• Keine Ergebnisse gefunden

Topological Entropy of Formal Languages

N/A
N/A
Protected

Academic year: 2022

Aktie "Topological Entropy of Formal Languages"

Copied!
23
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

arXiv:1507.03393v1 [math.DS] 13 Jul 2015

Topological Entropy of Formal Languages

Friedrich Martin Schneider, Daniel Borchmann July 14, 2015

In this work we shall consider a notion of complexity of formal languagesLthat is inspired by the concept ofentropyfrom dynamical systems. More precisely, we shall define the topological entropyof L to be the exponential growth-rate of the restrictions of the Nerode congruence relation of L to words of length at mostn. We shall show that the topological entropy of regular languages is always zero, but that there are also non-regular languages with vanishing en- tropy, for example Dyck-languages. Furthermore, we shall establish a way of how to compute the entropy of a formal language that is given by a topological automaton accepting it. Finally, we shall point out that the topological entropy of a formal language can be seen as the entropic dimension of a suitable pre- compact pseudo-metric space.

1 Introduction

A variety of notions has been developed to assess different aspects of complexity of formal languages. Most of these notions have been devised with an understanding of complexity in mind that comes with classical complexity theory, and thus these notions are formu- lated as decision problems. Examples for this are the word problem and the equivalence problem for formal languages, and the complexity of the formal languages is measured by the complexity class for which these problems are complete. Other notions quantify complexity by other means. Examples are thestate complexity [18] of a regular language, which gives the complexity of the language as the number of states in its minimal automa- ton, or thesyntactic complexityof a regular language, which instead considers the size of corresponding syntactic semigroup [10].

The core idea of the present article is to expand the methods of measuring a formal lan- guage’s complexity by a topological approach in terms oftopological entropy, which proved tremendously useful to dynamical systems. Topological entropy was introduced by Adler et al. [1] for single homeomorphisms (or continuous transformations) on a compact Haus- dorff space. The literature provides several essentially different extensions of this concept for continuous group and semigroup actions. Among others, there is an approach towards topological entropy for continuous actions of finitely generated (pseudo-)groups due to Ghys et al. [13] (see also [3, 5, 7, 16]), which has also been investigated for continuous semigroup actions in [4, 6, 14].

(2)

By a dynamical system we mean a continuous semigroup action on a compact Hausdorff topological space. Topological entropy measures the ability of an observer to distinguish between points of the dynamical system just by recognizing transitionsat equal time inter- vals, i.e., with respect to a fixed generating system of transformations, starting from the initial state. Since the above notion of dynamical system may very well be regarded as the topological counterpart of a finite automaton, it seems natural to utilize the dynamical approach for applications to automata theory.

To view a formal language Lover an alphabet Σ as some kind of dynamical system we take inspiration from the characterization of regular languages as languages whose Myhill- Nerode congruence relationΘ(L)has finite index. Recall that foru,v∈Σwe have

(u,v)∈Θ(L) ⇐⇒ ∀wΣ: (uwL ⇐⇒ vwL).

The relationΘ(L)can be seen as some way of measuring the complexity ofL: ifLis regular, the number of equivalence classes is finite and equals the number of states in the minimal automaton ofL. This is the idea behind the notion of state complexity.

However, if L is not regular, Θ(L) does not have finite index, and asking for the state complexity ofLis not a reasonable undertaking. The notion oftopological entropy that we shall introduce in this paper tries to overcome this issue in the following way. Instead of considering the relation Θ(L)alone, we shall also take into account how Θ(L) arises as a limit of a sequences of certain equivalence relationsΘ(Σ(n),L). The behavior of this sequence(Θ(Σ(n),L)|nN)then gives rise to our notion of topological complexity ofL.

It is the purpose of the this paper to present a first investigation of the notion of topologi- cal entropy of formal languages. We shall show that all regular languages have vanishing topological entropy, but that there are also non-regular languages whose topological en- tropy is zero, most notably Dyck languages [2]. We shall also give examples of context-free languages with non-vanishing entropy. Furthermore, we shall discuss how the topologi- cal entropy of a formal language can be computed if the language itself is represented by means of a topological automaton [15].

This paper is structured as follows. We shall first introduce and investigate the topolog- ical entropy of formal languages in Section 2. In Section 3 we shall have a closer look into the connection of topological entropy of formal languages andtopological complexity of semigroup actions on compact Hausdorff spaces. In particular we shall show how the topological entropy of a language can be obtained if the language is given by a topological automaton accepting it. Finally, Section 4 shall show that the topological entropy coincides with the entropic dimension of a suitable precompact pseudo-ultrametric space.

2 Topological entropy of formal languages

In this section we shall introduce our new notion of topological entropy of formal lan- guages. As already sketched in the introduction, this notion is inspired by the characteri- zation of regular languages as those languages whose Nerode congruence has finite index.

To make our argumentation easier to follow, we shall thus first recall this famous result.

