee \nolimits _{0\le j,j^{\prime } \le \mathsf {cmax}-1,j^{\prime \prime } \ne j^{\prime }} [\mathsf {F}_{(0,1]}(\mathsf {int}_{j}) \rightarrow [\lbrace \mathsf {F}_{(3,4)}(\mathcal {P}((b+\mathsf {int}_{j^{\prime \prime }})^*.\mathsf {int}{j^{\prime }})\rbrace \wedge \lbrace \lbrace \mathcal {F}(Even_a.\mathsf {int}_j) \wedge \mathsf {F}_{(1,2)}(\mathcal {P}(Even_a.\mathsf {int}_j) \wedge \mathcal {F}(b^*.\mathsf {int}_{j\oplus 2})\rbrace \vee \lbrace \mathcal {F}(Odd_a.\mathsf {int}_j) \wedge \mathsf {F}_{(1,2)}(\mathcal {P}(Odd_a.\mathsf {int}_j) \wedge \mathcal {F}(b^*.\mathsf {int}_{j^{\prime }})))\rbrace \rbrace ]]\).□

Formal Proof: We argue for correctness as follows:

(1)

Let \(c^{\prime }\) be any point between \(g^{\prime }\) and \(h^{\prime }\). Then, any accepting run from point \(g^{\prime }\) to \(h^{\prime }\) ending at an accepting state \(f \in F\) will pass through point \(c^{\prime }\) such that after reading the point \(c^{\prime }\) the run ends up at some state q. Thus, the behavior from \(g^{\prime }\) to \(c^{\prime }\) is given by all the runs starting from init and ending at state q (hence in \(\mathsf {A}^{\prime }_i[init, q]\)). Similarly, the remaining part of the run from \(c^{\prime }+1\) to \(h^{\prime }\) is characterized by those continuing from q to f (hence, in \(\mathsf {A}^{\prime }_i(q, f)\)). Disjuncting over all possible values of \(q \in Q\) and \(f \in F\), we get all the possible accepting runs. Hence, \(\forall c^{\prime }. g^{\prime }\lt c^{\prime }\lt h^{\prime }\), we have (1) \(\begin{equation} \begin{aligned} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff &\bigvee \limits _{q \in Q_i}\bigvee \limits _{f \in F_i}&\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i[init, q]) \\ &&\wedge \mathsf {Seg}^+(\rho ^{\prime },c^{\prime }+1,h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]). \end{aligned} \end{equation}\)

(2)

As \(\lceil \tau ^{\prime }_{g^{\prime }} \rceil \ne \lceil \tau ^{\prime }_{h^{\prime }} \rceil\) and \(g^{\prime } \lt h^{\prime }\), \(\tau ^{\prime }_{h^{\prime }} \gt \lceil \tau ^{\prime }_{g^{\prime }} \rceil\). Moreover, as \(\rho ^{\prime } \models \varphi _{\mathsf {ovs}}\), there is an oversampling point c with timestamp \(\tau ^{\prime }_c = \lceil \tau ^{\prime }_{g^{\prime }} \rceil\) where \(\mathsf {int}_j\) holds. Hence, by Equation (1) and as \(g^{\prime } \le c \le h^{\prime }\), we have (2) \(\begin{equation} \begin{aligned} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff &\bigvee \limits _{q \in Q_i}\bigvee \limits _{f \in F_i}&\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c,S^{\prime }\cup) \in L(\mathsf {A}^{\prime }_i[init, q]) \\ &&\wedge \mathsf {Seg}^+(\rho ^{\prime },c+1,h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]). \end{aligned} \end{equation}\)

(3)

As \(\mathsf {A}^{\prime }_i\) has a self loop over \(\lnot \bigvee \Sigma\), the states do not change on reading (or not reading) the oversampling point c. Hence, \(\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,Q_f]) \iff \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,Q_f])\). This implies: (3) \(\begin{equation} \begin{aligned} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff & \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F}&\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[init, q]) \\ &&\wedge \mathsf {Seg}^+(\rho ^{\prime },c+1,h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]). \end{aligned} \end{equation}\)

(4)

By definition of \(\mathsf {Seg}^+\) and \(\mathsf {Seg}^-\), \(\mathsf {Seg}^+(\rho ^{\prime },c+1,h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]) \iff \mathsf {Seg}^{-}(\rho ^{\prime },h^{\prime },c+1,S^{\prime }) \in L((\mathsf {A}^{\prime }_i[f,q]))\). This, along with Equation (3), implies, (4) \(\begin{equation} \begin{aligned} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff &\bigvee \limits _{q \in Q_i}\bigvee \limits _{f \in F_i}&[\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[init_i, q]) \\ &&\wedge \mathsf {Seg}^{-}(\rho ^{\prime },h^{\prime },c+1,S^{\prime }) \in L((\mathsf {A}^{\prime }_i[q,f])]. \end{aligned} \end{equation}\)

(5)

As \(\rho ^{\prime } \models \varphi _{\mathsf {ovs}}\). If \(j = \lceil \tau ^{\prime }_{g^{\prime }} \rceil \%\mathsf {cmax}\), then point c is labelled with \(\mathsf {int}_j\). Moreover, there is no oversampling point between \(g^{\prime }\) and c. Hence, (5) \(\begin{equation} \begin{aligned} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[init_i,q]) \iff \exists g^{\prime }\lt c^{\prime }. \\ \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c^{\prime },S^{\prime } \cup \lbrace \mathsf {int}_j\rbrace) \in L\left(\mathsf {A}_i[init_i, q]\cdot \lbrace \mathsf {int}_j, \bigvee \Sigma \rbrace \right)\!. \end{aligned} \end{equation}\) Note that we use \(\mathsf {A}_i\) instead of \(\mathsf {A}^{\prime }_i\), as \(\mathsf {A}_i\) will make sure that in the initial part from \(g^{\prime }\) to \(c^{\prime }-1\) there is no oversampling point, as it has no self loops on \(\lnot \bigvee \Sigma\). This will ensure that the \(c^{\prime }\) point is the very first oversampling point after \(g^{\prime }\). Hence, there is only one choice for \(c^{\prime }\), i.e., c. Moreover, concatenating \(\lbrace \mathsf {int}_j, \bigvee \Sigma \rbrace\) at the end makes sure that c is labelled as \(\mathsf {int}_j\).

(6)

Similarly, (6) \(\begin{equation} \begin{aligned} \mathsf {Seg}^{-}(\rho ^{\prime },h^{\prime },c+1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]) \iff \exists c^{\prime }\lt h^{\prime }. \\ \mathsf {Seg}^{-}(\rho ^{\prime },h^{\prime },c^{\prime },S^{\prime } \cup \lbrace \mathsf {int}_j\rbrace) \in L\left(\left(\mathsf {A}^{\prime }_i[q, f]\right)\cdot \left\lbrace \mathsf {int}_j, \bigvee \Sigma \right\rbrace \right)\!. \end{aligned} \end{equation}\) As \(\mathsf {A}^{\prime }_i\) does not contain symbols from \(\mathsf {Int}\), \(c^{\prime }\) is the nearest such point before \(h^{\prime }\) where \(\mathsf {int}_j\) holds. As \(\tau ^{\prime }_{h^{\prime }}\) - \(\tau ^{\prime }_{g^{\prime }} \le \mathsf {cmax}\) and the counters are incremented modulo \(\mathsf {cmax}\) at integer timestamps by \(\varphi _{ovs}\), if c is labelled as \(\mathsf {int}_j\), then there is no other point between \(g^{\prime }\) and \(h^{\prime }\) that will be labelled \(\mathsf {int}_j\). Hence, there is only one choice for \(c^{\prime }\), i.e., c.

(7)

By semantics of \(\mathcal {F}\) and Equation (5), we have: (7) \(\begin{equation} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },c-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[init_i,q]) \iff \rho ^{\prime }, g^{\prime } \models \mathcal {F}\left(\mathsf {A}_i[init_i, q]\cdot \left\lbrace \mathsf {int}_j, \bigvee \Sigma \right\rbrace \right)\!. \end{equation}\) Similarly, by semantics of \(\mathcal {P}\) and Equation (6), we have: (8) \(\begin{equation} \mathsf {Seg}^{-}(\rho ^{\prime },h^{\prime },c+1,S^{\prime }) \in L(\mathsf {A}^{\prime }_i[q,f]) \iff \rho ^{\prime },h^{\prime } \models \mathcal {P}\left(\left(\mathsf {A}_i[q, f]\right)\cdot \left\lbrace \mathsf {int}_j, \bigvee \Sigma \right\rbrace \right)\!. \end{equation}\)

(8)

By Equations (4), (7), and (8) and disjuncting over all possible values of \(q\in Q_i\) and \(f\in F_i\), if \(j = \lceil \tau ^{\prime }_{g^{\prime }} \rceil \%\mathsf {cmax}\), we have: (9) \(\begin{equation} \mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F} [\rho ^{\prime }, g^{\prime } \models \psi ^+(i,init_i,q,j) \wedge \rho ^{\prime }, h^{\prime } \models \psi ^-(i,q,f,j). \end{equation}\) Finally, disjuncting over all possible values of \(j \in \lbrace 0,\ldots ,\mathsf {cmax}-1\rbrace ,\) we have the required result: (10) \(\begin{equation} \begin{aligned}\mathsf {Seg}^+(\rho ^{\prime },g^{\prime },h^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_i) \iff \bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F} [\rho ^{\prime }, g^{\prime } \models \psi ^+(i,init_i,q,j) \wedge \rho ^{\prime }, h^{\prime } \models \psi ^-(i,q,f,j)]. \end{aligned} \end{equation}\)

We now start with the base case of the construction.

Lemma 8.5 (Base Lemma).

If \(\phi ^{\prime }_f = \mathcal {F}^1_{\langle l, u \rangle }(\mathsf {A}^{\prime }_1, \mathsf {A}^{\prime }_2)(S^{\prime })\)(i.e., \(n=1\)), then we can construct an \(\text{EMITL}_{0,\infty }\) formula \(\psi _f\) such that \(\rho ^{\prime }, i^{\prime }\) satisfies \(\phi ^{\prime }_f\) if and only if it satisfies \(\psi _f\). Moreover, the total number of operators (temporal and Boolean), \(N= \mathcal {O}(\mathsf {cmax}\times |\mathsf {A}_1| \times |\mathsf {A}_1| \times |\phi _f|)\).

Proof.

To reiterate the semantics of \(\phi ^{\prime }_f\): (11) \(\begin{equation} \rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f \iff \exists i_1^{\prime }. \tau _{i^{\prime }_1} - \tau _{i^{\prime }} \in \langle l, u \rangle \wedge \mathsf {Seg}^+(\rho ^{\prime },i_0^{\prime },i_1^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_1) \wedge \mathsf {Seg}^+(\rho ^{\prime },i^{\prime }_1,m^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_2). \end{equation}\)

Case 1: \(l=0\) or \(u = \infty\).

Intuition: This case is straightforward. As we have only \(\text{PnEMTL}\) modality with unit arity and the intervals are either of the form \(\langle 0,u\rangle\) or \(\langle l, \infty \rangle\), we can use an \(\mathcal {F}_{\langle l,u \rangle }\) \(\text{EMITL}_{0,\infty }\) formula to assert the check for \(\mathsf {A}^{\prime }_1\), which has a nested untimed \(\text{EMITL}\) formulae to check \(\mathsf {A}^{\prime }_2\).

