On the order of lazy cellular automata (original) (raw)
Edgar Alcalá-Arroyo111Email: edgar.alcala7434@alumnos.udg.mx Alonso Castillo-Ramirez222Email: alonso.castillor@academicos.udg.mx Centro Universitario de Ciencias Exactas e Ingenierías, Universidad de Guadalajara, México.
Abstract
We study the most elementary family of cellular automata defined over an arbitrary group universe GG and an alphabet AA: the lazy cellular automata, which act as the identity on configurations in AGA^{G}, except when they read a unique active transition p∈ASp\in A^{S}, in which case they write a fixed symbol a∈Aa\in A. As expected, the dynamical behavior of lazy cellular automata is relatively simple, yet subtle questions arise since they completely depend on the choice of pp and aa. In this paper, we investigate the order of a lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G}, defined as the cardinality of the set {τk:k∈ℕ}\{\tau^{k}:k\in\mathbb{N}\}. In particular, we establish a general upper bound for the order of τ\tau in terms of pp and aa, and we prove that this bound is attained when pp is a quasi-constant pattern.
keywords: Lazy cellular automaton; unique active transition; order of a cellular automaton; quasi-constant pattern.
1 Introduction
A cellular automaton (CA) is a mathematical model over a discrete space defined by a local map that is applied homogeneously and simultaneously to the whole space. The underlying discrete space is a configuration space AGA^{G}, which consists of all maps from a group universe GG to an alphabet AA. Following [6], a cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} is a function such that there exist a finite subset S⊆GS\subseteq G, called a neighborhood of τ\tau, and a local map μ:AS→A\mu:A^{S}\to A satisfying
| τ(x)(g)=μ((g⋅x)|S),∀x∈AG,g∈G,\tau(x)(g)=\mu((g\cdot x)|_{S}),\quad\forall\ x\in A^{G},g\in G, | | ----------------------------------------------------------------------------------------------------------- |
where ⋅\cdot denotes the shift action of GG on AGA^{G}:
(g⋅x)(h):=x(hg),∀h∈G.(g\cdot x)(h):=x(hg),\quad\forall h\in G. |
---|
We say that a cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} is lazy if there is a local defining map μ:AS→A\mu:A^{S}\to A for τ\tau such that e∈Se\in S, where ee is the identity of the group GG, and there exists a pattern p∈ASp\in A^{S}, called the unique active transition of τ\tau, satisfying the following:
∀z∈AS,μ(z)=z(e)⟺z≠p.\forall z\in A^{S},\quad\mu(z)=z(e)\Longleftrightarrow z\neq p. |
---|
Intuitively, a lazy cellular automaton acts almost as the identity function of AGA^{G}, except when it reads the pattern pp, in which case it writes the symbol a:=μ(p)∈A∖{p(e)}a:=\mu(p)\in A\setminus\{p(e)\}.
As expected, the dynamical behavior of a lazy CA is relatively simple, yet subtle questions arise because their evolution depends entirely on the choice of pp and aa. In a certain sense, lazy cellular automata are even more elementary than the well-known elementary cellular automata (ECA) studied by Wolfram [14], where complex behavior already emerges. Since there are only 256 ECA, a complete case-by-case computational analysis is feasible, whereas there are infinitely many lazy CAs (over an infinite group universe), as their neighborhood size can be arbitrarily large. Despite this, we believe that a deep understanding of lazy CAs is possible and may lead to new insights into broader families of cellular automata.
Lazy cellular automata were introduced in [3] as a tool to study idempotent CAs: this is, cellular automata τ:AG→AG\tau:A^{G}\to A^{G} satisfying τ2=τ\tau^{2}=\tau. It was observed that if the unique active transition p∈ASp\in A^{S} is constant (i.e., p(e)=p(s)p(e)=p(s), ∀s∈S\forall s\in S) or symmetric (i.e., S=S−1S=S^{-1} and p(s)=p(s−1)p(s)=p(s^{-1}), ∀s∈S\forall s\in S), then τ\tau is idempotent. Moreover, the idempotence of τ\tau was completely characterized when pp is quasi-constant, meaning that there is r∈Sr\in S such that p|S∖{r}p|_{S\setminus\{r\}} is constant.
Here we study the order of a lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G}, denoted by ord(τ)\textup{ord}\left(\tau\right), as the cardinality of the set of all powers of τ\tau:
| ord(τ):=|{τk:k∈ℕ}|.\textup{ord}\left(\tau\right):=|\{\tau^{k}:k\in\mathbb{N}\}|. | | ------------------------------------------------------------------------------------------- |
Besides being an important algebraic concept, the order also captures part of the dynamical behaviour of a cellular automaton. For example, it was shown by Kůrka [12, Theorem 4] that a one-dimensional cellular automaton (i.e, when the underlying universe is the group of integers ℤ\mathbb{Z}) is of finite order if and only if it is equicontinuous.
The dynamical behavior of one-dimensional lazy CAs was examined in [4]. In this case, when the neighborhood S⊆ℤS\subseteq\mathbb{Z} of τ\tau is an interval, an interesting dichotomy arises: either τ\tau is idempotent or of infinite order, which is equivalent to being strictly almost equicontinuous as a dynamical system.
In this paper, we establish several results on the order of lazy CAs in the general setting of an arbitrary group universe GG. In Section 2, we show that for any lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G}, if ord(τ)\textup{ord}\left(\tau\right) is finite, then τ\tau has period 11, which means that there is n∈ℕn\in\mathbb{N} such that τn=τn+1\tau^{n}=\tau^{n+1}. Moreover, in Theorem 2, we derive a general upper bound for ord(τ)\textup{ord}\left(\tau\right) in terms of its unique active transition p∈ASp\in A^{S}. As a consequence, in Corollary 3, we provide a sufficient condition on p∈ASp\in A^{S} that guarantees that τ\tau is idempotent.