(3)

Thereafter, we shall introduce and investigate our notion of topological entropy. In par- ticular, we shall show that all regular languages have entropy zero, but that there are also non-regular languages with vanishing entropy. This latter discussion shall be embedded in a more general observation about languages defined via groups with sub-exponential growth. Finally, we shall discuss some more examples of languages and determine their entropy.

Let us recall some basic notation. LetΘbe an equivalence relation on a setY. ForyY we put [y]Θ := {xY | (x,y) ∈ Θ}. ThenY/Θ := {[y]Θ | yY}. Furthermore, theindex of Θ onY is defined as ind(Θ) := |Y/Θ|. For a mapping f: XY we set f1(Θ) := {(s,t) ∈ X×X | (f(x),f(y)) ∈ Θ}. Clearly, f1(Θ) then constitutes an equivalence relation onX.

Now let Σ be an alphabet, i.e., a finite and non-empty set. The Nerode congruence of a languageLΣis the equivalence relation

Θ(L):={(u,v)∈Σ×Σ | ∀wΣ: uwLvwL}.

Recall thatLis regular if and only if it is accepted by an automaton. The following charac- terization of regular languages in terms of the Nerode congruence relation is well-known.

Theorem 2.1 (Myhill-Nerode) LetΣbe a finite alphabet. A language L ⊆ Σ is regular if and only ifΘ(L)has finite index.

For regular languages Lthe number of equivalence classes of the Nerode congruence re- lation can thus be considered as a measure of complexity of the L. However, if L is not regular this measure is not available anymore. We shall remedy this by not considering thenumber of equivalence classes ofΘ(L), but by considering the growth of the number of equivalence classes of a particular approximation of Θ(L). Based on this growth we introduce our notion of topological entropy ofL.

Definition 2.2 LetΣbe an alphabet. For FΣfinite and LΣdefine

Θ(F,L):={(u,v)∈Σ×Σ | ∀wF: uwLvwL}. △ Now, the equivalence relationsΘ(F,L)constitute an approximation ofΘ(L)in the sense that

Θ(L) =\{Θ(F,L)|FΣfinite}. (1) Furthermore, it can be seen quite easily thatΘ(F,L)always has finite index. The mapping ΦF,L/Θ(F,L)→ {0, 1}Fgiven by

ΦF,L([u]Θ(F,L))(w):=

(1 ifuwL, 0 otherwise

is a well-defined injection. Because of this we have indΘ(F,L) ≤ 2|F|, and thus Θ(F,L) has finite index. Thus the following definition is reasonable.

(4)

Definition 2.3 LetΣbe an alphabet, and denote withF(Σ)the set of finite subsets ofΣ. Define γ:F(Σ)× P(Σ)→N, (F,L)7→ indΘ(F,L).

Given LΣ, we call

γL: F(Σ)→N, F 7→γ(F,L)

thetopological complexity function of L. The topological entropyof a language LΣ is defined to be

h(L):=lim sup

n

log2γL(Σ(n))

n ,

whereΣ(n)is the set of all words overΣof length at most n. △ Next we want to collect some obvious but useful properties of topological complexity func- tions.

Proposition 2.4 LetΣbe a finite alphabet, let E,FΣbe finite, and let L,L0,L1Σ. Then (a) γ(F,∅) =γ(F,Σ) =1, and thus h() =h(Σ) =0.

(b) γ(EF,L)≤γ(E,Lγ(F,L). If EF, thenγ(E,L)≤γ(F,L). (c) γ(F,L) =γ(F,Σ\L), and hence h(L) =h(Σ\L).

(d) γ(F,L0L1)≤γ(F,L0γ(F,L1), and thus h(L0L1)≤h(L0) +h(L1). (e) γ(F,L0L1)≤γ(F,L0γ(F,L1), and thus h(L0L1)≤h(L0) +h(L1).

Proof ClearlyΘ(F,∅) =Θ(F,Σ) =Σ×Σand henceγ(F,∅) =γ(F,Σ) =1. Secondly, because ofΘ(EF,L) = Θ(E,L)∩Θ(F,L)we obtain γ(EF,L) ≤ γ(E,Lγ(F,L). In particular, if EF, then Θ(F,L) ⊆ Θ(E,L) and hence γ(E,L) ≤ γ(F,L). Moreover, Θ(F,L) =Θ(F,Σ\L)and thereforeγ(F,L) =γ(F,Σ\L). Finally, it is easy to check that Θ(F,L0)∩Θ(F,L1)⊆ Θ(F,L0L1). Consequently,γ(F,L0L1)≤γ(F,L0γ(F,L1). Utilizing the previous observations, we can furthermore conclude that

Θ(F,L0)∩Θ(F,L1) =Θ(F,Σ\L0)∩Θ(F,Σ\L1)

Θ(F,(Σ\L0)∪(Σ\L1))

=Θ(F,Σ\(L0L1))

=Θ(F,L0L1)

and thereforeγ(F,L0L1)≤ γ(F,L0γ(F,L1). As a consequence of this proposition we immediately obtain the result that the class of lan- guages with zero entropy and the class of languages with finite entropy are closed under Boolean operations. Since it can be seen easily that finite languages have zero entropy we immediately obtain that all regular languages must have zero entropy as well.