Formal Construction and proof: In this case, we can trivially reduce \(\phi ^{\prime }_f\) into an equivalent \(\psi _f\) that is already in \(\text{EMITL}_{0,\infty }\) using nesting: \(\psi _f = \mathcal {F}_{\langle l, u \rangle }({\mathsf {A}^{\prime }_{1}}^{\mathsf {last}, \beta })\) where \(\beta = \mathcal {F}(\mathsf {A}^{\prime }_2\cdot {\mathsf {Last}}) (S^{\prime }\cup \lbrace \mathsf {Last}\rbrace)\). By semantics of \(\mathcal {F}\) modality, (12) \(\begin{equation} \rho ^{\prime },i^{\prime } \models \psi _f \iff \exists i_1^{\prime }. \tau _{i^{\prime }_1} - \tau _{i^{\prime }} \in \langle l, u \rangle \wedge \mathsf {Seg}^+(\rho ^{\prime },i^{\prime },i_1^{\prime },S^{\prime }\cup \lbrace \beta , \mathsf {last}\rbrace) \in L({\mathsf {A}^{\prime }_{1}}^{\mathsf {last}, \beta }). \end{equation}\) Moreover, by definition of \(\mathsf {A}^{last, a}\), (13) \(\begin{equation} \rho ^{\prime },i^{\prime } \models \psi _f \iff \tau _{i^{\prime }_1} - \tau _{i^{\prime }} \in \langle l, u \rangle \wedge \mathsf {Seg}^+(\rho ^{\prime },i^{\prime },i_1^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_1) \wedge \rho ^{\prime }, i^{\prime }_1 \models \beta . \end{equation}\) Note that (14) \(\begin{equation} \rho ^{\prime }, i^{\prime }_1 \models \beta \iff \mathsf {Seg}^+(\rho ^{\prime },i^{\prime }_1,m^{\prime }-1,S^{\prime }) \in L(\mathsf {A}^{\prime }_2)\ \text{(by semantics)} \iff \mathsf {Seg}^+(\rho ^{\prime },i^{\prime }_1,m^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_2). \end{equation}\) The equivalence in the right is due to the observation the last point is always an oversampling point, as \(\rho ^{\prime } \models \varphi _{\mathsf {ovs}}\), and \(\mathsf {A}^{\prime }_2\) loops over oversampling points when evaluated on segments over \(S^{\prime }\) (hence, the set of states reached at \(m^{\prime }-1\) and m are the same). By Equations (14) and (13), we get \(\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f \iff \rho ^{\prime }, i^{\prime } \models \psi _f\), where \(\psi _f\) is an \(\text{EMITL}_{0,\infty }\) formula. Note that in this case \(|\psi _f| = \mathcal {O}(|\phi _f|).\)

Case 2: \(\langle l, u \rangle\) is a bounded interval where \(l \gt 0\). Hence, \(\mathsf {cmax}\ge \tau ^{\prime }_{i^{\prime }_1} - \tau ^{\prime }_{i^{\prime }} \ge 1\). As \(\tau ^{\prime }_{i^{\prime }_1} - \tau ^{\prime }_{i^{\prime }} \ge 1\) implies \(\lceil \tau ^{\prime }_{i^{\prime }_{1}} \rceil \ne \lceil \tau ^{\prime }_{i^{\prime }} \rceil\), we can apply the Lemma 8.3. Let \(\mathsf {A}_1 = (Q,init, 2^{S^{\prime }}, \delta , F)\).

Intuition: In this case, to check \(\mathsf {Seg}^+(\rho ^{\prime },i_0^{\prime },i_1^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_1)\), we use Lemma 8.3, which gives us \(\text{EMITL}_{0,\infty }\) \(\mathcal {F}\) formulae (of the form \(\psi ^+\), as mentioned in Lemma 8.3) to be asserted at \(i^{\prime }\) and \(\mathcal {P}\) formulae of the form \(\psi ^-\) to be asserted at \(i_1^{\prime }\). The former can be asserted directly, as \(i^{\prime }\) is the present point. For asserting formulae at \(i^{\prime }_1\), we jump from \(i^{\prime }\) to \(i^{\prime }_1\) using \(\mathsf {F}_{\langle l, u \rangle }\) modality and assert the corresponding \(\mathcal {P}\) modality. For checking \(\mathsf {A}^{\prime }_2\), we assert formulae \(\beta\), as constructed in the previous case at \(i_1^{\prime }\).

Formal Construction and Proof: (15) \(\begin{equation} \mathsf {Seg}^+(\rho ^{\prime },i_0^{\prime },i_1^{\prime },S^{\prime }) \in L(\mathsf {A}^{\prime }_1) \iff \bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F}[\rho ^{\prime }, g^{\prime } \models \psi ^+(1,init,q,j) \wedge \rho ^{\prime }, h^{\prime } \models \psi ^-(1,q,f,j)] \end{equation}\) Using Equations (15) and (14) in Equation (11), we get: (16) \(\begin{equation} \begin{aligned}\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f \iff & \\ \exists i_1^{\prime }. \tau _{i^{\prime }_1} - \tau _{i^{\prime }} \in \langle l, u \rangle \wedge & \bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F}[\rho ^{\prime }, i_0^{\prime } \models \psi ^+(1,init,q,j) \wedge \rho ^{\prime }, i_1^{\prime } \models \psi ^-(1,q,f,j)] \wedge \rho ^{\prime }, i_1^{\prime } \models \beta , \end{aligned} \end{equation}\) where \(\beta\) is the same as one used in Equation (14). We can eliminate the quantifier guarded by the timing constraint \(\exists i_1^{\prime }. \tau _{i^{\prime }_1} - \tau _{i^{\prime }} \in \langle l, u \rangle\) using \(\mathsf {F}_{\langle l, u \rangle }\) modality. Hence, by semantics of \(\mathsf {F}_{\langle l, u \rangle }\) modality, if \(\psi _f = \bigvee \nolimits _{j=0}^{\mathsf {cmax}-1} \bigvee \nolimits _{q \in Q}\bigvee \nolimits _{f \in F}\psi ^+(1,init,q,j) \wedge \mathsf {F}_{I_1}(\psi ^-(1,q,f,j) \wedge \beta ],\) then by semantics of \(\rho ^{\prime },i^{\prime } \models \phi ^{\prime }_f \iff \rho ^{\prime }, i^{\prime } \models \psi _f\). Note that \(\psi _f\) contains only \(\text{EMITL}_{0,\infty }\) and \(\mathsf {F}_{np}\) modalities and hence is the required formulae. Also note that, as each \(\psi ^+(1,init,q,j)\) and \(\psi ^-(1,q,f,j)\) formulae are of the size of \(\mathcal {O}(\phi _f)\), we have \(|\psi _f| = \mathcal {O}(\mathsf {cmax}\times |Q_1|\) \(\times\) \(|F_1| \times |\phi _f|)\).

We now give the recursive reduction. We show that, given any non-adjacent \(\text{PnEMTL}\) formula of arity k, we can construct an equisatisfiable formulae non-adjacent \(\text{PnEMTL}\) formula of arity \(k-1\) or less.

Lemma 8.6 (Recursive Reduction Lemma).

If \(\phi ^{\prime }_f = \mathcal {F}^{k}_{I_1,\ldots I_{k}}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k+1})\), then we can construct an \(\text{PnEMTL}\) formula \(\psi ^{k-1}_f\) with arity at most \(k-1\) such that \(\rho ^{\prime }, i^{\prime }\) satisfies \(\phi ^{\prime }_f\) if and only if it satisfies \(\psi ^{k-1}_f\). Moreover, the total number of operators \(N = \mathcal {O}(\mathsf {cmax}\times |\mathsf {A}_k| \times |\mathsf {A}_k| \times |\phi _f|)\).

Proof.

To rephrase the semantics of \(\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_f\) (by pushing \(\exists i^{\prime }_{k-1}\le i^{\prime }_{k}. (\tau ^{\prime }_{i^{\prime }_k} -\tau ^{\prime }_{i^{\prime }} \in \mathsf {I}_k\) inside): (17) \(\begin{equation} \begin{aligned}\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_k \!{\iff }\!\!&\exists i^{\prime }\le i_1^{\prime } \le \cdots \le i^{\prime }_{k-1}. \bigwedge \limits _{g=1}^{k-1} (\tau ^{\prime }_{i^{\prime }_g} -\tau ^{\prime }_{i^{\prime }} \in \mathsf {I}_g {\wedge }\rho ^{\prime },i^{\prime }_g \models \bigvee \Sigma \wedge \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{g-1}, i^{\prime }_g, S^{\prime }) \in L(\mathsf {A}^{\prime }_g)) \\ & \wedge \exists i^{\prime }_{k-1}\le i^{\prime }_{k}. (\tau ^{\prime }_{i^{\prime }_k} -\tau ^{\prime }_{i^{\prime }} \in \mathsf {I}_k \!\wedge \! \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, i^{\prime }_{k}, S^{\prime }) \in L(\mathsf {A}^{\prime }_{k}) \!\wedge \! \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{k}, m^{\prime }, S^{\prime }) \in L(\mathsf {A}^{\prime }_{k+1}). \end{aligned} \end{equation}\) Let \(\mathsf {A}^{\prime }_k = (Q,\mathsf {init}, 2^{S^{\prime }}, \delta , F)\), \(I_{k-1} = \langle l_{k-1}, u_{k-1} \rangle\) and \(I_{k} = \langle l_k, u_k \rangle\).

Case 1: \(I_{k-1}\) and \(I_{k}\) are non-overlapping. That is, \(u_{k-1} \lt l_k\) (strict \(\lt\) is implied by the fact that the set of intervals \(\lbrace \mathsf {I_1, \ldots , I_k}\rbrace\) is non-adjacent). Hence, \(\tau ^{\prime }_{i^{\prime }_{k-1}} -\tau ^{\prime }_{i^{\prime }_k} \ge 1\) for any possible value of \(i^{\prime }_{k-1}\) and \(i^{\prime }_k\).

Case 1.1: \(I_{k}\) is bounded, i.e., \(u_k \ne \infty\). Then, \(\tau ^{\prime }_{i^{\prime }_{k-1}} -\tau ^{\prime }_{i^{\prime }_k} \le \mathsf {cmax}\) holds.

Intuition: Refer to Figure 10. Similar to case 2 of Lemma 8.5, we apply Lemma 8.3 to split the check for \(\mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, i^{\prime }_{k}, S^{\prime }) \in L(\mathsf {A}^{\prime }_{k})\) at the nearest oversampling point c after \(i^{\prime }_{k-1}\). The first part of the check (from \(i^{\prime }_{k-1}\) to c) can be asserted using the kth tail automata of \(k-1\)-ary \(\text{PnEMTL}\) formula, where the first \(k-2\) arguments are identical to that of \(\phi ^{\prime }_f\). The second part of the check (from c to \(i^{\prime }_k\)) can be asserted in the reverse direction from \(i^{\prime }_k\) by jumping to it from \(i^{\prime }\) using \(\mathsf {F}_{I_k}\) modality.

Fig. 10.

Fig. 10. Figure showing reduction of k-ary \(\text{PnEMTL}\) formulae to \(k-1\) -ary \(\text{PnEMTL}\) formulae when \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\) are not within same integer time points (Case 1.1). The behavior from \(i^{\prime }\) to nearest oversampling point c after \(i^{\prime }_{k-1}\) (labelled \(\mathsf {int}_j\) ) is given by \(\mathcal {F}_{I_1, \ldots , I_{k-1}} (\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}_k[init,q]\cdot \mathsf {int}_j.\Sigma ^*),\) and the corresponding behavior from c is given by \(\lbrace \mathsf {F}_{I_k}(\mathcal {P}((\mathsf {A}^{\prime }_k[q,f])\cdot \mathsf {int}_j)) \wedge \mathcal {F}(\mathsf {A}_{k+1}\cdot \mathsf {Last})\rbrace\) , where q is any state of \(\mathsf {A}^{\prime }_{k}\) reached when read till \(\mathsf {int}_j\) and f is the final state that is reached when \(i^{\prime }_k\) is read. Disjuncting over all possible (but finitely many) \(j \in \lbrace 0,\ldots , \mathsf {cmax}-1\rbrace\) , \(q \in Q_i\) and \(f \in F_i\) , we get the required formulae.