In Section 3, we determine the order of a lazy CA whose unique active transition is a quasi-constant pattern, which is a substantial generalization of Theorem 2 in [3]. In order to state our result, we introduce the following notation about words on groups. A word on S⊆GS\subseteq G of length nn is simply an element of the Cartesian power Sn:={(s1,…,sn):si∈S}S^{n}:=\{(s_{1},\dots,s_{n}):s_{i}\in S\}. We consider the evaluation function θ\theta from words on S⊆GS\subseteq G to elements of GG given by θ(s1,…,sn):=s1…sn\theta(s_{1},\dots,s_{n}):=s_{1}\dots s_{n}. We say that vv is a subword of w=(s1,…,sn)w=(s_{1},\dots,s_{n}), denoted by v⊑wv\sqsubseteq w, if v=(si,si+1,…,sj)v=(s_{i},s_{i+1},\dots,s_{j}) for some i≤ji\leq j.
Theorem 1.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy cellular automaton with unique active transition p∈ASp\in A^{S} and writing symbol a∈AS∖{p(e)}a\in A^{S}\setminus\{p(e)\}. Assume that pp is quasi-constant with non-constant element r∈Sr\in S.
- If a≠p(s)a\neq p(s) for all s∈Ss\in S, then ord(τ)=2\textup{ord}\left(\tau\right)=2.
- If r≠er\neq e and a=p(r)a=p(r), then ord(τ)\textup{ord}\left(\tau\right) is finite if and only if there exists n≥2n\geq 2 such that rn∈Sr^{n}\in S. Moreover, in this case,
ord(τ)=min{n≥2:rn∈S}.\textup{ord}\left(\tau\right)=\min\{n\geq 2:r^{n}\in S\}.
- If r≠er\neq e and a=p(r)a=p(r), then ord(τ)\textup{ord}\left(\tau\right) is finite if and only if there exists n≥2n\geq 2 such that rn∈Sr^{n}\in S. Moreover, in this case,
- If r=er=e and a=p(s)a=p(s) for all s∈S∖{e}s\in S\setminus\{e\}, then ord(τ)\textup{ord}\left(\tau\right) is finite if and only if there exists n≥2n\geq 2 such that for all words w∈(S∖{e})n−1w\in(S\setminus\{e\})^{n-1} there exists a subword v⊑wv\sqsubseteq w such that θ(v)−1∈S\theta(v)^{-1}\in S. In such case, the order of τ\tau is the minimum nn satisfying this property.
As an easy consequence of Theorem 1 we find that, in contrast with the dichotomy obtained in [4] when the neighborhood S⊆ℤS\subseteq\mathbb{Z} is an interval, for every n≥2n\geq 2, there exists a lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} such that ord(τ)=n\textup{ord}\left(\tau\right)=n.
Finally, in Section 3, we present two open problems related to the study of lazy cellular automata.
2 The order of lazy CA
We assume that the alphabet AA has at least two different elements 0 and 11. A pattern is a function p∈ASp\in A^{S}, where SS is a finite subset of the group universe GG.
The term active transitions of a local map μ:AS→A\mu:A^{S}\to A, with e∈Se\in S, has been recently introduced by several authors [1, 2, 8, 9] as the patterns z∈ASz\in A^{S} such that μ(z)≠z(e)\mu(z)\neq z(e). The activity value [7] of μ:AS→A\mu:A^{S}\to A is simply the number of active transitions of μ\mu.
Definition 1.
A cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} is called lazy if there is a local defining map μ:AS→A\mu:A^{S}\to A for τ\tau such that e∈Se\in S and there exists p∈ASp\in A^{S} satisfying
∀z∈AS,μ(z)=z(e)⟺z≠p.\forall z\in A^{S},\quad\mu(z)=z(e)\Longleftrightarrow z\neq p. |
---|
In such case, we say that pp is the unique active transition of τ\tau and μ(p)∈A∖{p(e)}\mu(p)\in A\setminus\{p(e)\} is the writing symbol of τ\tau.
The adjective “lazy” is justified for this class of cellular automata as they have the smallest non-zero activity value (i.e., activity value 11).
Example 1.
Let G=ℤG=\mathbb{Z} and A={0,1}A=\{0,1\}. The ECA rule 236 (see [11, Sec. 2.5] for an explanation of the rule labeling) is lazy, because it may be defined via the local map μ:A{−1,0,1}→A\mu:A^{\{-1,0,1\}}\to A given by the following table:
z∈A{−1,0,1}111110101100011010001000μ(z)∈A11101100\begin{array}[]{c|cccccccc}z\in A^{\{-1,0,1\}}&111&110&101&100&011&010&001&000\\ \hline\cr\mu(z)\in A&1&1&1&0&1&1&0&0\end{array} |
---|
The unique active transition of rule 236 is p=101∈A{−1,0,1}p=101\in A^{\{-1,0,1\}}.
Example 2.
Let G=ℤG=\mathbb{Z} and A={0,1}A=\{0,1\}. The ECA rule 136 may be defined via the local map μ:A{−1,0,1}→A\mu:A^{\{-1,0,1\}}\to A given by the following table
z∈A{−1,0,1}111110101100011010001000μ(z)∈A10001000\begin{array}[]{c|cccccccc}z\in A^{\{-1,0,1\}}&111&110&101&100&011&010&001&000\\ \hline\cr\mu(z)\in A&1&0&0&0&1&0&0&0\end{array} |
---|
It may seem that rule 136 is not lazy because the above local map has two active transitions 110110 and 010010. However, this local map may be reduced to its minimal neighborhood, so we obtain a local defining map μ′:A{0,1}→A\mu^{\prime}:A^{\{0,1\}}\to A for rule 136 given by the following table:
z∈A{0,1}11100100μ′(z)∈A1000\begin{array}[]{c|cccccccc}z\in A^{\{0,1\}}&11&10&01&00\\ \hline\cr\mu^{\prime}(z)\in A&1&0&0&0\end{array} |
---|
Therefore, rule 136 is lazy and has unique active transition p=10∈A{0,1}p=10\in A^{\{0,1\}}.