The following result gives a precise formulation of this fact, and provides an alternative proof for the claim.

Theorem 2.5 LetΣbe an alphabet and LΣ. The following are equivalent:

(5)

(a) L is regular, (b) γLis bounded, and

(c) there exists some finite subset FΣsuch thatΘ(F,L) =Θ(L).

Proof (a) =⇒(b). Due to 2.1,Θ(L)has finite index. Note thatΘ(L)⊆ Θ(F,L)and hence γL(F)≤indΘ(L)for allFΣfinite. Thus,γLis bounded.

(b) =⇒ (c). Suppose thatγLis bounded. Then there exists some finiteF0Σ such that γL(F0) = sup{γL(F) | FΣfinite}. We shall show that Θ(F0,L) = Θ(L). Of course, Θ(L)⊆Θ(F0,L). Let(u,v)∈(Σ×Σ)\Θ(L). By (1) there exists some finiteF1Σsuch that(u,v) ∈/ Θ(F1,L). Obviously, F0F1Σ is finite andΘ(F0F1,L) ⊆ Θ(F0,L). By assumption,γL(F0F1) ≤ γL(F0). Consequently,Θ(F0F1,L) = Θ(F0,L)and therefore (u,v) ∈ (Σ ×Σ)\Θ(F1,L) ⊆ (Σ×Σ)\Θ(F0F1,L) = (Σ×Σ)\Θ(F0,L). This substantiates thatΘ(F0,L) =Θ(L).

(c) =⇒ (a). By assumptionΘ(L) =Θ(F,L), and since Θ(F,L)has finite index,Θ(L)has

finite index as well. Hence,Lis regular due to 2.1.

Corollary 2.6 LetΣbe an alphabet. If LΣis regular, then h(L) =0.

The converse of this corollary does not hold, i.e., there are non-regular languages with vanishing topological entropy. To see this we shall show thatDyck languagesalways have zero entropy (c.f. 2.12). We shall put the corresponding argumentation in a more general framework, by estimating the entropy of languages defined by groups. For this purpose, we recall the concept ofgrowthin groups. Consider a finitely generated groupG. LetSbe a finite symmetric generating subset ofGcontaining the neutral element. Theexponential growth rateofGwith respect toSis defined to be

egr(G,S):=lim sup

n

log2|Sn|

n .

Note that this quantity is finite as|Sn| ≤ |S|nfor everynN. Furthermore, egr(G,S):= lim

n

log2|Sn| n

due to a well-known result by Fekete [12]. Of course, the precise value of the exponential growth rate depends upon the particular choice of a generating set.

However, if T is another finite symmetric generating subset of G containing the neutral element, then

1

k ·egr(G,T)≤ egr(G,S)≤ l·egr(G,T)

wherek :=inf{mN\ {0} |TSm}andl :=inf{mN\ {0} |STm}. This justifies the following definition: Gis said to havesub-exponential growthif egr(G,S) = 0 for some (and thus any) symmetric generating setSofGcontaining the neutral element. The class of finitely generated groups with sub-exponential growth encompasses all finitely generated abelian groups. In fact, ifGis abelian, then

Snn

sS

sα(s)

α: S→ {0, . . . ,n}o

and thus|Sn| ≤(n+1)|S|for allnN. Now let us return to formal languages.

(6)

Theorem 2.7 LetΣbe an alphabet. Let G be a group, ϕ: ΣG a homomorphism, HG, and EG finite. Define

Pϕ(H):={wΣ | ∀u prefix of w: ϕ(u)∈ H}, Lϕ(H,E):=Pϕ(H)∩ϕ1(E).

Thenγ(F,Lϕ(H,E))≤ |E| · |ϕ(F)|+1for all finite FΣ. In particular, h(Lϕ(H,E))≤lim sup

n

log2|ϕ(Σ(n))|

n ≤log2|Σ|.

Furthermore, if S is a finite symmetric generating subset of G containing the neutral element and k:=inf{mN\ {0} | ϕ(Σ)⊆Sm}, then

h(Lϕ(H,E))≤ k·egr(G,S).

Proof We abbreviateP:= Pϕ(H)andL:= Lϕ(H,E). Consider a finite subsetFΣ. Then Q:= (F)1is a finite subset ofG. Fix any object∞∈/Qand defineQ := Q∪ {}. Let us consider the mapψ: ΣQgiven by

ψ(u):=

(ϕ(u) ifuPϕ1(Q),

∞ otherwise (uΣ).

We show kerψΘ(F,L). To this end, let(u,v)∈ kerψ. We proceed by case analysis.