Construction: By Lemma 8.3, (18) \(\begin{equation} \begin{aligned} \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, i^{\prime }_{k}, S^{\prime }) \in L(\mathsf {A}^{\prime }_{k}) & \equiv & \bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F}\! \rho ^{\prime }, i^{\prime }_{k-1} \models \exists c^{\prime }. {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, c^{\prime }, S^{\prime })& \\ && \in L(\mathsf {A}_k[init, q] \cdot \mathsf {int}_j) \wedge \rho ^{\prime }, i^{\prime }_{k} \models \psi ^-(k,q,f,j)&. \end{aligned} \end{equation}\) For the sake of brevity, we make following abuses of notation: For any NFA \(\mathsf {A}\) and any proposition \(\mathsf {int}_j \in \mathsf {Int}\), \(\mathsf {A}\cdot \lbrace nt_j, \lnot \bigvee \Sigma \rbrace\) is denoted by \(\mathsf {A}\cdot \mathsf {int}_j\). Moreover, automata accepting all possible behaviors over any given set of subformulae is denoted by \(\Sigma ^*\). Moreover, (19) \(\begin{equation} \exists c^{\prime }. {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, c, S^{\prime }) \in L(\mathsf {A}_k[init, q]\cdot \mathsf {int}_j) \iff {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, m^{\prime }, S^{\prime }) \in L(\mathsf {A}_k[init, q]\cdot \mathsf {int}_j\cdot \Sigma ^*). \end{equation}\) Hence, by Equations (17), (18), and (19), we have \(\rho ^{\prime }, i^{\prime } \models \phi ^{\prime }_k \iff\) (20) \(\begin{equation} \begin{aligned}\rho ^{\prime },i^{\prime } \models &\bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F} \left[\left\lbrace \exists i^{\prime }\le i_1^{\prime } \le \cdots \le i^{\prime }_{k-1}. \bigwedge \limits _{g=1}^{k-1} (\tau ^{\prime }_{i^{\prime }_g} -\tau ^{\prime }_{i^{\prime }}) \in \mathsf {I}_g \wedge \rho ^{\prime },i^{\prime }_g \models \bigvee \Sigma \right.\right.\\ & \left.\wedge \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{g-1}, i^{\prime }_g, S^{\prime }) \in L(\mathsf {A}^{\prime }_g)) \wedge {Seg^+}(\rho ^{\prime }, i^{\prime }_{k-1}, m^{\prime }, S^{\prime }) \in L(\mathsf {A}_k[init, q]\cdot \mathsf {int}_j\cdot \Sigma ^*)\right\rbrace \\ &\left.\wedge \left\lbrace \exists i^{\prime }_{k}. (\tau ^{\prime }_{i^{\prime }_k} -\tau ^{\prime }_{i^{\prime }}) \in \mathsf {I}_k. \rho ^{\prime }, i^{\prime }_k \models \psi ^-(k,q,f,j) \wedge \mathsf {Seg^+}(\rho ^{\prime }, i^{\prime }_{k}, m^{\prime }, S^{\prime }) \in L(\mathsf {A}^{\prime }_{k+1})\right\rbrace \right]\!. \end{aligned} \end{equation}\) By semantics of \(\text{PnEMTL}\) and \(\text{EMITL}\) logic, the above condition is equivalent to \(\rho ^{\prime }, i^{\prime } \models \psi ^{1.1}_f=\) (21) \(\begin{equation} \bigvee \limits _{j=0}^{\mathsf {cmax}-1} \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F}[\lbrace \mathcal {F}^{k-1}_{\mathsf {I_1, \dots ,I_{k-1}}}(\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_{k-1}, \mathsf {A}_k[init, q]\cdot \mathsf {int}_j \cdot \Sigma ^*)\rbrace \wedge \lbrace \mathsf {F}_{I_k} (\psi ^-(k,q,f,j) \wedge \mathcal {F}(\mathsf {A}^{\prime }_{k+1}\cdot \mathsf {Last}))\rbrace ]. \end{equation}\)

Case 1.2: \(I_{k}\) is unbounded, i.e., \(u = \infty\). There are two possibilities. Either (1) \(i^{\prime }_{k}\) occurs within \(\langle l_k, l_k+1)\) from \(i^{\prime }\) or (2) \(i^{\prime }_k\) occurs beyond timestamp \(\lceil \tau ^{\prime }_{i^{\prime }}+l_k \rceil\):

Possibility 1: \(i^{\prime }_{k}\) occurs within \(\langle l_k, l_k+1)\). Hence, \(\tau ^{\prime }_{i^{\prime }_{k}} \in \langle \tau ^{\prime }_{i^{\prime }} + l_{k}, \tau ^{\prime }_{i^{\prime }} + l_k+1)\). The case can be handled using the following formulae: (22) \(\begin{equation} \phi ^{1.2.1} = \mathcal {F}^{k-1}_{\mathsf {I_1, \dots ,I_{k-1}},\langle l_k, l_k+1)} (\mathsf {A}_1, \ldots , \mathsf {A}_{k+1}). \end{equation}\) This formulae now falls under case 1.1, as \(I_k\) and \(I_{k-1}\) are non-overlapping and bounded and, \(l_{k} + 1 \le \mathsf {cmax}\). Hence, it can be handled similarly. Let \(\psi ^{1.2.1}\) be the formula that we get after applying the reduction as mentioned in case 1.1 on \(\phi {1.2.1}_f\).

Possibility 2: \(\tau ^{\prime }_{i^{\prime }_k} \gt \lceil \tau ^{\prime }_{i^{\prime }}+l_k \rceil\). In this case, we use the oversampling point, c, at integer timestamp \(\lceil \tau ^{\prime }_{i^{\prime }}+l_k \rceil\) to break the check for \(\mathsf {A}_k\). Note that if the nearest oversampling integer point next to \(i^{\prime }\) is labelled \(\mathsf {int}_j\), then (by \(\varphi _{ovs}\)), \(\lceil \tau ^{\prime }_{i^{\prime }}+l_k \rceil\) is labelled as \(\mathsf {int}_{j^{\prime }}\) where \(j^{\prime } = j \oplus l_k\). Moreover, as the time difference between \(\mathsf {int}_{j^{\prime }}\) and \(i^{\prime }\) is less than \(l_k+1 \le \mathsf {cmax}\), there is no other oversampling point between \(i^{\prime }\) and c (hence, between \(i^{\prime }_{k-1}\) and c) with the same label as \(\mathsf {int}_{j^{\prime }}\). Hence, to assert \(\mathsf {A}_{k}\) from \(i^{\prime }_{k-1}\) to some point beyond c, we start with checking for \(\mathsf {A}^{\prime }_{k}\) from \(i^{\prime }_k\) to c reaching some state q and continuing the run from c to check for the remaining part starting from q and ending up at some final state f at some point (i.e., the required \(i^{\prime \prime }\)) in the future of c from where we again assert that the behavior till the end of the word is accepted by \(\mathsf {A}^{\prime }_{k+1}\) (in which case \(i^{\prime \prime }\) becomes the required \(i^{\prime }_{k}\)). This can be expressed using the following formula: (23) \(\begin{equation} \psi ^{1.2.2} = \bigvee \limits _{j=0}^{\mathsf {cmax}-1}\left[(\mathsf {F}_{[0,1)} \mathsf {int}_j) \rightarrow \left\lbrace \bigvee \limits _{q \in Q}\bigvee \limits _{f \in F} [\mathcal {F}^{k-1}_{\mathsf {I_1, \dots ,I_{k-1}}} (\mathsf {A}_1, \ldots , \mathsf {A}_{k-1}, \mathsf {A}^{\prime }_k[init,q] \cdot \mathsf {int}_{j^{\prime }},\mathsf {A}^{\prime }_k[q,f] \cdot \mathsf {A}_{k+1})\right]\right\rbrace \!. \end{equation}\) We encourage readers to refer to Figure 11. Hence, \(\psi ^{1.2} = \psi ^{1.2.1} \vee \psi ^{1.2.2}\) is the required formula for this case. Note that Possibilities 1 and 2 are not disjoint. That is, there are positions of \(i^{\prime }_k\) that fall within both sets of possibilities. This simply means that there are models for which both \(\psi ^{1.2.1}, \psi ^{1.2.2}\) hold. Hence, the restrictions imposed by both these formulae might be redundant, but together both these formulae cover all the possibilities for occurrence of \(i^{\prime }_k\).

Fig. 11.

Fig. 11. Figure showing reduction of k-ary \(\text{PnEMTL}\) formulae to \(k-1\) -ary \(\text{PnEMTL}\) formulae when the intervals \(I_{k-1}\) and \(I_{k}\) are non-overlapping but \(I_k\) is unbounded (Case 1.2). In this case, we will not be able to make sure that time difference between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\) is bounded by \(\mathsf {cmax}\) . This figure highlights the reduction for only one of the possibilities ( \({i^{\prime }_k}\) occurs after time \(\lceil \tau ^{\prime }_{i^{\prime }}+ l_{k} \rceil\) and hence beyond \(l_k\) time from \(i^{\prime }\) ) of this case and hence gives one of the disjunct for required formula. The behavior from \(i^{\prime }\) to oversampling point at time \(\lceil \tau ^{\prime }_{i^{\prime }}+ l_{k} \rceil\) (say, c) is given by the part of the following formula colored blue \(\color {blue}{\mathcal {F}_{I_1, \ldots , I_{k-1}} (\mathsf {A}^{\prime }_1, \ldots , \mathsf {A}^{\prime }_k[init,q]\cdot \mathsf {int}_{j^{\prime }}}\color {red}{,\mathsf {A}^{\prime }_k[q,f]\cdot \mathsf {A}^{\prime }_{k+1}})\) and the part beyond c is given by the part of the formula colored in red where \(j^{\prime } = j \oplus l_k\) , q is any state of \(\mathsf {A}^{\prime }_{k}\) reached when read till c and f is the final state that is reached when \(i^{\prime }_k\) is read. Disjuncting over all possible (but finitely many) \(j \in \lbrace 0,\ldots , \mathsf {cmax}-1\rbrace\) , \(q \in Q_i\) and \(f \in F_i\) , we get the required formulae.

Case 2: \(I_{k-1}\) and \(I_k\) overlap each other. That is, \(l_k \lt u_{k-1}\) (again, the strict \(\lt\) is due to the fact that \(\phi ^{\prime }_f\) is a non-adjacent formula). Hence, it is possible that there is no oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_k\), because of which we can not only rely on Lemma 8.3. There are following subcases depending on how the intervals \(I_{k-1}\) and \(I_{k}\) overlap and whether \(I_k\) is bounded or not:

Case 2.1: \(I_k\) is bounded, \(l_{k-1} = l_{k}\) and \(u_{k-1} \lt u_{k}\). There are two possibilities based on the relative positions of \(i^{\prime }_{k-1}\) and \(i^{\prime }_k\).

Possibility 1: There is an oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\). As \(I_k\) is bounded, the time difference between the former and the latter is bounded by \(\mathsf {cmax}\). Hence, using Lemma 8.3 and identical reasoning used in case 1.1, the same formula \(\psi ^{1.1}\) takes care of this possibility.

Possibility 2: There is no oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\). If \(i^{\prime }_{k-1}\) lies within \(I_{k-1}\) from \(i^{\prime }\), then \(\tau ^{\prime }_{i^{\prime }_k} \le \lceil \tau ^{\prime }_{i^{\prime }_{k}} \rceil = \lceil \tau ^{\prime }_{i^{\prime }_{k-1}} \rceil\)(timestamps of both the points have same integer parts) \(\lt \tau ^{\prime }_{i^{\prime }_{k-1}}+1\)(property of ceiling function) \(\le \tau ^{\prime }_{i^{\prime }}+1+u_{k-1}\) (\(i^{\prime }_{k-1}\) lies within \(I_{k-1}\) of \(i^{\prime }\)) \(\le \tau ^{\prime }_{i^{\prime }}+u_{k}\) (\(u_k \lt u_{k-1}\) and \(u_k\) is an integer). Similarly, \(i^{\prime }_{k-1}\) lies within \(I_{k-1}\), and the \(i^{\prime }_k\) occurs after \(i^{\prime }_{k-1}\) implies the time difference between \(i^{\prime }_k\) and \(i^{\prime }\) is more than \(l_{k-1} = l_k\) units. Hence, (a) \(i^{\prime }_{k-1}\) within \(I_{k-1}\) from \(i^{\prime }\), (b) \(i^{\prime }_k\) occurs after \(i^{\prime }_{k-1}\), and (c) there is no oversampling point between them, implies that \(i^{\prime }_{k}\) is within \(I_k\) from \(i^{\prime }\). We check this inline using the following \(k-1\) ary \(\text{PnEMTL}\) formula: (a) is checked using the last interval \(I_{k-1}\), (b) is asserted by concatenating \(\mathsf {A}_k\) with \(\mathsf {A}^{\prime }_{k+1}\) appearing in the last argument, and (c) is asserted by using \(\mathsf {A}_k\) (which disallows any oversampling points) rather than \(\mathsf {A}^{\prime }_k\) for concatenation with \(\mathsf {A}^{\prime }_{k+1}\) in the last argument. Hence, the following formula covers this possibility: (24) \(\begin{equation} \psi ^{2.1.2} = \mathcal {F}^{k-1}_{I_1,\ldots I_{k-1}}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k-1},\mathsf {A}_k\cdot \mathsf {A}^{\prime }_{k+1}). \end{equation}\) We encourage the readers to go through Figure 12. Finally, the formula for this kind of overlapping of \(I_{k-1}\) and \(I_k\) is \(\psi ^{2.1} = \psi ^{1.1} \vee \psi ^{2.1.2}\).