Recall that the minimal neighborhood of a cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} is a neighborhood admitted by τ\tau of smallest cardinality, which always exists and is unique by [6, Proposition 1.5.2]. Equivalently, the minimal neighborhood of τ\tau is equal to the set of all elements of GG that are essential in order to define a local map for τ\tau (see [5, Proposition 1]). We call the minimal local map of τ\tau to the local defining map of τ\tau associated to its minimal neighborhood.
It was shown in [3, Lemma 1] that if μ:AS→A\mu:A^{S}\to A has a unique active transition and |S|≥2|S|\geq 2, then the minimal neighborhood of the lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} defined by μ\mu is precisely SS. Hence, the next result follows.
Proposition 1.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a non-constant cellular automaton with minimal local map μ:AS→A\mu:A^{S}\to A. Then, τ\tau is lazy if and only if e∈Se\in S and μ\mu has a unique active transition.
We say that a pattern p∈ASp\in A^{S} appears in x∈AGx\in A^{G} if there is g∈Gg\in G such that (g⋅x)|S=p(g\cdot x)|_{S}=p.
Lemma 1.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA with unique active transition p∈ASp\in A^{S}. Then, pp appears in a configuration x∈AGx\in A^{G} if and only if x≠τ(x)x\neq\tau(x).
Proof.
Let μ:AS→A\mu:A^{S}\to A be the minimal local map of τ\tau. By Definition 1, there is g∈Gg\in G such that (g⋅x)|S=p(g\cdot x)|_{S}=p if and only if
| τ(x)(g)=μ((g⋅x)|S)≠p(e)=(g⋅x)(e)=x(g).\tau(x)(g)=\mu((g\cdot x)|_{S})\neq p(e)=(g\cdot x)(e)=x(g). | | ------------------------------------------------------------------------------------------------------------------ |
The result follows. ∎
For any cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} and n∈ℕn\in\mathbb{N}, we denote by τn\tau^{n} the nn-th composition of τ\tau with itself:
τn=τ∘⋯∘τ⏟n times .\tau^{n}=\underbrace{\tau\circ\dots\circ\tau}_{n\text{ times }}. |
---|
By [6, Prop. 1.4.9], τn:AG→AG\tau^{n}:A^{G}\to A^{G} is also a cellular automaton.
Corollary 1.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA with unique active transition p∈ASp\in A^{S}. For any n∈ℕn\in\mathbb{N}, τn≠τn+1\tau^{n}\neq\tau^{n+1} if and only if there exists x∈AGx\in A^{G} such that pp appears in τn(x)\tau^{n}(x).
The next result follows by [3, Lemma 3] or [4, Lemma 3.2], but we add its proof here for completeness.
Lemma 2.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA with unique active transition p∈ASp\in A^{S} and writing symbol a∈Aa\in A. If there exists x∈AGx\in A^{G} such that τ(x)|S=p\tau(x)|_{S}=p, then (s⋅x)|S=p(s\cdot x)|_{S}=p for some s∈S∖{e}s\in S\setminus\{e\} such that p(s)=a≠p(e)p(s)=a\neq p(e).
Proof.
The hypothesis τ(x)|S=p\tau(x)|_{S}=p is equivalent to
| p(s)=τ(x)(s)=μ((s⋅x)|S),∀s∈S,p(s)=\tau(x)(s)=\mu((s\cdot x)|_{S}),\quad\forall s\in S, | | ---------------------------------------------------------------------------------------------------- |
where μ:AS→A\mu:A^{S}\to A is the minimal local map of τ\tau.
If x|S=px|_{S}=p, then p(e)=μ(x|S)=ap(e)=\mu(x|_{S})=a is a contradiction. Hence, x|S≠px|_{S}\neq p and p(e)=μ(x|S)=x(e)p(e)=\mu(x|_{S})=x(e). Suppose that (s⋅x)|S≠p(s\cdot x)|_{S}\neq p for all s∈S∖{e}s\in S\setminus\{e\}. Then,
| p(s)=μ((s⋅x)|S)=(s⋅x)(e)=x(s),p(s)=\mu((s\cdot x)|_{S})=(s\cdot x)(e)=x(s), | | --------------------------------------------------------------------------------------- |
which contradicts that x|S≠px|_{S}\neq p. Therefore, there must exist s∈S∖{e}s\in S\setminus\{e\} such that (s⋅x)|S=p(s\cdot x)|_{S}=p. This implies that p(s)=μ((s⋅x)|S)=a≠p(e)p(s)=\mu((s\cdot x)|_{S})=a\neq p(e). ∎
Given x∈AGx\in A^{G} and b∈Ab\in A, define
suppb(x):={g∈G:x(g)=b}.\textup{supp}_{b}\left(x\right):=\{g\in G:x(g)=b\}. |
---|
The following result presents some elementary properties of supports of configurations under lazy CAs.
Lemma 3.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA with unique active transition p∈ASp\in A^{S} and writing symbol a∈Aa\in A. Let i,j∈ℕ,i≤ji,j\in\mathbb{N},i\leq j. Then:
- suppa(τi(x))⊆suppa(τj(x))\textup{supp}_{a}\left(\tau^{i}(x)\right)\subseteq\textup{supp}_{a}\left(\tau^{j}(x)\right) for all x∈AGx\in A^{G}
- suppb(τi(x))⊇suppb(τj(x))\textup{supp}_{b}\left(\tau^{i}(x)\right)\supseteq\textup{supp}_{b}\left(\tau^{j}(x)\right), for all x∈AGx\in A^{G}, and b∈A∖{a}b\in A\setminus\{a\}.
- If suppa(x)=suppa(τ(x))\textup{supp}_{a}\left(x\right)=\textup{supp}_{a}\left(\tau(x)\right) for some x∈AGx\in A^{G}, then x=τ(x)x=\tau(x).
Proof.