First case: ψ(u) = ψ(v) 6= . Now,u,vPϕ1(Q)andϕ(u) = ψ(u) = ψ(v) = ϕ(v). LetwFand suppose thatuwL. We showvwL. We observe that

ϕ(vw) = ϕ(v)ϕ(w) =ϕ(u)ϕ(w) = ϕ(uw)∈E,

i.e.,vwϕ1(E). In order to prove that vwP, let x be a prefix of vw. If x is a prefix of v, then ϕ(x) ∈ H asvP. Otherwise, there exists a prefixy of w such that x = vy, and so we conclude that ϕ(x) = ϕ(vy) = ϕ(v)ϕ(y) = ϕ(u)ϕ(y) = ϕ(uy) ∈ H, because uwPanduyis a prefix ofuw. Hence,vwL. On account of symmetry, it follows that (u,v)∈ Θ(F,L).

Second case: ψ(u) = ψ(v) = . Let x ∈ {u, v}. If x ∈/ ϕ1(Q), then we conclude that ϕ(xw) = ϕ(x)ϕ(w)∈/ Eand thusxw ∈/ Lfor anywF. Ifx ∈/ P, thenxw ∈/ Pand hence xw ∈/ Lfor anywF. This proves that{uw, vw} ∩L = for all wF. Consequently, (u,v)∈ Θ(F,L).

This substantiates that kerψΘ(F,L). Therefore

γ(F,L) =indΘ(F,L)≤ind(kerψ)≤ |Q| ≤ |Q|+1≤ |E| · |ϕ(F)|+1.

In particular, it follows that h(L) =lim sup

n

log2γL(Σ(n))

n ≤lim sup

n

log2(|E| · |ϕ(Σ(n))|+1) n

=lim sup

n

log2|ϕ(Σ(n))|

n ≤lim sup

n

log2(n· |Σ|n)

n =log2|Σ|.

(7)

Finally, supposeS to be a finite symmetric generating subset ofGcontaining the neutral element. SinceΣis finite,M :={mN\ {0} | ϕ(Σ) ⊆Sm}is not empty. Letk := infM.

Our considerations above now readily imply that h(Lϕ(S,E))≤lim sup

n

log2|ϕ(Σ(n))|

nk·lim sup

n

log2|Sn|

n =k·egr(G,S). For groups whose growth is sub-exponential the previous theorem yields that the corre- sponding languagesLϕ(S,E)have zero entropy.

Corollary 2.8 LetΣbe an alphabet, let G be a group with sub-exponential growth, andϕ: ΣG a homomorphism. Then for each SG and finite EG, it is true that h(Lϕ(S,E)) =0.

We immediately obtain the following statement.

Corollary 2.9 LetΣbe an alphabet, let G be a finitely generated abelian group, andϕ: ΣG a homomorphism. Then for each SG and finite EG, it is true that h(Lϕ(S,E)) =0.

The following corollaries are immediate consequences of Theorem 2.7 forS= G.

Corollary 2.10 LetΣbe a finite alphabet and LΣ. Let G be a group, ϕ: ΣG a homo- morphism and EG finite such that L= ϕ1(E). Thenγ(F,L)≤ |E| · |ϕ(F)|+1for all finite FΣ. In particular,

h(L)≤lim sup

n

log2|ϕ(Σ(n))|

n ≤log2|Σ|.

Corollary 2.11 LetΣ be a finite alphabet, LΣ. Let G be an abelian group, ϕ: ΣG a homomorphism and EG finite such that L= ϕ1(E). Then h(L) =0.

With the previous results in place, we are now able to argue thatDyck languages have fi- nite entropy. Recall that theDyck language with k sorts of parenthesesconsists of all balanced strings over{(1,)1, . . . ,(k,)k}. Alternatively, we can view the Dyck language withksorts of parentheses as the set of all strings that can be reduced to the empty word by succes- sively eliminating matching pairs of parentheses.

We can formalize this as follows. LetΣ,Σbe two alphabets,∆ := ΣΣ, and letκ: Σ→ Σ be a bijection. Consider the the free group F(Σ) with generator set Σ, and denote with ϕ: ∆Σ the unique homomorphism satisfying ϕ(a) = a and ϕ(κ(a)) = a1 for all aΣ. Define

D(κ):= {w | ϕ(w) =e∧(∀uprefix ofw: |w|a ≥ |w|κ(a))}.

IfΣ = {(1, . . . ,(k},Σ = {)1, . . . ,)k}, andκ (i = )i, then the setD(κ)coincides with the Dyck language withksorts of parentheses.

(8)

Theorem 2.12 Letκ: Σ→Σbe a bijection between finite sets. Then log2|Σ| ≤h(D(κ))≤egr(F(Σ),S) for S:=ΣΣ1∪ {e}, where e denotes the neutral element of F(Σ).

Proof We first show indΘ(Σ(n),L)≥ |Σn|, since this implies log2|Σ| ≤h(D(κ)). For this let u,vΣn,u6= v. Defineκ(u):=κ(u|u|). . .κ(u1), whereu = u1. . .u|u|. Thenu·κ(u)∈ L, butv·κ(u)∈/ L. Thus(u,v)∈/Θ(Σ(n),L)and therefore indΘ(Σ(n),L)≥ |Σn|as required.