Fig. 12.

Fig. 12. Figure showing reduction of k-ary \(\text{PnEMTL}\) formulae to \(k-1\) -ary \(\text{PnEMTL}\) formulae when the intervals \(I_{k-1}\) and \(I_{k}\) are overlapping but \(I_k\) is bounded \(l_k = l_{k-1}\) and \(u_k \gt u_{k-1}\) (Case 2.1). Hence, in this case, we will not be able to make sure that there is an oversampling point between \(i_{k-1}\) and \(i_{k}\) . This diagram covers the situation where there is no oversampling point between \(i_{k-}\) and \(i_k\) . For the situation where there is an oversampling point between these points, we can use the formula identical to case 1.1. If \(i^{\prime }_{k-1}\) is within \(I_{k-1}\) and \(i^{\prime }_{k}\) occurs after that but before the next oversampling point (that is, the integer part of their timestamps are same), then \(i_k\) is within \(I_k\) . Hence, we just need to check that from \(i^{\prime }_{k-1}\) there exists a point in the future c before the next oversampling point such that the behavior from \(i^{\prime }_{k-1}\) till that point is given by \(\mathsf {A}_{k}\) rather than \(\mathsf {A}^{\prime }_k\) (as the former disallows occurrence of all oversampling points) and from c onwards \(\mathsf {A}^{\prime }_{k+1}\) holds till the last point of the word. Then, c is a valid candidate for the required \(i^{\prime }_{k}\) . Hence, we remove the interval \(I_k\) from the given formula and replace the last argument \(\mathsf {A}^{\prime }_{k+1}\) by \({\mathsf {A}_{k}\cdot \mathsf {A}^{\prime }_{k+1}}\) .

Case 2.2: \(I_k\) is bounded, \(u_{k-1} = u_{k}\) and \(l_{k-1} \lt l_{k}\). This case is similar to the case above. There are two possibilities as in case 2.1.

Possibility 1: There is an oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\). Similar to case 1.1 and 2.1, \(\psi ^{1.1}\) covers this possibility.

Possibility 2: There is no oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\). The argument here is symmetric. But we just need to check \(i^{\prime }_{k}\)’s position rather than \(i^{\prime }_{k-1}\). If \(i^{\prime }_k\) lies within \(I_{k}\) from \(i^{\prime }\), then \(\tau ^{\prime }_{i^{\prime }_{k-1}} \ge \lfloor \tau ^{\prime }_{i^{\prime }_{k}} \rfloor\) (timestamps of both the points have same integer parts) \(\gt \tau ^{\prime }_{i^{\prime }_{k}}-1\) (property of floor function) \(\ge \tau ^{\prime }_{i^{\prime }}-1+l_{k}\) (\(i^{\prime }_{k}\) lies within \(I_{k}\) of \(i^{\prime }\)) \(\le \tau ^{\prime }_{i^{\prime }}+l_{k-1}\) (\(l_{k-1} \lt l_{k}\) and \(l_{k-1}\) is an integer). Similarly, \(i^{\prime }_{k}\) lies within \(I_{k}\) and the \(i^{\prime }_{k-1}\) occurs before \(i^{\prime }_{k}\) implies the time difference between \(i^{\prime }_{k-1}\) and \(i^{\prime }\) is less than \(u_k\) = \(u_{k-1}\) units. Hence, (a) \(i^{\prime }_k\) occurring within \(I_{k}\) from \(i^{\prime }\), (b) \(i^{\prime }_{k-1}\) occurs after \(i^{\prime }_{k}\), and (c) there is no oversampling point between them, implies that \(i^{\prime }_{k-1}\) is within \(I_{k-1}\) from \(i^{\prime }\). We check this inline using the following \(k-1\)-ary \(\text{PnEMTL}\) formula: (a) is checked setting the last interval \(I_{k}\), (b) is asserted by concatenating \(\mathsf {A}^{\prime }_{k-1}\) with \(\mathsf {A}_{k}\) in the second last argument, and (c) is asserted by using \(\mathsf {A}_k\) (which disallows any oversampling points) rather than \(\mathsf {A}^{\prime }_k\) for concatenation with \(\mathsf {A}^{\prime }_{k-1}\) in the last but second argument. Hence, the following formula covers this possibility: (25) \(\begin{equation} \psi ^{2.2.2} = \mathcal {F}^{k-1}_{I_1,\ldots , I_{k-2}, I_{k}}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k-1}\cdot \mathsf {A}_k, \mathsf {A}^{\prime }_{k+1}). \end{equation}\) Finally, the formula for this kind of overlapping between \(I_{k-1}\) and \(I_k\) is \(\psi ^{2.2} = \psi ^{1.1} \vee \psi ^{2.2.2}\).

Case 2.3: \(I_k\) is bounded, \(l_{k-1} \lt l_{k}\) and \(u_{k-1} \lt u_{k}\). As above, there are two possibilities. Possibility 1, when point \(i^{\prime }_{k-1}\) and \(i^{\prime }_k\) have an oversampling point in between them. This possibility is identical to case 1.1 and hence \(\psi ^{1.1}\) covers it. For possibility 2, when both the points are within same integer timestamps, consider the following: Let \(I^{\prime } = I_{k-1} \cap I_{k} = \langle l_k, u_{k-1} \rangle .\)12 Then, for possibility 2 to occur, either (a) \(i^{\prime }_{k-1}\) occurs within \(I^{\prime }\) from \(i^{\prime }\) or (b) \(i^{\prime }_k\) occur within \(I^{\prime }\) from \(i^{\prime }\), because if both of them do not occur within the intersection, then there is at least one oversampling point between them (which is already covered by \(\psi ^{1.1})\)). Hence, it suffices to reduce arity of formula \(\phi ^{2.3} = \phi ^{2.3}_{a} \vee \phi ^{2.3}_{b}\) for this possibility where (26) \(\begin{equation} \phi ^{2.3}_{a} = \mathcal {F}^{k-1}_{I_1,\ldots ,I_{k-2}, I^{\prime }, I_{k}}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k-1},\mathsf {A}^{\prime }_k, \mathsf {A}^{\prime }_{k+1}), \end{equation}\) (27) \(\begin{equation} \phi ^{2.3}_{b} = \mathcal {F}^{k-1}_{I_1,\ldots ,I_{k-2}, I_{k-1}, I^{\prime }}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k-1},\mathsf {A}^{\prime }_k, \mathsf {A}^{\prime }_{k+1}). \end{equation}\) Note that these k-ary \(\text{PnEMTL}\) formulae can be reduced individually, as \(\phi ^{2.3}_{a}\) falls under the case 2.1, while \(\phi ^{2.3}_{b}\) falls under the case 2.2. Let \(\psi ^{2.3}_{a}\) and \(\psi ^{2.3}_{b}\) be the formulae we get after applying the reduction from cases 2.1 and 2.2 to the formulae \(\phi ^{2.3}_{a}\) and \(\phi ^{2.3}_{b}\), respectively. Then, the required formula covering this case is \(\psi ^{2.3} = \psi ^{1.1} \vee \psi ^{2.3}_{a} \vee \psi ^{2.3}_{b}\).

Case 2.4: \(l_{k-1} = l_k = l\) and \(u_{k-1} = u_k = u\). Let \(I_{k-1} = (l,u) = I_k\).13 Like the previous subcases, possibility in which there is an oversampling point between \(i^{\prime }_k\) and \(i^{\prime }_{k-1}\), is handled by formula \(\psi ^{1.1}\). There are two other possibilities.

Possibility 2: (a) There is no oversampling point between \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\), and (b)\(\tau ^{\prime }_{i^{\prime }_{k-1}} \in (\tau ^{\prime }_{i^{\prime }} + l, \lfloor \tau ^{\prime }_{i^{\prime }} + u \rfloor)\). Note that (a) and (b) implies \(i^{\prime }_{k-1}\) and \(i^{\prime }_{k}\) are within \((l,u)\) from \(i^{\prime }\). To check (a), we nest a \(\mathcal {F}\) modality within the \(k-1\)-ary \(\text{PnEMTL}\) formula asserting \(\mathsf {A}_k\) instead of \(\mathsf {A}^{\prime }_k\) from point \(i^{\prime }_{k-1}\) in the kth argument of \(k-1\)-ary \(\text{PnEMTL}\) formula (see formula \(\Gamma _j\)). To check (e), we just have to assert that if \(\rho ^{\prime }, i^{\prime } \models \mathsf {F}_{[0,1)} (\mathsf {int}_j)\), then \(\tau ^{\prime }_{i^{\prime }_{k-1}} - \tau ^{\prime }_{i^{\prime }} \in (l, u) \wedge \rho ^{\prime }, i^{\prime }_{k-1} \models \lnot \mathsf {F}_{[0,1)}(\mathsf {int}_{j\oplus u_{k}})\) (again, see formula \(\Gamma _j\)). Let \(\Gamma _j = \lnot \mathsf {F}_{[0,1)}(\mathsf {int}_{j\oplus u_k}) \wedge \mathcal {F}(\mathsf {A}_k \cdot \mathsf {A}^{\prime }_{k+1})\) and \(S^{\prime \prime } = S^{\prime } \cup \Gamma _j\). Let \(\mathcal {S}\) sets of subsets of \(S^{\prime \prime }\) containing \(\Gamma _j\). (28) \(\begin{equation} \psi ^{2.4(i)} = \bigvee \limits _{X \in \mathcal {S}} \subseteq \bigvee \limits _{j=1}^{\mathsf {cmax}} (\mathsf {F}_{[0,1)}(\mathsf {int}_j) \rightarrow \mathcal {F}^{k-1}_{I_1,\ldots ,I_{k-1}}(\mathsf {A}^{\prime \prime }_1,\ldots , \mathsf {A}^{\prime \prime }_{k-1}, X \cdot \Sigma ^*)(S^{\prime \prime }), \end{equation}\) where \(\mathsf {A}^{\prime \prime }_i = \mathsf {A}^{\prime \Gamma _j}_i\). That is, the transitions of \(\mathsf {A}^{\prime \prime }_i\) do not depend on the truth value of \(\Gamma _j\).