For parts (1.) and (2.), we shall prove the base case i=0i=0 and j=1j=1, as the rest follows by induction. Let g∈suppa(x)g\in\textup{supp}_{a}\left(x\right). Observe that (g⋅x)|S≠p(g\cdot x)|_{S}\neq p, because (g⋅x)(e)=x(g)=a≠p(e)(g\cdot x)(e)=x(g)=a\neq p(e). Definition 1 implies
| τ(x)(g)=μ((g⋅x)|S)=(g⋅x)(e)=x(g)=a,\tau(x)(g)=\mu((g\cdot x)|_{S})=(g\cdot x)(e)=x(g)=a, | | ------------------------------------------------------------------------------------------------------- |
where μ:AS→A\mu:A^{S}\to A is the minimal local map of τ\tau. This shows that g∈suppa(τ(x))g\in\textup{supp}_{a}\left(\tau(x)\right). Now, assume b∈A∖{a}b\in A\setminus\{a\} and g∈suppb(τ(x))g\in\textup{supp}_{b}\left(\tau(x)\right). Since b≠ab\neq a, Definition 1 implies τ(x)(g)=x(g)\tau(x)(g)=x(g), so g∈suppb(x)g\in\textup{supp}_{b}\left(x\right).
For part (3), first note that x(g)=a=τ(x)(g)x(g)=a=\tau(x)(g) for all g∈suppa(x)=suppa(τ(x))g\in\textup{supp}_{a}\left(x\right)=\textup{supp}_{a}\left(\tau(x)\right). Let h∈G∖suppa(τ(x))h\in G\setminus\textup{supp}_{a}\left(\tau(x)\right). This means that μ((h⋅x)|S)=τ(x)(h)≠a\mu((h\cdot x)|_{S})=\tau(x)(h)\neq a, so by Definition 1, (h⋅x)|S≠p(h\cdot x)|_{S}\neq p and
| τ(x)(h)=μ((h⋅x)|S)=(h⋅x)(e)=x(h).\tau(x)(h)=\mu((h\cdot x)|_{S})=(h\cdot x)(e)=x(h). | | -------------------------------------------------------------------------------------------------- |
Therefore, x=τ(x)x=\tau(x). ∎
If ord(τ)<∞\textup{ord}\left(\tau\right)<\infty, let m∈ℕm\in\mathbb{N} and n∈ℤ+n\in\mathbb{Z}_{+} be as small as possible such that τm=τm+n\tau^{m}=\tau^{m+n}. Then, mm and nn are referred to as the index and period of τ\tau, respectively. Note that ord(τ)=m+n\textup{ord}\left(\tau\right)=m+n.
Proposition 2.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA. If ord(τ)<∞\textup{ord}\left(\tau\right)<\infty, then the period of τ\tau is 1.
Proof.
Let μ:AS→A\mu:A^{S}\to A be the minimal local map of τ\tau with unique active transition p∈ASp\in A^{S} and writing symbol a:=μ(p)∈Aa:=\mu(p)\in A. If ord(τ)<∞\textup{ord}\left(\tau\right)<\infty, let mm and nn be the index and period of τ\tau, so τm=τm+n\tau^{m}=\tau^{m+n}. Lemma 3 (1.) implies that for all x∈AGx\in A^{G}
suppa(τm(x))=suppa(τm+i(x)),∀i∈{1,…,n}.\textup{supp}_{a}\left(\tau^{m}(x)\right)=\textup{supp}_{a}\left(\tau^{m+i}(x)\right),\quad\forall i\in\{1,\dots,n\}. |
---|
Hence, Lemma 3 (3.) implies that τm=τm+1\tau^{m}=\tau^{m+1}, so, by the minimality of nn, we must have n=1n=1. ∎
Corollary 2.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA and n∈ℤ+n\in\mathbb{Z}_{+}. Then,
ord(τ)>n⇔τj−1≠τj,∀j∈1,…,n.\textup{ord}\left(\tau\right)>n\iff\tau^{j-1}\neq\tau^{j},\quad\forall j\in{1,\ldots,n}. |
---|
Furthermore,
ord(τ)=min{n≥2:τn−1=τn}.\textup{ord}\left(\tau\right)=\min\{n\geq 2:\tau^{n-1}=\tau^{n}\}. |
---|
We define the product of two subsets S,K⊆GS,K\subseteq G as SK:={sk:s∈S,k∈K}SK:=\{sk:s\in S,k\in K\} and the inverse as S−1:={s−1:s∈S}S^{-1}:=\{s^{-1}:s\in S\}. In the following result we provide a general upper bound for the order of a lazy cellular automaton in terms of its unique active transition.
Theorem 2.
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy CA with minimal neighborhood S⊆GS\subseteq G, unique active transition p∈ASp\in A^{S} and writing symbol a∈A∖{p(e)}a\in A\setminus\{p(e)\}. For any b∈Ab\in A, we define
Sb:=p−1{b}={s∈S:p(s)=b}.S_{b}:=p^{-1}\{b\}=\{s\in S:p(s)=b\}. |
---|
Then, ord(τ)\textup{ord}\left(\tau\right) is at most the minimum n≥2n\geq 2 such that for every word (s1,…,sn−1)∈(Sa)n−1(s_{1},\ldots,s_{n-1})\in(S_{a})^{n-1} there exist 1≤i≤j≤n−11\leq i\leq j\leq n-1 satisfying at least one of the following:
- (sj⋯si)−1∈Sb−1Sa(s_{j}\cdots s_{i})^{-1}\in S_{b}^{-1}S_{a}, for some b∈A∖{a}b\in A\setminus\{a\};
- (sj⋯si)−1∈Sb1−1Sb2(s_{j}\cdots s_{i})^{-1}\in S_{b_{1}}^{-1}S_{b_{2}}, for some b1,b2∈A∖{a}b_{1},b_{2}\in A\setminus\{a\}, with b1≠b2b_{1}\neq b_{2}.
Proof.