For the second inequality let us consider the unique homomorphismψ: F(Σ)→ZΣsatis- fying

ψ(b)(a):=

(1 ifa=b,

0 otherwise (a,bΣ).

We observeD(κ) =Lϕ(ψ1(NΣ),{e}), where the mapping ϕis as above. Hence, we have

h(D(κ))≤egr(F(Σ),S)by 2.9.

The reason that Dyck languages with more than one type of parentheses have non-zero positive entropy is cause by the requirement that in a wordwthe different types of paren- theses need to mutually balanced, i.e.,ϕ(w) =e. In other words, if we replace this require- ment by the weaker condition that each opening parenthesis has to be closed eventually, then we obtain a class of languages with zero entropy.

Theorem 2.13 Letκ: Σ → Σbe a bijection between finite sets, let∆ := ΣΣ, and consider the language

D(κ):={w | ∀aΣ: (|w|a =|w|κ(a))∧(∀u prefix of w:|w|a ≥ |w|κ(a))}. Then h(D(κ)) =0.

Proof Let us consider the homomorphismϕ: ∆ZΣgiven by ϕ(w)(a):=|w|a− |w|κ(a) (w, aΣ).

We observe thatD(κ) =Lϕ(NΣ,{0}), whereforeh(D(κ)) =0 by 2.9.

Note that for|Σ|=1 we haveD(κ) =D(κ) =0, and thus we obtain an example of a lan- guage with zero entropy that is not regular. Other non-regular languages with vanishing entropy are discussed in the following examples.

Example 2.14 LetΣbe an alphabet.

(a) Let mN and a,bΣ, a 6= b. Then L := {wΣ | |w|a = |w|b+m} is not regular. However, h(L) = 0 by Corollary 2.11. To see this, note that the mapping ϕZ, w7→ |w|a− |w|bconstitutes a homomorphism whereL= ϕ1({m}).

(9)

(b) Suppose that Σ = {a,b,c}. Then L := {anbncn | nN}is not context-free, but h(L) =0. To see this we show thatΘ=Θ(Σ(n),L)has the equivalence classes

[ck]Θ, kn/2

[bck]Θ, 1≤ ℓ≤ k, 2k−ℓ≤n [abkck]Θ, 1≤ ℓ≤ k, k−ℓ≤n

[b]Θ.

(2)

From this it follows that indΘ(Σ(n),L)∈ O(n2), and thush(L) =0.

To see that the sets in (2) are indeed all equivalence classes ofΘ(Σ(n),L), letwΣ such thatwis not an element of the first three types of classes in (2). We need to show that thenw ∈ [b]Θ(Σ(n),L). We do this by showing that there is no uΣ(n)such that uwL.

Assume by contradiction that such a word u exists. Then u must be of one of the following forms

u= akbkc, 2k+ℓ≤n, 0≤ℓ≤k, u= akb, k+ℓ≤n, 0≤ℓ <k, u= a, 0 ≤ℓ <n

Ifu = akbkc, 2k+ℓ ≤ n, 0 ≤ ℓ ≤ k, thenw = ck, k−ℓ ≤ n/2, and therefore w∈[ck]Θ(Σ(n),L), a contradiction. Ifu= akb,k+ℓ≤ n, 0≤ ℓ <k, thenw=bkck, andk−ℓ > 0, 2k−(k−ℓ) ≤ n, thus w ∈ [bkck]Θ(Σ(n),L), again a contradiction.

Ifu = a, thenw = akbkck, and k−(k−ℓ) ≤ n, so w ∈ [abkck]Θ(Σ(n),L), another contradiction.

Thus, our assumption thatuexists is false. The same is true for the wordb, and thus

w∈[b]Θ(Σ(n),L), as required.

The following example shows that there are natural examples of “simple” languages whose entropy is not zero, but still finite.

Example 2.15 Suppose|Σ| ≥2. Then thepalindrome language L:={wwR |wΣ} is not regular, but context-free, andh(L)∈(0,∞).

To seeh(L) >0, observe that for each nNand allu,vΣn, if(u,v)∈ Θ(Σ(n),L), then u=v. This is because ofvvRL, we also haveuvRL, and henceu=v. Thus

[u]Θ(Σ(n),L)6= [v]Θ(Σ(n),L) (u6= v) Thus indΘ(Σ(n),L)≥ |Σn|= |Σ|n, and we obtain

h(L) =lim sup

n

log2|Σ|n

n =log2|Σ|>0.

(10)

To seeh(L)<we shall consider the relationΘ defined by

(u,v)∈Θ ⇐⇒ (u,v)∈Θ(Σ(n),L)and(|u| ≤n ⇐⇒ |v| ≤n). Then indΘ(Σ(n),L)≤indΘ. We shall show

lim sup

n

log2(indΘ) n <.