Possibility 3: (a) holds and (c) \(\tau ^{\prime }_{i^{\prime }_{k}} \in (\lfloor \tau ^{\prime }_{i^{\prime }} + u_k \rfloor , \tau ^{\prime }_{i^{\prime }} + u_k)\). Then, (a) and (c) implies \(i^{\prime }_{k-1}\) and \({i^{\prime }_{k}}\) are within \((l,u)\) from \(i^{\prime }\). Like the previous possibility, to check (a), concatenate \(\mathsf {A}^{\prime }_{k-1}\) with \(\mathsf {A}_k\) instead of \(\mathsf {A}^{\prime }_k\) in the \(k-1\)th argument of \(k-1\)-ary \(\text{PnEMTL}\) formula. To check (c), we just have to assert that if \(\rho ^{\prime }, i^{\prime } \models \mathsf {F}_[0,1) (\mathsf {int}_j)\), then \(\tau ^{\prime }_{i^{\prime }_{k}} - \tau ^{\prime }_{i^{\prime }} \in (l, u) \wedge \rho ^{\prime }, i^{\prime }_{k-1} \models \mathsf {F}_{[0,1)}(\mathsf {int}_{j\oplus u_{k}})\) (check \(\Gamma ^{\prime }_j\)). Let \(\Gamma ^{\prime }_j = \mathsf {F}_{[0,1)}(\mathsf {int}_{j\oplus u_k}) \wedge \mathcal {F}(\mathsf {A}^{\prime }_{k+1})\), \(S^{\prime \prime } = S^{\prime } \cup \Gamma ^{\prime }_j\). Let \(\mathcal {S}^{\prime }\) sets of subsets of \(S^{\prime \prime }\) containing \(\Gamma ^{\prime }_j\). (29) \(\begin{equation} \psi ^{2.4(ii)} = \bigvee \limits _{X \in \mathcal {S}^{\prime }} \wedge \Gamma _j \in S \subset \bigvee \limits _{j=1}^{\mathsf {cmax}} (\mathsf {F}_{[0,1)}(\mathsf {int}_j) \rightarrow \mathcal {F}^{k-1}_{I_1,\ldots ,I_{k-2}, I_{k}}(\mathsf {A}^{\prime \prime }_1,\ldots , \mathsf {A}^{\prime \prime }_{k-1}\cdot \mathsf {A}_k^{\Gamma ^{\prime }_j}, X\cdot \Sigma ^*), \end{equation}\) where \(\mathsf {A}^{\prime \prime }_i = \mathsf {A}^{\prime \Gamma ^{\prime }_j}_i\) for \(i\le k-1\). Let \(\psi ^{2.4} = \psi ^{1.1} \vee \psi ^{2.4(i)} \vee \psi ^{2.4(ii)}\).

Case 2.5: \(I_{k}\) is an unbounded interval. We break this case into two possibilities. (1) \(i^{\prime }_{k}\) occurs within \(J_1 = \langle l_{k}, u_{k-1} + 1)\).14 (2) \(i^{\prime }_{k}\) occurs within \(J_2 = [u_{k-1} + 1, \infty)\). Hence, \(\phi ^{\prime }_f\) can be rewritten as \(\phi ^{\prime }_{f,1} \vee \phi ^{\prime }_{f,2}\) where for \(i \in \lbrace 1,2\rbrace\), (30) \(\begin{equation} \phi ^{\prime }_{f,i} = \mathcal {F}^{k}_{I_1,\ldots ,I_{k-1}, J_i}(\mathsf {A}^{\prime }_1,\ldots , \mathsf {A}^{\prime }_{k-1},\mathsf {A}^{\prime }_k, \mathsf {A}^{\prime }_{k+1}). \end{equation}\) Note that \(\phi ^{\prime }_{f,1}\) falls under case 2.3 if \(l_k \gt l_{k-1}\) and case 2.1 if \(l_{k} = l_{k-1}\). Moreover, \(\phi ^{\prime }_{f,2}\) is within the case 1.2 and, hence, can be handled accordingly. Let \(\psi ^{\prime }_{f,1}\) and \(\psi ^{\prime }_{f,2}\) be the formulae we get after applying corresponding reductions to \(\phi ^{\prime }_{f,1}\) and \(\phi ^{\prime }_{f,2}\), respectively. Then, \(\psi ^{2.5} = \psi ^{\prime }_{f,1} \vee \psi ^{\prime }_{f,2}\).

Note that all other cases are disallowed by Assumptions 1 and 2.

Hence, the required formula \(\phi ^{k-1}_f\) depends on the type of intervals \(I_k\) and \(I_{k-1}\). For example, if \(I_{k}\) is bounded and does not have an intersection with \(I_{k-1}\), then it falls within case 1.1 and \(\psi ^{1.1}\) is the required \(\text{PnEMTL}\). Moreover, note that the total number of operators (temporal, Boolean, etc.) in \(\text{PnEMTL}\) and \(\text{EMITL}_{0,\infty }\) is \(\mathcal {O}(\mathsf {cmax}\times |Q_k| \times |F_k| \times |\phi _f|)\).□

After recursively applying the above reductions, we get a formula \(\psi _f\) that contains modalities from \(\text{EMITL}_{0,\infty }\) and \(\mathsf {F}_{np}\) of the size in \(\mathcal {O}((\mathsf {cmax}\times |\phi _f|)^{n}) = \mathcal {O}(|\phi _f|^{Poly(n,M)}),\) where n is the arity of \(\phi _f\) and M is the number of bits required to store the constants of its timing interval.

8.4 Eliminating 𝖥l, u Modalities Where l > 0 And u ≠ ∞

Let \(I = \langle l, u \rangle\) be any interval appearing in \(\psi ^{\prime }\) where \(l\gt 1\) and \(u\lt \infty\) (hence, \(u\le \mathsf {cmax}\)). Let \(\rho ^{\prime } = (\sigma ^{\prime }_1,\tau ^{\prime }_1) \ldots (\sigma ^{\prime }_{m^{\prime }}, \tau ^{\prime }_{m^{\prime }})\) such that \(\rho ^{\prime } \models \varphi _{ovs}\). In this section, given any formula of the form \(\mathsf {F}_{I}(\alpha)\), we construct a specification \(\delta ({{I}},\alpha)\) using \(\alpha\) and modalities from \(\text{MITL}_{0,\infty }\) such that, for any \(i^{\prime }\in dom(\rho ^{\prime })\), \(\rho ^{\prime },i^{\prime } \models \mathsf {F}_{I}(\alpha)\) iff \(\rho ^{\prime },i^{\prime }\models \delta (\mathsf {F}_{I},\alpha)\). Notice that \(\rho ^{\prime },i^{\prime } \models \mathsf {F}_{I} (\alpha)\) iff there exists a point \(j\gt i^{\prime }\) such that \(\tau ^{\prime }_j - \tau ^{\prime }_{i^{\prime }} \in \langle l, u \rangle\) and \(\rho ^{\prime },j \models \alpha\). Let c be the nearest integer oversampling point after \(i^{\prime }\). Let \(c^{\prime }\) be the integer oversampling point with timestamp \(\lceil \tau ^{\prime }_{i^{\prime }} + l \rceil\). There are two possibilities, depending on the occurrence of j.

Case 1: Either \(\tau ^{\prime }_j \in \langle \tau ^{\prime }_{i^{\prime }} + l, \lceil \tau ^{\prime }_{i^{\prime }} + l \rceil ]\). This implies j occurs before \(c^{\prime }\). If \(\mathsf {int}_{j}\) is true at point c, then \(\mathsf {int}_{j\oplus l}\) is true at point \(c^{\prime }\). Moreover, \(c^{\prime }\) will be the very first point after \(i^{\prime }\) with \(\mathsf {int}_{j\oplus l}\) counter value as \(l \le \mathsf {cmax}\). Hence, if j is any point in \(\langle l,\infty)\) from \(i^{\prime }\) that occurs before c, then timestamp of j is within \(\langle \tau ^{\prime }_{i^{\prime }} + l, \lceil \tau ^{\prime }_{i^{\prime }} + l \rceil ]\). This could be easily expressed using formula \(\delta _1({I}, \alpha) = \bigvee \nolimits _{i=0}^{\mathsf {cmax}-1} [\mathsf {F}_{[0,1)}(\mathsf {int}_j) \rightarrow (\lnot \mathsf {int}_{j\oplus l} \mathsf {U}_{\langle l, \infty)} \alpha)].\)

Case 2: Or \(\tau ^{\prime }_j \in (\lceil \tau ^{\prime }_{i^{\prime }} + l \rceil , \tau ^{\prime }_{i^{\prime }}+ u \rangle\). Hence, j occurs after \(c^{\prime }\). Notice that u is not greater than \(\mathsf {cmax}\). Hence, all the oversampling points in \(\langle \tau ^{\prime }_{i^{\prime }}, \tau ^{\prime }_{i^{\prime }}+ u \rangle\) are labelled with unique counters as the counters increment modulo \(\mathsf {cmax}\). Hence, the counter values at all the oversampling points between c and \(c^{\prime }\) are different than the counter values (or labels) at all the oversampling points with timestamp in \(\tau ^{\prime }_{i^{\prime }} + l \rceil , \tau ^{\prime }_{i^{\prime }}+ u \rangle\). More precisely, if c is labelled with \(\mathsf {int}_j\), then all the oversampling points within \((\tau ^{\prime }_{i^{\prime }}, \lceil \tau ^{\prime }_{i^{\prime }} + l \rceil ]\) will be labelled with propositions in \(\lbrace \mathsf {int}_{j\oplus 1},\ldots \mathsf {int}_{j \oplus l}\rbrace\), while the oversampling points within \((\lceil \tau ^{\prime }_{i^{\prime }} + l \rceil , \lceil \tau ^{\prime }_{i^{\prime }}+ u \rceil \rangle\) will be labelled with propositions from \(E = \lbrace \mathsf {int}_{j\oplus l+1},\ldots , \mathsf {int}_{j \oplus u}\rbrace\). Hence, any point within \([0,u\rangle\) of \(i^{\prime }\) where \(\mathsf {F}_{[0,1)} \bigvee E\) holds, occurs within time \((\lceil \tau ^{\prime }_{i^{\prime }} + l \rceil , \tau ^{\prime }_{i^{\prime }}+ u \rangle\). Hence, to assert that \(j \in (\lceil \tau ^{\prime }_{i^{\prime }} + l \rceil , \tau ^{\prime }_{i^{\prime }}+ u \rangle\), we construct formula \(\delta _2({I}, \alpha) = \bigvee \nolimits _{i=0}^{\mathsf {cmax}-1} [\mathsf {F}_{[0,1)}(\mathsf {int}_j) \rightarrow \mathsf {F}_{[0,u)}(\alpha \wedge \mathsf {F}_{[0,1)} (\bigvee E))]\).

Hence, \(\delta ({I}, \alpha) = \delta _1({I}, \alpha) \vee \delta _2({I}, \alpha)\) is the required formula free from any bounded interval with non-zero lower bound (provided \(\alpha\) is free from such intervals). Notice that size of \(\delta (I, \alpha) = \mathcal {O}(\mathsf {cmax}\times |\mathsf {F}_I(\alpha)|)\).

Hence, when this step is applied to formula \(\psi _f\) from the previous step, we get a formula \(\psi ^{\prime }_f\) in \(\text{EMITL}_{0,\infty }\) (\(\text{MITL}_{0,\infty }\) is a sublogic of \(\text{EMITL}_{0,\infty }\)). Moreover, in the DAG corresponding to \(\psi _f\) there will be at most \(|\psi _f|\) \(\mathsf {F}_{np}\) operators. Hence, size of \(\psi ^{\prime }_f\) is \(\mathcal {O}(|\psi _f| \times \mathsf {cmax}) = \mathcal {O}(|\phi _f|^{Poly(n,M)})\).

Applying all the above steps to every \(\text{PnEMTL}\) modality in \(\phi _{flat}\), we get a formula \(\psi ^{\prime } \in \text{EMITL}_{0,\infty }\) that is equisatisfiable to \(\phi\). Moreover, the size of \(\psi ^{\prime }\) is at most \(|\psi |\) times the size of max \(\psi ^{\prime }_f\). Hence, \(\psi ^{\prime }\) is of the size in \(\mathcal {O}(|\phi |^{Poly(n,M)})\).

8.5 Concluding Proof

The above three steps of construction show that:

The equisatisfiable \(\text{EMITL}_{0,\infty }\) formula \(\psi\) is of the size \((\mathcal {O}(|\phi |^{Poly(n,M)})\), where n is the arity \(\phi\) and M is the number of bits required to store the constants appearing in the timing intervals. This is because reducing the arity of each subformula \(\phi _f\) by 1 results in formula of \(O(\mathsf {cmax}\times |\phi _f|^3)\) size. Hence, after recursively applying the reduction to get the final \(\text{EMITL}_{0,\infty } + \mathsf {F}_{np}\) formula, we get a formula of the \(O((\mathsf {cmax}\times |\phi |^3)^n)\) (\(|\phi | \gt |\phi _f|\)) size. Eliminating \(\mathsf {F}_{np}\) will blow up the size by \(\mathsf {cmax}\). (i) Hence, the required formula is of \(O(\mathsf {cmax}\times (\mathsf {cmax}\times |\phi |^3)^n)\) size. As \(\mathsf {cmax}= 2^{M}\) (M defined above), the size of the final required formula is bounded by \(O(2^{Poly(n,M)})\).