We will show that for any n≥2n\geq 2, if τn−1≠τn\tau^{n-1}\neq\tau^{n}, then there exists a word (s1,…,sn−1)∈(Sa)n−1(s_{1},\ldots,s_{n-1})\in(S_{a})^{n-1} such that (1.) and (2.) do not hold for all 1≤i≤j≤n−11\leq i\leq j\leq n-1, so the result follows from Corollary 2.
If τn−1≠τn\tau^{n-1}\neq\tau^{n}, then, by Corollary 1, there exists x∈AGx\in A^{G} such that pp appears in τn−1(x)\tau^{n-1}(x). By the GG-equivariance of cellular automata (i.e., τ(g⋅x)=g⋅τ(x)\tau(g\cdot x)=g\cdot\tau(x), for all g∈Gg\in G, see [6, Prop. 1.4.4]), we may assume that τn−1(x)|S=p\tau^{n-1}(x)|_{S}=p. Applying iteratively Lemmas 1 and 2 yields the existence of a word (s1,…,sn−1)∈(Sa)n−1(s_{1},\ldots,s_{n-1})\in(S_{a})^{n-1} such that
| ((sj⋯s1)⋅τ(n−1)−j(x))|S=p,∀j∈{1,…,n−1}.\big((s_{j}\cdots s_{1})\cdot\tau^{(n-1)-j}(x)\big)|_{S}=p,\quad\forall j\in\{1,\ldots,n-1\}. | | ------------------------------------------------------------------------------------------------------------------------------------------------------- |
It follows from the definition of SbS_{b} that for all g∈Gg\in G, y∈AGy\in A^{G},
| (g⋅y)|S=p⇔Sbg⊆suppb(y),∀b∈A,(g\cdot y)|_{S}=p\quad\iff\quad S_{b}g\subseteq\textup{supp}_{b}\left(y\right),\quad\forall b\in A, | | ------------------------------------------------------------------------------------------------------------------------------------------------- |
Hence, for all b∈Ab\in A,
Sb(sj⋯s1)\displaystyle S_{b}(s_{j}\cdots s_{1}) | ⊆suppb(τ(n−1)−j(x)),∀j∈{1,…,n−1},\displaystyle\subseteq\textup{supp}_{b}\left(\tau^{(n-1)-j}(x)\right),\quad\forall j\in\{1,\ldots,n-1\}, |
---|
Now, Lemmas 3 (1.) and (2.) imply that for all k∈{1,…,n−1}k\in\{1,\dots,n-1\},
⋃j=1k(Sas(n−1)−j⋯s1)\displaystyle\bigcup_{j=1}^{k}\big(S_{a}s_{(n-1)-j}\cdots s_{1}\big) | ⊆suppa(τk(x)),\displaystyle\subseteq\textup{supp}_{a}\left(\tau^{k}(x)\right), |
---|---|
⋃j=kn−1(Sbs(n−1)−j⋯s1)\displaystyle\bigcup_{j=k}^{n-1}\big(S_{b}s_{(n-1)-j}\cdots s_{1}\big) | ⊆suppb(τk(x)),∀b∈A∖{a}.\displaystyle\subseteq\textup{supp}_{b}\left(\tau^{k}(x)\right),\quad\forall b\in A\setminus\{a\}. |
Since supports with respect to different elements of AA are disjoint, we obtain that
[⋃i=1k(Sas(n−1)−i⋯s1)]\displaystyle\left[\bigcup_{i=1}^{k}\big(S_{a}s_{(n-1)-i}\cdots s_{1}\big)\right] | ∩[⋃j=kn−1(Sbs(n−1)−j⋯s1)]=∅,∀b∈A∖{a},\displaystyle\cap\left[\bigcup_{j=k}^{n-1}\big(S_{b}s_{(n-1)-j}\cdots s_{1}\big)\right]=\emptyset,\quad\forall b\in A\setminus\{a\}, | (1) |
---|---|---|
[⋃i=kn−1(Sb1s(n−1)−i⋯s1)]\displaystyle\left[\bigcup_{i=k}^{n-1}\big(S_{b_{1}}s_{(n-1)-i}\cdots s_{1}\big)\right] | ∩[⋃j=kn−1(Sb2s(n−1)−j⋯s1)]=∅,∀b1,b2∈A∖{a},b1≠b2.\displaystyle\cap\left[\bigcup_{j=k}^{n-1}\big(S_{b_{2}}s_{(n-1)-j}\cdots s_{1}\big)\right]=\emptyset,\quad\forall b_{1},b_{2}\in A\setminus\{a\},b_{1}\neq b_{2}. | (2) |
Finally, (1) is equivalent to (sj⋯si)−1∉Sb−1Sa(s_{j}\cdots s_{i})^{-1}\notin S_{b}^{-1}S_{a} for all 1≤i≤j≤n−11\leq i\leq j\leq n-1 and b∈A∖{a}b\in A\setminus\{a\}, while (2) is equivalent to (sj⋯si)−1∉Sb1−1Sb2(s_{j}\cdots s_{i})^{-1}\notin S_{b_{1}}^{-1}S_{b_{2}}, for all 1≤i≤j≤n−11\leq i\leq j\leq n-1, and all b1,b2∈A∖{a}b_{1},b_{2}\in A\setminus\{a\}, b1≠b2b_{1}\neq b_{2}. The result follows.
∎
As an easy consequence of Theorem 2 we obtain the following sufficient conditions on a lazy cellular automaton for being idempotent.
Corollary 3.
With the notation of Theorem 2, assume at least one of the following holds:
- Sa=∅S_{a}=\emptyset;
- Sa−1⊆Sb−1SaS_{a}^{-1}\subseteq S_{b}^{-1}S_{a}, for some b∈A∖{a}b\in A\setminus\{a\};
- Sa−1⊆Sb1−1Sb2S_{a}^{-1}\subseteq S_{b_{1}}^{-1}S_{b_{2}}, for some b1,b2∈A∖{a}b_{1},b_{2}\in A\setminus\{a\} such that b1≠b2b_{1}\neq b_{2}.
Then, τ\tau is idempotent.
Example 3.