There are at most|Σ(n)|many equivalence classes[u]Θ foruΣ,|u| < n. To count the number of equivalence classes for|u| ≥ nwe define

n(u):={a1. . .ai |1≤in,a1, . . . ,aiΣ,u=a1. . .aiu,uL}. Then foru,vΣ\Σ(n)we have

(u,v)∈Θ ⇐⇒ (u,v)∈Θ(Σ(n),L) ⇐⇒ ℓn(u) =ℓn(v). (3) The first equivalence is clear. To see the second equivalence let(u,v)∈ Θ(Σ(n),L), and let a1. . .ai ∈ℓn(u). By definition ofℓn(u)it is then true that

u(a1. . .ai)RL.

Because(u,v)∈Θ(Σ(n),L)we therefore obtain v(a1. . .ai)RL,

i.e., v is of the form v = a1. . .aiv for some vL. This yields a1. . .ai ∈ ℓn(v). By symmetry we obtainℓn(u) =ℓn(v)as required.

Conversely, assumeℓn(u) =ℓn(v), and letwΣ(n)be such thatuwL. Because|u| ≥ n, there existsuLwithuw = wRuw. ThenwR ∈ ℓn(u) = ℓn(v), and thereforev = wRv for somevL. But thenvwL. By symmetryvwL =⇒ uwLfor eachwΣ(n), and therefore(u,v)∈ Θ(Σ(n),L), as required.

Using the characterization from Equation (3) we have

Σ\Σ(n)Θ

={ℓn(u)|u∈Σ(n)}

Now every setℓn(u)withu =u1. . .uk,kn, can be represented by the prefixu1. . .unof uof lengthntogether with a tuplet∈ {0, 1}ndefined by

ti =1 ⇐⇒ u1. . .ui ∈ ℓn(u). Therefore,

Σ\Σ(n)Θ

={ℓn(u)|uΣ\Σ(n)} ≤ |Σ|n·2n.

(11)

This yields

indΘ =Σ(n)Θ

+Σ\Σ(n)Θ

≤ |Σ(n)|+|Σ|n·2n, and thus

lim sup

n

log2(indΘ)

n =log2(2|Σ|)<∞.

It is not hard to see that the entropy of a formal language can very well be infinite. This is illustrated by the following example

Example 2.16 Let|Σ| ≥ 2, and choose mappingsϕn: Σ2n → P(Σn)for each nNsuch that|im(ϕn)|=|Σ|2n =22n. Then define a languageLΣ by

LΣm :=

({uv|uΣ2n,vϕn(u)} ifm=2n+nfor somenN,

∅ otherwise.

Then 22nγL(n), i.e.,

22n ≤indΘ(Σn,L). (4)

To see this we shall show that each wordϕn(u)defines its own equivalence class, i.e., for wordsu0,u1Σ2n with ϕn(u0) 6= ϕn(u1) we have (u0,u1) ∈/ Θ(Σn,L). This is because if ϕn(u0)6= ϕn(u1)we can assume without loss of generality that there exists some word vϕn(u0)\ϕn(u1). By definition ofLwe then haveu0vL, but since|u1v|=2n+nand v∈/ ϕn(u1)we also getu1v ∈/L. Thus(u0,u1)∈/ Θ(Σn,L).

But then (4) implies

lim sup

n

log2γL(n)

n ≥lim sup

n

log222n n =,

and thush(L) =.

3 Entropy of semigroup actions and topological automata

Regular languages are exactly those accepted by finite automata. For some non-regular languages, there exist similar characterizations in terms of finite state machines, e.g., push- down automata, linearly bounded Turing machines, and Turing machines. However, there is a different approach in terms oftopological automata[15], which in contrast to finite au- tomata have an infinite state set which itself is equipped with a compact Hausdorff topol- ogy. In this case,everylanguageLis accepted by a topological automaton, and one can ask whether the topological entropy ofLcan be expressed in terms of a topological automaton accepting it. In this section we shall show that this question has a positive answer.

The concept of topological automata arises from the observation that the transition func- tion of a finite automata is amonoid actionon the setΣof all words overΣ. Recall that an actionof a monoidSon a setXis a mappingα: X×SXsuch that

α(x,eS) =x,

α(x,st) =α(α(x,s),t)

(12)

for allxX,s,tS.

Recall that a (deterministic) automaton over an alphabet Σ is a tuple A = (Q,Σ,δ,q0,F) consisting of a finite setQofstates, atransition functionδ: Q×ΣQ, a setFQoffinal states, and aninitial state q0Q. The transition function is usually extended to the set of all words overΣby virtue of

δ(q,ε):=q,

δ(q,wa):=δ(δ(q,w),a)

forqQ,aΣ, andwΣ. It is not difficult to see that in this caseδis a monoid action ofΣonQ. Thelanguage acceptedbyAis then

L(A):={wΣ |δ(q0,w)∈ F}.

We can extend the notion of deterministic finite automata to an infinite state set as fol- lows [15].