For a non-adjacent 1-\(\text{TPTL}\) formula \(\gamma\), applying the reduction in Section 5 yields \(\phi\) of size (ii) \(\mathcal {O}(2^{Poly|\gamma |})\) and, arity of \(\phi = \mathcal {O}(|\gamma |^2)\) and the set of constants remain the same. Note the set of constants used in the timing interval of output formula \(\phi\) is the same as that of \(\gamma\). Hence, the number of bits required to store the constants in \(\phi\) is (iii) \(M = \mathcal {O}(|\gamma |)\). Also, after applying the reduction of Section 8 by plugging the value of \(|\phi |\) and its arity from (ii) and value of M from (iii) in (i), we get the \(\text{EMITL}_{0,\infty }\) formula \(\psi\) of size \(\mathcal {O}(2^{Poly(|\gamma |)*Poly(n,M)})= \mathcal {O}(2^{Poly(|\gamma |)})\).

By Lemma 7.3, given any formula \(\gamma\) in NA-GQMSO, we can construct an equivalent formula \(\phi\) in NA-PnEMTL (with non-elementary blow-up) that can then be analyzed for satisfaction, as presented above. Hence, satisfiability for NA-GQMSO is decidable. Non-elementary lower bound for NA-GQMSO is inherited by the subclass FO[\(\lt\)].

Skip 9A NOTE ON INFINITE TIMED WORDS Section

9 A NOTE ON INFINITE TIMED WORDS

Up until this point, we have restricted our models to be finite timed words. Let \(\Sigma\) be any finite set of propositions. An infinite or \(\omega\)-timed word over \(\Sigma\) is an infinite sequence of the form \((\sigma _1, \tau _1) (\sigma _2, \tau _2)\ldots\) where \(\forall \in \mathbb {N}\), \(\sigma _i\) is a non-empty subset of \(\Sigma\), \(\tau _1 = 0\) and \(i,j \in \mathbb {N}\) \(i \lt j\) implies \(\tau _i \le \tau _j\). An \(\omega\)-timed word is said to be zeno if the limit of the sequence \(\tau _1, \tau _2, \ldots ,\) is not infinite. This means there are infinite actions within a finite duration. For example, \((a,0) (a, 0.5) (a, 0.75) (a, 0.875) \ldots\) is a zeno timed word. It is a common practice in the literature to restrict the models to non-zeno words, as physical systems do not exhibit zeno behavior: It would take infinite amount of energy to carry out infinitely many actions in finite time. Hence, we restrict ourselves to non-zeno models.15 The set of all non-zeno \(\omega\)-timed words over \(\Sigma\) is denoted by \(T\Sigma ^{\omega }\). On closer inspection, it can be seen that all the results for finite timed words in the previous sections can be easily lifted to infinite timed words. We point out the required modifications for this lifting. In the rest of this section, let \(\rho = (\sigma _1, \tau _1) \ldots\) be any non-zeno \(\omega\)-timed word. Let \(\mathsf {A}\) be any Büchi Automata. Let \(L^\omega (\mathsf {A})\) denote the set all untimed \(\omega\)-words accepted by \(\mathsf {A}\).

9.1 Definition of Logics over Infinite Timed Words

The syntax and semantics for logic \(\text{LTL}\), \(\text{MTL}\), \(\text{TPTL},\) and \(\text{MSO}\) remain the same. For \(\text{EMTL}\) and \(\text{PnEMTL},\) the following changes are required:

9.1.1 EMTL Extended with Büchi Automata Modalities.

We extend EMTL with a new modality, \(\mathcal {F}^\omega (\mathsf {A})(S)\), where \(\mathsf {A}\) is a Büchi Automata modality over subformulae S. Intuitively, this modality asserts that from the given point the suffix is accepted by \(\mathsf {A}\). Let \(S = \lbrace \varphi _1, \ldots , \varphi _n\rbrace\). For any \(x \in \mathbb {N}\), let \(S_x\) be the exact subset of formulae in S that holds at point x of \(\rho\). Then, for any \(i \in \mathbb {N}\), \(\mathsf {Seg}^{\omega }(\rho , i, S)\) is an untimed \(\omega\)-word \(S_i S_{i+1} \ldots\). Define \(\rho , i\models \mathcal {F}^\omega (\mathsf {A})(S)\) iff \(\mathsf {Seg}^{\omega }(\rho , i, S) \in L^\omega (\mathsf {A})\).

9.1.2 PnEMTL Extended with Büchi Automata Modalities.

Syntactically, in the modality \(\mathcal {F}^{k}(\mathsf {A}_1, \ldots , \mathsf {A}_{k+1})\), \(\mathsf {A}_{k+1}\) is a Büchi Automata, while the rest \(\mathsf {A}_1, \ldots , \mathsf {A}_{k}\) are classical non-deterministic automata with reachability objective. The new semantics of the \(\mathcal {F}^{k}_{\mathsf {I_1,\ldots , I_k}}\) modality is as follows:

\(\rho ,i_0 \models \mathcal {F}^k_{I_1,\ldots ,I_k}(\mathsf {A}_1,\ldots ,\mathsf {A}_{k+1})(S)\) iff \({\exists } {i_0 \le i_1\le i_2 \ldots \le i_k}\) s.t. \(\bigwedge \nolimits _{w=1}^{k}{[(\tau _{i_w} - \tau _{i_0} \in I_w)} \wedge \mathsf {Seg^+}(\rho , i_{w-1}, i_w,S) \in L({\mathsf {A}_w})] \wedge \mathsf {Seg^\omega }(\rho , i_{k}, S) \in L^\omega ({\mathsf {A}_{k+1}})\).

Intuitively, as the suffix from \(i_k\) onwards is infinite, it is natural to check the behavior in that region by a Büchi Automaton. The syntax and semantics of the \(\mathcal {P}^{k}_{\mathsf {I_1,\ldots , I_k}}\) modality does not change.

9.2 Anchored Interval Infinite Timed Words

As the name suggests, Anchored Interval \(\omega\)-words are \(\omega\) extension of interval words. For the sake of completeness, we define these formally. Let \(I_\nu \subseteq \mathcal {I}_\mathsf {int}\). An \(I_\nu\) anchored \(\omega\)-interval word over \(\Sigma\) is an \(\omega\)-word \(\kappa\) of the form \(\sigma _1 \sigma _2 \ldots \in 2^{\Sigma \cup \lbrace anch\rbrace \cup I_\nu }\) such that there is a unique point \(i \in \mathbb {N}\) where \(\mathsf {anch}\) holds. As before, this point is called an anchor point of \(\kappa\) and denoted by \(\mathsf {anch}(\kappa)\). Moreover, for every \(i \in \mathbb {N}\), \(\Sigma \cap \sigma _i \ne \emptyset\). That is, at every point in \(\kappa\), at least one of the propositions from \(\Sigma\) holds. Let \(\rho = (\sigma _1, \tau _1) \ldots\) be any \(\omega\)-timed word. \(\rho ,i\) is consistent with an \(I_\nu\) \(\omega\)-interval word \(\kappa = \sigma ^{\prime }_1 \ldots\) if and only if for any \(j \in \mathbb {N}\), \(\sigma ^{\prime }_j \cap \Sigma = \sigma _j\), \(\tau _j - \tau _i \in \sigma ^{\prime }_j \cap I_\nu\) and \(i = \mathsf {anch}(\kappa)\). Let \(\mathsf {Time}(\kappa)\) be all the non-zeno pointed \(\omega\)-timed word \(\rho ,i\) consistent with \(\kappa\). In what follows, let \(\kappa\) be an \(I_\nu\) \(\omega\)-interval word. Let \(I \in I_\nu\) be any interval of the form \(\langle l,u \rangle\), where \(u \ne \infty\). If \(\lbrace j | I \in \kappa [j]\rbrace\) is an infinite set (i.e., I occurs infinitely often in \(\kappa\)), then we call \(\kappa\) a Zeno Interval Word. The following proposition is straightforward:

Proposition 9.1.

If \(\kappa\) is a Zeno Interval Word and if \(\rho ,i\) is consistent with \(\kappa ,\) then \(\rho\) is a zeno word.

This is because there will be infinitely many points j is \(\rho = (\sigma _1, \tau _1) \ldots\) such that \(\tau _j - \tau _i \le u\). This, by definition, implies that \(\rho\) is a zeno word. Hence, if \(\kappa\) is a Zeno Interval Word, then \(\mathsf {Time}(\kappa)\) is an empty set. The definition of Collapsed \(\omega\)-Interval word is the same as Collapsed Interval words appearing in Section 4. As the proof of Lemma 4.2 does not require \(\kappa\) to be a finite word, it holds for non-zeno \(\omega\)-interval words, too.

9.2.1 Normalization.

For a collapsed non-zeno \(\omega\)-interval word \(\kappa\) and \(I \in I_\nu\), let \(\mathsf {first}(\kappa ,I)\) and \(\mathsf {last}(\kappa , I)\) denote the positions of first and last occurrence of I (as defined in Section 4). If I occurs infinitely often, then \(\mathsf {last}(\kappa ,I)\) is undefined. \(\mathsf {Norm}(\kappa) = \sigma ^{\prime }_1 \sigma ^{\prime }_2 \ldots\) is an \(I_\nu\) \(\omega\)-interval word built from \(\kappa\) as follows:

Reduction 1: For every unbounded interval \(I \in I\nu\), delete all the occurrences of I except the first one. Let this be denoted as \(R_1(\kappa)\)

Reduction 2: For every unbounded interval \(I \in I\nu\), delete all the occurrences of I except the first and the last one.

For any non-zeno word \(\kappa\), unbounded interval \(I = \langle l, \infty) \in I_\nu\) and \(x \in \mathbb {N}\), \(x = \mathsf {first}(\kappa , I)\) implies for all \(\rho ,i \in \mathsf {Time}(\kappa)\) and \(y \gt x\), \(\tau _y - \tau _i \in I\). Hence, any occurrence of interval I after its first occurrence is redundant, as the same restriction is imposed by the first occurrence of I. Hence, we have the following proposition:

Proposition 9.2.

For any non-zeno interval word \(\kappa\), if \(\kappa ^{\prime }\) is obtained from \(\kappa\) by applying Reduction 1 defined above, then \(\kappa \cong \kappa ^{\prime }\).

Hence, for any collapsed word non-zeno \(\kappa\), \(\kappa ^{\prime } = R_1(\kappa)\) will contain only finitely many time-restricted points. As every interval \(I\in I_\nu\) appears finitely often in \(\kappa ^{\prime }\), Lemma 4.4 is now applicable for \(\kappa ^{\prime }\). Hence, by Lemmas 4.2, 4.4, and 9.2, we have, for any non-zeno word \(\kappa\), \(\kappa \cong \mathsf {Norm}(\kappa)\).

9.3 Translation from 1-TPTL to PnEMTL

All the reduction in Section 5.1.1, i.e., translation from simple \(\text{TPTL}\) formulae to LTL over interval words, remains the same, as all the lemmas in that section hold for non-zeno infinite words, too. Translation from LTL to Büchi Automata over Collapsed Interval Words remains the same, as those techniques and results are standard for both finite and infinite timed words.

9.3.1 Partitioning of Interval Words.

While the general idea of partitioning the Language of NFA over the interval words into finitely many type sequences remains the same, we need to make some changes to the construction from A to \(Aut_{\mathsf {seq}}\) to incorporate Reduction 1 of normalization of \(\omega\)-interval words. In particular, we need to make sure that the \(Status(I_j)\) for any unbounded interval \(I_j \in I_\nu\) does not change from mid to post. This is because we essentially want to erase all the occurrences of \(I_j\) after the first one. Hence, \(Choice 1\) transition is deleted for unbounded intervals in \(Aut_{seq}\).