Consider the lazy CA τ:Aℤ→Aℤ\tau:A^{\mathbb{Z}}\to A^{\mathbb{Z}} with minimal neighborhood S={−1,0,1,2}S=\{-1,0,1,2\}, unique active transition p=1010∈ASp=1010\in A^{S} and writing symbol a=1∈Aa=1\in A. Since S1={−1,1}=S1−1S_{1}=\{-1,1\}=S_{1}^{-1}, then condition (2.) of Corollary 3 holds, so τ\tau is idempotent.
Example 4.
A counterexample of the converse of Corollary 3 was found by María G. Magaña-Chávez in private communication. Consider the lazy CA τ:Aℤ→Aℤ\tau:A^{\mathbb{Z}}\to A^{\mathbb{Z}} with minimal neighborhood S={−3,−1,0,3,4}S=\{-3,-1,0,3,4\}, unique active transition p=11010∈ASp=11010\in A^{S} and writing symbol a=1∈Aa=1\in A. Then S0={0,4}S_{0}=\{0,4\} and S1={−3,−1,3}S_{1}=\{-3,-1,3\}, so none of the conditions of Corollary 3 holds. However, direct computations show that τ\tau is idempotent.
3 Lazy CA with a quasi-constant active transition
Let τ:AG→AG\tau:A^{G}\to A^{G} be a lazy cellular automaton with minimal local map μ:AS→A\mu:A^{S}\to A, unique active transition p∈ASp\in A^{S}, and writing symbol a:=μ(p)∈A∖{p(e)}a:=\mu(p)\in A\setminus\{p(e)\}. In this section, we assume that pp is quasi-constant, which means that pp is not constant and there exists r∈Sr\in S such that p|S∖{r}p|_{S\setminus\{r\}} is constant. We call r∈Sr\in S the non-constant element of pp. Without loss of generality, we assume that p∈ASp\in A^{S} is defined as follows: for all s∈Ss\in S,
p(s):={1 if s=r0 if s≠r.p(s):=\begin{cases}1&\text{ if }s=r\\ 0&\text{ if }s\neq r.\end{cases} |
---|
Because of Corollary 3, we only need to consider the case when a=p(s)a=p(s) for some s∈Ss\in S (as otherwise, Sa=∅S_{a}=\emptyset).
3.1 Non-constant element r≠er\neq e
In this section, we assume that the non-constant element r∈Sr\in S of p∈ASp\in A^{S} is different from the group identity e∈Se\in S. Then, we must have 1=p(r)=a=μ(p)1=p(r)=a=\mu(p), since we are assuming that p(e)=0p(e)=0.
Proposition 3.
For any n≥2n\geq 2, ord(τ)>n\textup{ord}\left(\tau\right)>n if and only if rj∉Sr^{j}\notin S for all 2≤j≤n2\leq j\leq n.
Proof.
Assume that ord(τ)>n\textup{ord}\left(\tau\right)>n. By Theorem 2, there exists a word w=(s1,…,sn−1)∈(S1)n−1w=(s_{1},\ldots,s_{n-1})\in(S_{1})^{n-1} such that for all i≤ji\leq j,
(sj⋯si)−1∉S0−1S1.(s_{j}\cdots s_{i})^{-1}\notin S_{0}^{-1}S_{1}. |
---|
By the construction of pp in this section, we have S0=S∖{r}S_{0}=S\setminus\{r\} and S1={r}S_{1}=\{r\}, so the word ww has the form w=(r,r,…r)w=(r,r,\dots r). Hence, the above condition is equivalent to rj∉Sr^{j}\notin S, for all 2≤j≤n2\leq j\leq n.
Conversely, assume that rj∉Sr^{j}\notin S for all 2≤j≤n2\leq j\leq n. Define x∈AGx\in A^{G} as follows:
x(g):={p(g) if g∈S0otherwise.x(g):=\begin{cases}p(g)&\textup{ if }g\in S\\ 0&\textup{otherwise}\end{cases}. |
---|
Claim 1.
For all m∈{0,…,n−1}m\in\{0,\ldots,n-1\}, (g⋅τm(x))|S=p(g\cdot\tau^{m}(x))|_{S}=p if and only if g=r−mg=r^{-m}.
Proof.