Definition 3.1 Atopological automaton over an alphabetΣis a tuple A = (X,Σ,α,x0,F) consisting of

a compact Hausdorff space X, called the set of statesofA

a continuous actionαofΣon X, called thetransition functionofA,

a point x0X, called theinitial stateofA, and

a clopen subset FX, called the set of final statesofA.

We say thatAistrimifα(x0)is dense in X. Thelanguage recognized byAis defined as L(A):= {wΣ |α(x0,w)∈F}.

LetB= (Y,Σ,β,y0,G)be another topological automaton. We shall say thatAandBareisomor- phic, and writeA ∼= B, if there exists a homeomorphism ϕ: XY such that

ϕ(α(x,σ)) =β(ϕ(x),σ)

for all xX,σΣ(x0) =y0, andϕ(F) =G. △ Evidently, isomorphic automata accept the same language.

Observe that every automaton acceptingLcan be turned into an automaton that is trim: if A= (X,Σ,α,x0,F)is a topological automaton acceptingL, then replacingXwithα(x0) andFwithFα(x0)always yields a trim automaton accepting the same languageL.

As already stated, and in contrast to regular languages, every formal languageLΣis accepted by a topological automaton, cf. [15, Proposition 2.1].

Proposition 3.2 Let LΣ andχL the characteristic function of L. Equip X := {0, 1}Σ with the product topology, and define the mappingδ: X×ΣX by

δ(f,u)(v):= f(uv).

Then L is accepted by the topological automaton(X,Σ,δ,χL,T)for T :={fX | f(ε) =1}.

(13)

With the notation of 3.2, we define theminimal automaton of Lto be AL= χL(Σ),Σ,δ,χL,TL

,

whereχL(Σ) is the closure ofχL(Σ) in{0, 1}Σ, and TL = TχL(Σ). Clearly, AL is trim. Indeed we have the following fact, cf. [15, Theorem 2.2].

Proposition 3.3 Let LΣ, and letA = (X,Σ,x0,δ,F)be a topological automaton accepting L. ThenA ∼=ALif and only if for every automatonB = (Y,Σ,y0,λ,G)accepting L there exists a uniquely determined surjective continuous function

ϕ:YX

satisfying ϕ(λ(y,σ)) =δ(ϕ(y),σ), ϕ(y0) =x0, and F= ϕ1(G).

SinceAL∼= AL, this proposition immediately yields that the minimal automaton is indeed minimal in the above sense. Moreover, in the case thatLis regular,AL is finite and is the usual minimal automaton of regular languages.

Example 3.4 Let Σbe a finite alphabet and let a,b ∈ Σ, a 6= b. We consider theAlexan- droff compactificationZ of the discrete space of integersZ, that is the setZ = Z∪ {} equipped with the topology

{MZ∪ {} | M=⇒Z\Mis finite}.

We define an action α of Σ onZ by setting α(m,a) = m+1, α(m,b) = m−1 and α(m,c) =mfor allmZandcΣ\ {a,b}. Thenαconstitutes a continuous action ofΣ onZ, and for eachnNthe topological automatonA = (Z,Σ,α, 0,{n})accepts the

languageL=wΣ |w|a =|w|b+n .

We now shall express the topological complexity of the languageLaccepted by a topologi- cal automatonA= (Q,Σ,α,x0,F)by thetopological entropyof the continuous actionαofΣ onQ[1, 8, 17]. To this end, we shall first fix some useful notation and recall some important definitions about continuous actions on compact Hausdorff spaces.

LetXagain be a compact Hausdorff space. We shall denote byC(X)the set of all finite open covers ofX. If f: XXis continuous andU ∈ C(X), then f1(U):= {f1(U)|U∈ U } is a finite open cover ofXas well. Given U,V ∈ C(X), we say thatV refinesU and write U V if

V∈ V ∃U ∈ U: VU,

and we say thatU andV arerefinement-equivalentand writeU ≡ V ifU V andV U. Furthermore, if(Ui |iI)is a finite family of finite open covers ofX, then

_

iI

Ui :=\

iI

Ui

(Ui)iI

iI

Ui .

(14)

is a finite open cover ofXas well. ForU ∈ C(X)let N(U):=inf

|V | V ⊆ U,X =[V .

In preparation for some later considerations, let us recall the following basic observations.

Remark 3.5 ([1]) Let X be a compact Hausdorff space, U,V ∈ C(X), I be a finite set, (Ui)iI,(Vi)iI ∈ C(X)I, and f: XX be a continuous map. Then the following state- ments hold:

(1) U V =⇒ N(U)≤N(V), (2) U V =⇒ f1(U) f1(V),

(3) (∀iI: Ui Vi) =⇒ WiIUi WiIVi.

Now we come to dynamical systems, i.e., continuous semigroup actions. LetSbe a semi- group and consider a continuous actionα: X×SXofSonX, that is,αis supposed to be an action ofSonXwhereαs: XX, αs(x) =α(x,s)is continuous for everysS. For U ∈ C(X)we write

s1(U):=αs1(U). For every finiteFSandU ∈ C(X)let

(F:U)α := N _

sF

s1(U).