9.3.2 Reducing NFA of Each Type to PnEMTL.

Section 5.3 remains the same. All the automata \(\mathsf {A}_1\) to \(\mathsf {A}_k\) are automata over finite words (as the intervals only appear within the finite prefix of accepted words). Automaton \(\mathsf {A}_{k+1}\) is a Büchi Automaton. Hence, in the \(\text{PnEMTL}\) formulae, the last argument will be a Büchi Automaton, which is in agreement with the new syntax and semantics introduced in Section 9.1.

9.4 Equivalence of PnEMTL and GQMSO

Here, too, the existing reduction from PnEMTL to GQMSO and vice versa works. The only difference is, we need to use the standard Büchi Elgot Trakhtenbrot Theorem for infinite words.

9.5 Satisfiability Checking for Non-adjacent PnEMTL

This remains the same, too. The only difference is, wherever \(\mathsf {A}_{n+1}\) appears, it is a Büchi Automaton. As in the reduction mentioned in Section 8, \(\mathsf {A}_{n+1}\) always appears as the last argument (or the tail automaton) in the modalities of \(\text{EMITL}\) and \(\text{PnEMTL}\) in output and all the intermediate formulae. This is in accordance with the new syntax and semantics of \(\text{PnEMTL}\) for infinite words, as mentioned in Section 9.1. And in the base case Lemma 8.5, \(\mathsf {A}^{\prime }_2\) is a Büchi Automaton. We reduce the logic to \(\text{EMITL}_{0,\infty }\) extended with \(\mathcal {F}^{\omega }\). On inspection of Reference [27], the \(\mathcal {F}^{\omega }\) modality could be trivially reduced to Büchi Timed Automaton with size polynomial to that of the formulae. Hence, the result.

Skip 10CONCLUSION Section

10 CONCLUSION

We generalized the notion of non-punctuality to non-adjacency in logics \(\text{TPTL}\) and GQMSO. We proved that satisfiability checking for the non-adjacent 1-variable fragment of \(\text{TPTL}\) is EXPSPACE Complete. This gives us a strictly more expressive logic than \(\text{MITL}\) while retaining the satisfiability complexity. We introduced a new logic called \(\text{PnEMTL}\) and used it to solve the satisfiability checking problem for both non-adjacent 1-\(\text{TPTL}\) and GQMSO. The added expressive power over MITL comes with a useful ability to specify complex sequences of timing constraints over regular behaviors (automata). All our results, including decidability, extend to infinite timed words, as outlined in Section 9.

We believe that our logics and decidability results are useful for the specification and design of real-time systems. In model-based temporal planning, timing constraints on events are specified using logical formulae. Satisfiability checking of such formulae return a model that essentially gives a schedule meeting all the planning constraints. Several papers have investigated the use of TPTL with past modalities in formulating time-line-based planning [9, 10, 11, 21, 36]. Our expressive logics subsume several of these, and they offer a possibility of modelling even more general timing constraints involving regular behaviors (see the example below). The satisfiability checking method of this article potentially gives us a technique for automatic synthesis of plans. In another line of work investigating top-down design of real-time systems, assumptions and commitments over real-time systems are specified in a real-time logic. Moreover, design decisions (in the form of desired constraints on the behavior of the system to be implemented) can also be encoded in logic. Verification of this design step involves showing that the commitment is logically implied by the conjunction of assumptions and design decisions (see References [13, 38] for early examples of this approach). Validity checking of logical formulae (equivalently, satisfiability checking of negated formulae) permits automatic verification of such design decisions. However, an experimental validation of usefulness of these methods for practical planning and verification remains to be investigated.

Example 10.1.

Consider a job (e.g., automated pizza-maker) containing some high-level activities (involving several sub-steps) given by a sequence of finite state automata \(P_1, P_2, \ldots P_k\) (e.g., kneading the dough, preheating oven, baking the pizza) that has to be performed in a given sequence atomically (i.e., without pre-emption). Each process \(P_k\) has a deadline \(u_k\) associated with it. Now, we need to plan these processes such that the job is successfully completed within m time units. Moreover, there are some extra restrictions specified by, for instance, an MITL formula \(\varphi\). This could be done by finding a finite word satisfying the following formula: \(\begin{equation*} \phi = \mathcal {F}^k_{[0,u_1), \ldots , [0,u_k)} (P_1,P_2, \ldots , P_k, \mathsf {Finish}) \wedge \varphi . \end{equation*}\)

In case of GQMSO, the fact that the alternation of metric quantifiers in an anchored block can be eliminated using extra non-metric quantifiers (see Theorem 6.4) is an interesting result, in our opinion.

Finally, we pose the following open problems that we believe are fundamentally interesting and worth solving:

Is non-adjacent 1-\(\text{TPTL}\) strictly more expressive than \(\text{MITL}\) with Pnueli modalities (and hence Q2MLO of Reference [26])? We conjecture a positive answer to the question. More precisely, we conjecture that the property “within interval \((1,2)\) from the present point, events a and b occur such that a is immediately followed by event b” is not expressible using MITL with Pnueli modalities and hence in Q2MLO. But this is easily expressible in non-adjacent 1-\(\text{TPTL}\) and hence in GQFO (first-order fragment of GQMSO) as follows: \(x.\mathsf {F}(a \wedge T-x \in (1,2) \wedge \oplus (b \wedge T-x \in (1, 2)))\).

How does the logic non-adjacent GQMSO compare with the class of two-way deterministic timed automata with reversal boundedness [3] and MIDL [17]? Is there any natural subclass of timed automata corresponding to GQMSO? If yes, then it will be the largest known subclass of timed automata (to the best of our knowledge) that is closed under complementation. Ferrère in Reference [17] gives a very elegant extension of MITL called Metric Interval Dynamic Logic (MIDL), where the timing constraints are associated with regular expressions (Metric Interval Regular Expressions) as opposed to the modalities. While EMITL is a syntactic subclass of both non-adjacent PnEMTL and MIDL of Reference [17], there are still gaps in the expressiveness relationships amongst these logics. Ferrère already proved that MIDL is strictly more expressive than EMITL with only future automata modalities. However, (i) is EMITL with past modalities strictly included in non-adjacent PnEMTL and (ii) how non-adjacent PnEMTL compares with MIDL of Reference [17] (if you allow/disallow past operators in both logics) in terms of expressiveness is still open.

Efficient Tool Development for NA-1-TPTL Satisfiability and Model Checking. While we show that the satisfiability checking problem for NA-1-TPTL is in EXPSPACE, the algorithm is merely a proof-of-concept. We believe that in spite of the inherent worst-case theoretical complexity of the problem, in practice, we can build scalable tools for automated verification of NA-1-TPTL properties. Our decidability proof relies on the reduction of any NA-1-TPTL formula to an equisatisfiable EMITL\(_{0,\infty }\) formula. Using techniques similar to References [32] and [35], we can reduce EMITL\(_{0,\infty }\) formulae to equisatisfiable MITL formulae. This can then be followed by using scalable tools, such as MightyL [12] (automata-based tool) and Reference [7] (SMT-based tool) for MITL satisfiability and model checking. Using the reduction by Ho in Reference [27], we can also reduce this EMITL\(_{0,\infty }\) formula to an equivalent timed automata, which can be analyzed using scalable tools such as UPPAAL [6] and TChecker [24]. A direct simpler reduction from the satisfiability checking problem of NA-1-TPTL is also an intriguing open problem.

We also leave open an exploration for a suitable definition for non-adjacency and its satisfiability checking problem in the context of \(\text{TPTL}\) with multiple variables.

Skip ACKNOWLEDGMENTS Section

ACKNOWLEDGMENTS

We thank the reviewers of FM 2021 and the journal Formal Aspects of Computing for their feedback to improve the article.

Footnotes

  1. 1 Here, T is a special variable that stores the timestamp of the present point and x is the clock that was frozen when \(x.\) was asserted.

    Footnote
  2. 2 Bounded variability is usually defined on timed signals rather than timed words. But, every timed word can be equivalently represented as timed signals. Moreover, the definition of bounded variability of Reference [19] for timed words boils down to the above-mentioned restriction.

    Footnote
  3. 3 While Reference [37] proves decidability of \(\text{MTL}[\mathsf {U}]\) via reduction to 1-clock Alternating Timed Automata (1-ATA) followed by proving decidability for emptiness checking problem of 1-ATA over finite models. The generalization of this reduction is provided in Reference [22], where the authors prove a stronger result showing that 1-TPTL[U] with least fix-point operator is expressively equivalent to 1-ATA over finite models.

    Footnote
  4. 4 Even if we restrict the syntax to disallow Boolean expressions over constraints having a unique solution, it is possible to get undecidability due to the power of freeze quantification. The main power of adjacency comes from the fact that it could express the following kind of properties: a holds at the last/first point within the next/previous unit interval. For example, \(x.[\mathsf {F}\lbrace a \wedge x \in (0,1) \wedge \oplus (x \in (1,\infty))\rbrace ]\) (symbol \(\oplus\) stands for the next operator) specifies a holds at the last point in the next unit interval. This property can then be used to encode runs of any arbitrary 2 counter machines. See Reference [35], Chapter 3, Section 3.4, for more details.

    Footnote
  5. 5 We exclude this empty-set for technical reasons. This simplifies definitions related to equisatisfiable modulo oversampled projections [35]. Note that this does not affect the expressiveness of the models, as one can add a special symbol denoting the empty-set.

    Footnote
  6. 6 Note that we assume that the constants are encoded in binary.

    Footnote
  7. 7 Note, if \(I= [a, b),\) then \(-I = (-b, -a]\).

    Footnote
  8. 8 Note that, \(\text{LTL}(\varphi _1 ~\mathsf {U}~ \varphi _2) = \text{LTL}(\varphi _1) ~\mathsf {U}~\text{LTL}(\varphi _2).\)

    Footnote
  9. 9 We mention \(\mathsf {A}^{\prime }_k = S_{k-1}\cdot {\mathsf {A}_{k}}\cdot S_k\) instead of \(\cdot \,{\mathsf {A}_{k}}\cdot S_k\) due to the non-strict inequalities in the semantics of \(\text{PnEMTL}\) modalities.

    Footnote
  10. 10 In Reference [33], a similar logic called QkMSO was defined. QkMSO had yet another restriction: It can only quantify positions strictly in the future, and hence was not able to express past timed specifications.

    Footnote
  11. 11 Unbounded intervals can be eliminated using \(\mathcal {F}^{k}_{\mathsf {I_1, I_2, \ldots , I_{k-2},} [l_1, \infty) [l_2, \infty)}(\mathsf {A}_1, \ldots , \mathsf {A}_{k+1}) \equiv \mathcal {F}^{k}_{\mathsf {I_1, I_2, \ldots , I_{k-2},} [l_1, \mathsf {cmax}) [l_2, \infty)}\) \((\mathsf {A}_1, \ldots , \mathsf {A}_{k+1}) \vee \mathcal {F}^{k-1}_{\mathsf {I_1, I_2, \ldots , I_{k-2},} [l_2, \infty)}(\mathsf {A}_1, \ldots , \mathsf {A}_{k-1},\mathsf {A}_{k} \cdot \mathsf {A}_{k+1})\).

    Footnote
  12. 12 The left bracket will depend on interval \(I_{k}\) and the right will depend on \(I_{k-1}.\)

    Footnote
  13. 13 The proof can be extended to handle other kinds of intervals similarly.

    Footnote
  14. 14 \(u_{k-1} + 1 \le \mathsf {cmax}.\)

    Footnote
  15. 15 That is, language of any formula \(\phi\) can only contain non-zeno timed words.