We prove this by strong induction on mm. For m=0m=0, it is clear by the construction of xx. Now, assume the existence of m∈{0,…,n−2}m\in\{0,\ldots,n-2\} such that the following holds
| (g⋅τℓ(x))|S=p⇔g=r−ℓ,∀ℓ∈{0,…,m}.(g\cdot\tau^{\ell}(x))|_{S}=p\iff g=r^{-\ell},\quad\forall\ell\in\{0,\ldots,m\}. | | -------------------------------------------------------------------------------------------------------------------------------- |
We shall prove the case m+1m+1. By the induction hypothesis and construction of xx,
| τm+1(x)(g)=μ((g⋅τm(x))|S)={1 if g∈{r−ℓ}ℓ=−1m0 otherwise .\tau^{m+1}(x)(g)=\mu((g\cdot\tau^{m}(x))|_{S})=\begin{cases}1&\text{ if }g\in\{r^{-\ell}\}_{\ell=-1}^{m}\\ 0&\text{ otherwise }\end{cases}. | (3) | | ------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ | --- |
We will show that (g⋅τm+1)(x)|S=p(g\cdot\tau^{m+1})(x)|_{S}=p if and only if g=r−(m+1)g=r^{-(m+1)}. First we will show that (r−(m+1)⋅τm+1)(x)|S=p(r^{-(m+1)}\cdot\tau^{m+1})(x)|_{S}=p. For any s∈Ss\in S, (3) implies that
τm+1(x)(sr−(m+1))=1⇔sr−(m+1)∈{r−ℓ}ℓ=−1m.\displaystyle\tau^{m+1}(x)(sr^{-(m+1)})=1\iff sr^{-(m+1)}\in\{r^{-\ell}\}_{\ell=-1}^{m}. | (4) |
---|
If sr−(m+1)=r−ℓsr^{-(m+1)}=r^{-\ell}, then s=rm+1−ℓs=r^{m+1-\ell}. Note that m+1−ℓ∈{1,…,m+2}m+1-\ell\in\{1,\ldots,m+2\}, where m+2≤nm+2\leq n by construction of mm. Since rj∉Sr^{j}\notin S for all 2≤j≤n2\leq j\leq n, it follows that s=rm+1−ℓs=r^{m+1-\ell} if and only if m+1−ℓ=1m+1-\ell=1. As a consequence, ℓ=m\ell=m. Then, by (3) and (4)
τm+1(x)(sr−(m+1))=1\displaystyle\tau^{m+1}(x)(sr^{-(m+1)})=1 | ⇔s=r,\displaystyle\iff s=r, |
---|---|
τm+1(x)(sr−(m+1))=0\displaystyle\tau^{m+1}(x)(sr^{-(m+1)})=0 | ⇔s∈S∖{r}.\displaystyle\iff s\in S\setminus\{r\}. |
Therefore, (r−(m+1)⋅τm+1(x))|S=p(r^{-(m+1)}\cdot\tau^{m+1}(x))|_{S}=p. Conversely, suppose that (g⋅τm+1(x))|S=p(g\cdot\tau^{m+1}(x))|_{S}=p for some g∈Gg\in G. It follows by construction of pp that
τm+1(x)(g)\displaystyle\tau^{m+1}(x)(g) | =p(e)=0,\displaystyle=p(e)=0, |
---|---|
τm+1(x)(rg)\displaystyle\tau^{m+1}(x)(rg) | =p(r)=1.\displaystyle=p(r)=1. |
As a consequence of (3), g∉{r−ℓ}ℓ=−1mg\notin\{r^{-\ell}\}_{\ell=-1}^{m} and rg∈{r−ℓ}ℓ=−1mrg\in\{r^{-\ell}\}_{\ell=-1}^{m}, which implies that g∈{r−ℓ}ℓ=0m+1g\in\{r^{-\ell}\}_{\ell=0}^{m+1}. Thus, g=r−(m+1)g=r^{-(m+1)} and the result follows. ∎
It follows by the previous claim and Lemma 1 that τj−1≠τj\tau^{j-1}\neq\tau^{j} for all j∈{1,…,n}j\in\{1,\ldots,n\}. Therefore, ord(τ)>n\textup{ord}\left(\tau\right)>n by Corollary 2. ∎
Theorem 1 (2.) follows as a direct consequence of Proposition 3.
Recall that the order of an element g∈Gg\in G, denoted by ord(g)\textup{ord}\left(g\right), is the minimum integer n≥1n\geq 1 such that gn=eg^{n}=e, or infinity, in case that no such integer exists. Recall that gi≠gjg^{i}\neq g^{j} holds for all 1≤i<j<ord(g)1\leq i<j<\textup{ord}\left(g\right).
Corollary 4.
Let n≥2n\geq 2 be an integer such that there is g∈Gg\in G with ord(g)>n\textup{ord}\left(g\right)>n. Then, there exists a lazy CA τ:AG→AG\tau:A^{G}\to A^{G} with ord(τ)=n\textup{ord}\left(\tau\right)=n.
Proof.
Let S:={e,g,gn}S:=\{e,g,g^{n}\} and define a quasi-constant pattern p∈ASp\in A^{S} by p(e)=p(gn)=0p(e)=p(g^{n})=0 and p(g)=1p(g)=1. Let τ:AG→AG\tau:A^{G}\to A^{G} be the lazy CA with unique active transition pp and writing symbol 1∈A1\in A. The result follows by Theorem 1 (2.), using the assumption ord(g)>n\textup{ord}\left(g\right)>n. ∎
When the universe GG has an element of infinite order, such as when G=ℤdG=\mathbb{Z}^{d}, the previous corollary implies that there is always a lazy CA of every given order n≥2n\geq 2.
3.2 Non-constant element r=er=e
In this section, we assume that r=er=e, so the writing symbol of τ\tau is a=0=p(s)a=0=p(s), for all s∈S∖{e}s\in S\setminus\{e\}.
Proposition 4.
For any n≥2n\geq 2, ord(τ)>n\textup{ord}\left(\tau\right)>n if and only if there exists a word (s1,…,sn−1)∈(S∖{e})n−1(s_{1},\dots,s_{n-1})\in(S\setminus\{e\})^{n-1} such that
(sj…si)−1∉S,∀i≤j.(s_{j}\dots s_{i})^{-1}\not\in S,\quad\forall i\leq j. |
---|
Proof.
Suppose that ord(τ)>n\textup{ord}\left(\tau\right)>n. By Theorem 2, there exists a word w=(s1,…,sn−1)∈(S0)n−1w=(s_{1},\ldots,s_{n-1})\in(S_{0})^{n-1} such that for all i≤ji\leq j,
(sj⋯si)−1∉S1−1S0.(s_{j}\cdots s_{i})^{-1}\notin S_{1}^{-1}S_{0}. |
---|
By the construction of pp in this section, we have that S0=S∖{e}S_{0}=S\setminus\{e\} and S1={e}S_{1}=\{e\}, so the direct implication follows.
Conversely, assume that there exists a word (s1,…,sn−1)∈(S∖{e})n−1(s_{1},\dots,s_{n-1})\in(S\setminus\{e\})^{n-1} such that
(sj…si)−1∉S,∀i≤j.(s_{j}\dots s_{i})^{-1}\not\in S,\quad\forall i\leq j. |
---|
Define x∈AGx\in A^{G} as follows: for all g∈Gg\in G
x(g):={0 if g∈{si⋯s1s0}i=0n−11otherwisex(g):=\begin{cases}0&\textup{ if }g\in\{s_{i}\cdots s_{1}s_{0}\}_{i=0}^{n-1}\\ 1&\textup{otherwise}\end{cases} |
---|
where s0=es_{0}=e.