AssumeF to be a finite generating subset ofS. IfU is a finite open cover of X, then we define

η(α,F,U):=lim sup

n

log2(Fn:U)α

n .

Furthermore, thetopological entropy ofαwith respect to Fis defined to be the quantity η(α,F):=sup{η(α,F,U)| U ∈ C(X)}.

Of course, the precise value of this quantity depends on the choice of a finite generating system. However, we observe the following fact.

Proposition 3.6 Let S be a semigroup and let α be a continuous action of S on some compact Hausdorff space X. Suppose E,FS to be finite subsets generating S. Then

1

m·η(α,F)≤η(α,E)≤n·η(α,F), where m:=inf{kN|FEk}and n:=inf{kN|EFk}.

Proof LetU ∈ C(X). Evidently,(Ek :U)≤ (Fkn :U)for allkN, whence η(α,E,U) =lim sup

k

log2(Ek :U)α

k ≤lim sup

k

log2(Fkn :U)α

k

=nlim sup

k

log2(Fkn :U)α

knnlim sup

k

log2(Fk :U)α

k =n·η(α,F,U)

(15)

Thus,η(α,E,U)≤ n·η(α,F,U). This shows thatη(α,E) ≤ (α,F). Due to symmetry, it

follows thatη(α,F)≤m·η(α,E)as well.

We shall now show that the entropy of a formal language is bounded from above by the entropy of any topological automaton accepting it. In the case that the automaton is trim, these two notions even coincide.

Theorem 3.7 SupposeA= (X,Σ,α,x0,F)to be a topological automaton. Consider S:= Σ∪ {ε} andU := {F, X\F}. Then h(L(A))≤η(α,S,U). IfAis trim, then h(L(A)) =η(α,S,U). We prove this theorem with the following three auxiliary statements.

Lemma 3.8 LetA= (X,Σ,α,x0,F)be a topological automaton. LetΦ: ΣX, w7→α(x,w) andU := {F, X\F}. Consider a finite subset EΣas well as the equivalence relation

ΛE :={(x,y)| ∀wE: α(x,w)∈ F⇐⇒α(y,w)∈ F}. Then the following statements hold:

(1) X/ΛE = WwEw1(U)\ {}. (2) Θ(E,L(A)) = (Φ×Φ)1(ΛE).

(3) IfAis trim, thenΦ(Σ)∩V 6= for every VX/ΛE.

Proof (1): We observe that V := WwEw1(U)\ {}constitutes a finite partition of X into clopen subsets. For anyV ∈ VandxV, we observe that

[x]ΛE = {yY| ∀wE: α(x,w)∈Fα(y,w)∈F}

= {yY| ∀wE: xw1(F)⇔yw1(F)}

= {yY| ∀wEU∈ U : xw1(U)⇔yw1(U)}

= {yY| ∀W ∈ V: xWyW}

= {yY|yV}

=V We conclude thatX/ΛE =V.

(2): LetL:=L(A). For any two wordsu,vΣ, it follows that (u,v)∈Θ(E,L) ⇐⇒ ∀wE: uwLvwL

⇐⇒ ∀wE: α(x0,uw)∈ Fα(x0,vw)∈ F

⇐⇒ ∀wE: α(α(x0,u),w)∈ Fα(α(x0,v),w)∈F

⇐⇒ ∀wE: α(Φ(u),w)∈Fα(Φ(v),w)∈ F

⇐⇒ (Φ(u),Φ(v))∈ΛE. That is,Θ(E,L) = (Φ×Φ)1(ΛE).

(3): By (1), the setX/ΛE is a collection of open, non-empty subsets ofX. IfAis trim, then Φ(Σ)is dense inX, and thusΦ(Σ)∩V6=∅for everyVX/ΛE.

Referenzen

ÄHNLICHE DOKUMENTE

In section 2 we calculate the topological pseudo entropy in various setups of a three-dimensional Chern-Simons gauge theory and provide its interpretations in the light of

Since compactness plays an important role in the theory of bounded sets, we will start this chapter by recalling some basic definitions and properties of compact subsets of a

Given an open subset Ω of R d with the euclidean topology, the space C(Ω) of real valued continuous functions on Ω with the so-called topology of uniform convergence on compact sets

3.2 Connection between local compactness and finite dimensionality 39 4 Locally convex topological vector spaces 41 4.1 Definition by

Indeed, starting from a base of neighbourhoods of X, we can define a topology on X by setting that a set is open i↵ whenever it contains a point it also contains a basic

Definition 4.5.1. Let X be an infinite dimensional vector space whose di- mension is countable. Let X be an infinite dimensional vector space whose dimen- sion is countable endowed

Until the definition of scale function in such groups given by George Willis, almost the only structure known was a theorem of van Dantzig, namely that a locally compact

TOPOLOGI CAL COMPLEXI TY OF MANI FOLDS OF PREFERENCES*. GRACI ELA CHI CHI