    Footnote

REFERENCES

  1. [1] Alur R., Feder T., and Henzinger T.. 1996. The benefits of relaxing punctuality. J. ACM 43, 1 (1996), 116146.Google ScholarGoogle ScholarDigital LibraryDigital Library
  2. [2] Alur Rajeev, Feder Tomás, and Henzinger Thomas A.. 1991. The benefits of relaxing punctuality. In 10th Annual ACM Symposium on Principles of Distributed Computing. ACM, 139152. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  3. [3] Alur Rajeev and Henzinger Thomas A.. 1992. Back to the future: Towards a theory of timed regular languages. In 33rd Annual Symposium on Foundations of Computer Science. IEEE Computer Society, 177186. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  4. [4] Alur Rajeev and Henzinger Thomas A.. 1993. Real-time logics: Complexity and expressiveness. Inf. Comput. 104, 1 (1993), 3577. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  5. [5] Alur Rajeev and Henzinger Thomas A.. 1994. A really temporal logic. J. ACM 41, 1 (Jan.1994), 181203. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  6. [6] Bengtsson Johan, Larsen Kim Guldstrand, Larsson Fredrik, Pettersson Paul, and Yi Wang. 1995. UPPAAL—A tool suite for automatic verification of real-time systems. In DIMACS/SYCON Workshop on Verification and Control of Hybrid Systems(Lecture Notes in Computer Science, Vol. 1066). Springer, 232243. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  7. [7] Bersani Marcello M., Rossi Matteo, and Pietro Pierluigi San. 2015. An SMT-based approach to satisfiability checking of MITL. Inf. Computat. 245 (2015), 7297. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  8. [8] Bouyer Patricia, Chevalier Fabrice, and Markey Nicolas. 2005. On the expressiveness of TPTL and MTL. In Conference on Foundations of Software Technology and Theoretical Computer Science. Springer Berlin, 432443.Google ScholarGoogle ScholarDigital LibraryDigital Library
  9. [9] Bozzelli Laura, Molinari Alberto, Montanari Angelo, and Peron Adriano. 2019. Model checking timeline-based systems over dense temporal domains. In 20th Italian Conference on Theoretical Computer Science(CEUR Workshop Proceedings, Vol. 2504). CEUR-WS.org, 235247. Retrieved from http://ceur-ws.org/Vol-2504/paper27.pdf.Google ScholarGoogle Scholar
  10. [10] Bozzelli Laura, Molinari Alberto, Montanari Angelo, Peron Adriano, and Woeginger Gerhard J.. 2020. Timeline-based planning over dense temporal domains. Theor. Comput. Sci. 813 (2020), 305326. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  11. [11] Bozzelli Laura, Montanari Angelo, and Peron Adriano. 2019. Taming the complexity of timeline-based planning over dense temporal domains. In 39th IARCS Annual Conference on Foundations of Software Technology and Theoretical Computer Science(Leibniz International Proceedings in Informatics (LIPIcs), Vol. 150). Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik, Dagstuhl, Germany, 34:1–34:14. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  12. [12] Brihaye Thomas, Geeraerts Gilles, Ho Hsi-Ming, and Monmege Benjamin. 2017. MightyL: A compositional translation from MITL to timed automata. In 29th International Conference on Computer Aided Verification(Lecture Notes in Computer Science, Vol. 10426). Springer, 421440. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  13. [13] Chaochen Zhou, Hoare Charles Anthony Richard, and Ravn Anders P.. 1991. A calculus of durations. Inf. Process. Lett. 40, 5 (1991), 269276.Google ScholarGoogle ScholarCross RefCross Ref
  14. [14] Demri Stéphane, Goranko Valentin, and Lange Martin. 2016. Temporal Logics in Computer Science: Finite-state Systems, Vol. 58. Cambridge University Press.Google ScholarGoogle ScholarCross RefCross Ref
  15. [15] Eisner Cindy and Fisman Dana. 2006. A Practical Introduction to PSL. Springer.Google ScholarGoogle Scholar
  16. [16] Elgot Calvin C.. 1961. Decision problems of finite automata design and related arithmetics. Trans. Amer. Math. Soc. 98 (1961), 2151.Google ScholarGoogle ScholarCross RefCross Ref
  17. [17] Ferrère Thomas. 2018. The compound interest in relaxing punctuality. In 22nd International Symposium on Formal Methods, Held as Part of the Federated Logic Conference(Lecture Notes in Computer Science, Vol. 10951). Springer, 147164. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  18. [18] Language IEEE P1850-Standard for PSL-Property Specification. 2005. https://ieeexplore.ieee.org/document/4040498.Google ScholarGoogle Scholar
  19. [19] Furia Carlo A. and Rossi Matteo. 2008. MTL with bounded variability: Decidability and complexity. In Formal Modeling and Analysis of Timed Systems. Springer Berlin, 109123.Google ScholarGoogle Scholar
  20. [20] Gastin Paul and Oddoux Denis. 2003. LTL with past and two-way very-weak alternating automata. In 28th International Symposium on Mathematical Foundations of Computer Science(Lecture Notes in Computer Science, Vol. 2747). Springer, 439448. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  21. [21] Gigante Nicola. 2019. Timeline-based planning: Expressiveness and complexity. CoRR abs/1902.06123 (2019).Google ScholarGoogle Scholar
  22. [22] Haase Christoph, Ouaknine Joël, and Worrell James. 2010. On process-algebraic extensions of metric temporal logic. In Reflections on the Work of C. A. R. Hoare. Springer, 283300. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  23. [23] Henzinger Thomas A., Raskin Jean-François, and Schobbens Pierre-Yves. 1998. The regular real-time languages. In 25th International Colloquium on Automata, Languages and Programming(Lecture Notes in Computer Science, Vol. 1443). Springer, 580591. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  24. [24] Herbreteau Frédéric, Srivathsan B., and Walukiewicz Igor. 2016. Better abstractions for timed automata. Inf. Comput. 251 (2016), 6790. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  25. [25] Hirshfeld Y. and Rabinovich A.. 2006. An expressive temporal logic for real time. In Mathematical Foundations of Computer Science Conference. 492504.Google ScholarGoogle ScholarDigital LibraryDigital Library
  26. [26] Hirshfeld Yoram and Rabinovich Alexander. 2006. Expressiveness of metric modalities for continuous time. In Computer Science – Theory and Applications. Springer Berlin, 211220.Google ScholarGoogle Scholar
  27. [27] Ho Hsi-Ming. 2019. Revisiting timed logics with automata modalities. In 22nd ACM International Conference on Hybrid Systems: Computation and Control. ACM, 6776. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  28. [28] Hunter P.. 2013. When is metric temporal logic expressively complete? In Computer Science Logic (CSL’13), 380394. https://drops.dagstuhl.de/opus/volltexte/2013/4209/.Google ScholarGoogle Scholar
  29. [29] Design Specification IEEE Standard for SystemVerilog: Unified Hardware and Language Verification. 2005. IEEE Std 1800-2005 (2005), 1648. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  30. [30] Büchi J. R.. 1962. On a Decision Method in Restricted Second-order Arithmetic. .Google ScholarGoogle ScholarCross RefCross Ref
  31. [31] Madnani S. N. Krishna, K., and Pandya P. K.. 2014. Partially punctual metric temporal logic is decidable. In International Workshop on Temporal Representation and Reasoning. 174183.Google ScholarGoogle Scholar
  32. [32] Krishna Shankara Narayanan, Madnani Khushraj, and Pandya Paritosh K.. 2017. Making metric temporal logic rational. In 42nd International Symposium on Mathematical Foundations of Computer Science(LIPIcs, Vol. 83). Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 77:1–77:14. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  33. [33] Krishna Shankara Narayanan, Madnani Khushraj, and Pandya Paritosh K.. 2018. Logics meet 1-clock alternating timed automata. In 29th International Conference on Concurrency Theory(LIPIcs, Vol. 118). Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 39:1–39:17. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  34. [34] Lasota Slawomir and Walukiewicz Igor. 2008. Alternating timed automata. ACM Trans. Comput. Log. 9, 2 (2008), 10:1–10:27. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  35. [35] Madnani Khushraj Nanik. 2019. On Decidable Extensions of Metric Temporal Logic. Ph. D. Dissertation. Indian Institute of Technology Bombay, Mumbai, India.Google ScholarGoogle Scholar
  36. [36] Monica Dario Della, Gigante Nicola, Montanari Angelo, Sala Pietro, and Sciavicco Guido. 2017. Bounded timed propositional temporal logic with past captures timeline-based planning with bounded constraints. In 26th International Joint Conference on Artificial Intelligence. ijcai.org, 10081014. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  37. [37] Ouaknine J. and Worrell J.. 2005. On the decidability of metric temporal logic. In Annual Symposium on Logic in Computer Science. 188197.Google ScholarGoogle Scholar
  38. [38] Pandya Paritosh K.. 2001. Specifying and deciding quantified discrete-time duration calculus formulae using DCVALID. In RTTOOLS2001 Workshop. BRICS. Retrieved from https://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.30.8532.Google ScholarGoogle Scholar
  39. [39] Pandya Paritosh K. and Shah Simoni S.. 2011. On expressive powers of timed logics: Comparing boundedness, non-punctuality, and deterministic freezing. In 22nd International Conference on Concurrency Theory(Lecture Notes in Computer Science, Vol. 6901). Springer, 6075. DOI:Google ScholarGoogle ScholarCross RefCross Ref
  40. [40] Prabhakar Pavithra and D’Souza Deepak. 2006. On the expressiveness of MTL with past operators. In International Conference on Formal Modeling and Analysis of Timed Systems. 322336.Google ScholarGoogle ScholarDigital LibraryDigital Library
  41. [41] Rabinovich A.. 2008. Complexity of metric temporal logic with counting and Pnueli modalities. In International Conference on Formal Modeling and Analysis of Timed Systems. 93108.Google ScholarGoogle ScholarDigital LibraryDigital Library
  42. [42] Rabinovich Alexander. 2010. Complexity of metric temporal logics with counting and the Pnueli modalities. Theor. Comput. Sci. 411, 22-24 (2010), 23312342. DOI:Google ScholarGoogle ScholarDigital LibraryDigital Library
  43. [43] Raskin Jean Francois. 1999. Logics, Automata and Classical Theories for Deciding Real Time. Ph. D. Dissertation. Universite de Namur.Google ScholarGoogle Scholar
  44. [44] Trakhtenbrot B. A.. 1961. Finite automata and logic of monadic predicates. Doklady Akad. Nauk SSSR, in Russian. 98 (1961), 326329.Google ScholarGoogle Scholar
  45. [45] Vardi Moshe Y. and Wolper Pierre. 1994. Reasoning about infinite computations. Inf. Computat. 115, 1 (1994), 137.Google ScholarGoogle ScholarDigital LibraryDigital Library
  46. [46] Wilke Thomas. 1994. Specifying timed state sequences in powerful decidable logics and timed automata. In 3rd International Symposium on Formal Techniques in Real-Time and Fault-Tolerant Systems, Organized Jointly with the Working Group Provably Correct Systems. 694715. DOI:Google ScholarGoogle ScholarCross RefCross Ref

Index Terms

  1. From Non-punctuality to Non-adjacency: A Quest for Decidability of Timed Temporal Logics with Quantifiers

        Recommendations

        Comments

        Login options

        Check if you have access through your login credentials or your institution to get full access on this article.

        Sign in

        Full Access

        • Published in

          cover image Formal Aspects of Computing
          Formal Aspects of Computing  Volume 35, Issue 2
          June 2023
          187 pages
          ISSN:0934-5043
          EISSN:1433-299X
          DOI:10.1145/3605783
          Issue’s Table of Contents

          Permission to make digital or hard copies of part or all of this work for personal or classroom use is granted without fee provided that copies are not made or distributed for profit or commercial advantage and that copies bear this notice and the full citation on the first page. Copyrights for third-party components of this work must be honored. For all other uses, contact the owner/author(s).

          Publisher

          Association for Computing Machinery

          New York, NY, United States

          Publication History

          • Published: 23 June 2023
          • Online AM: 6 December 2022
          • Accepted: 31 October 2022
          • Revised: 21 September 2022
          • Received: 28 April 2022
          Published in fac Volume 35, Issue 2

          Permissions

          Request permissions about this article.

          Request Permissions

          Check for updates

          Qualifiers

          • research-article
        • Article Metrics

          • Downloads (Last 12 months)561
          • Downloads (Last 6 weeks)212

          Other Metrics

        PDF Format

        View or Download as a PDF file.

        PDF

        eReader

        View online with eReader.

        eReader