Claim 2.
For all i∈{1,…,n}i\in\{1,\ldots,n\}, (g⋅τi(x))|S=p(g\cdot\tau^{i}(x))|_{S}=p if and only if g=sn−i⋯s1s0g=s_{n-i}\cdots s_{1}s_{0}.
Proof.
First of all, observe that the hypothesis (sj…si)−1∉S(s_{j}\dots s_{i})^{-1}\not\in S, for all i≤ji\leq j is equivalent to
{sn−j⋯s1s0}j=kn∩⋃i=1k((S∖{e})sn−i⋯s1s0)=∅,∀1≤k≤n.\{s_{n-j}\cdots s_{1}s_{0}\}_{j=k}^{n}\cap\bigcup_{i=1}^{k}\big((S\setminus\{e\})s_{n-i}\cdots s_{1}s_{0}\big)=\emptyset,\quad\forall 1\leq k\leq n. | (5) |
---|
We prove Claim 2 by strong induction on ii. For the base case i=1i=1, fix k=1k=1 in (5), so we see that x(ssn−1⋯s1)=1x(ss_{n-1}\cdots s_{1})=1 for all s∈S∖{e}s\in S\setminus\{e\} and x(sn−1⋯s1)=0x(s_{n-1}\cdots s_{1})=0 by the construction of xx. It follows that (sn−1⋯s1⋅x)|S=p(s_{n-1}\cdots s_{1}\cdot x)|_{S}=p. Conversely, assume that (g⋅x)|S=p(g\cdot x)|_{S}=p for some g∈Gg\in G, so x(g)=0x(g)=0 and x(sg)=1x(sg)=1 for all s∈S∖{e}s\in S\setminus\{e\}. If follows by the construction of pp that g=sℓ⋯s1s0g=s_{\ell}\cdots s_{1}s_{0} for some ℓ∈{0,…,n−1}\ell\in\{0,\ldots,n-1\} and sj⋯s1s0∉(S∖{e})gs_{j}\cdots s_{1}s_{0}\notin(S\setminus\{e\})g for all j∈{0,…,n−1}j\in\{0,\ldots,n-1\}. If ℓ<n−1\ell<n-1,
sℓ+1=(sℓ+1⋯s1s0)(sℓ⋯s1s0)−1∉S∖{e},s_{\ell+1}=(s_{\ell+1}\cdots s_{1}s_{0})(s_{\ell}\cdots s_{1}s_{0})^{-1}\notin S\setminus\{e\}, |
---|
which contradicts that sℓ+1∈S∖{e}s_{\ell+1}\in S\setminus\{e\}. Therefore, g=sn−1⋯s1s0g=s_{n-1}\cdots s_{1}s_{0}.
Now, assume the existence of m∈{1,…,n−1}m\in\{1,\ldots,n-1\} such that the following holds
| (g⋅τi(x))|S=p⇔g=sn−i⋯s1s0,∀i∈{1,…,m}.(g\cdot\tau^{i}(x))|_{S}=p\iff g=s_{n-i}\cdots s_{1}s_{0},\quad\forall i\in\{1,\ldots,m\}. | | ---------------------------------------------------------------------------------------------------------------------------------------------------- |
We shall prove the case m+1m+1 by showing that (g⋅τm+1(x))|S=p(g\cdot\tau^{m+1}(x))|_{S}=p if and only if g=sn−(m+1)⋯s1s0g=s_{n-(m+1)}\cdots s_{1}s_{0}. By the construction of xx, the induction hypothesis, and Definition 1,
τm(x)(g)={0 if g∈{si⋯s1s0}i=0n−m1otherwise.\tau^{m}(x)(g)=\begin{cases}0&\textup{ if }g\in\{s_{i}\cdots s_{1}s_{0}\}_{i=0}^{n-m}\\ 1&\textup{otherwise}\end{cases}. |
---|
Since g⋅τm(x))|S=pg\cdot\tau^{m}(x))|_{S}=p if and only if g=sn−m⋯s1s0g=s_{n-m}\cdots s_{1}s_{0}, it follows that τm(x)(g)≠τm+1(x)(g)\tau^{m}(x)(g)\neq\tau^{m+1}(x)(g) if and only if g=sn−m⋯s1s0g=s_{n-m}\cdots s_{1}s_{0}. Then
τm+1(x)(g)={0 if g∈{si⋯s1s0}i=0n−(m+1)1otherwise.\displaystyle\tau^{m+1}(x)(g)=\begin{cases}0&\textup{ if }g\in\{s_{i}\cdots s_{1}s_{0}\}_{i=0}^{n-(m+1)}\\ 1&\textup{otherwise}\end{cases}. | (6) |
---|
As a consequence of fixing k=m+1k=m+1 in (5),
sj⋯s1s0∉(S∖{e})sn−(m+1)⋯s1s0,∀j=0,…,n−(m+1).s_{j}\cdots s_{1}s_{0}\notin(S\setminus\{e\})s_{n-(m+1)}\cdots s_{1}s_{0},\quad\forall j=0,\ldots,n-(m+1). |
---|
Thus, τm+1(x)(ssn−(m+1)⋯s1s0)=1\tau^{m+1}(x)(ss_{n-(m+1)}\cdots s_{1}s_{0})=1 for all s∈S∖{e}s\in S\setminus\{e\}. Since τm+1(x)(sn−(m+1)⋯s1s0)=0\tau^{m+1}(x)(s_{n-(m+1)}\cdots s_{1}s_{0})=0, it follows that
| (sn−m+1⋯s1s0⋅τm+1(x))|S=p.\big(s_{n-{m+1}}\cdots s_{1}s_{0}\cdot\tau^{m+1}(x)\big)|_{S}=p. | | --------------------------------------------------------------------------------------------------------- |
Conversely, assume that (g⋅τm+1)(x)|S=p(g\cdot\tau^{m+1})(x)|_{S}=p for some g∈Gg\in G. Observe that τm+1(x)(g)=0\tau^{m+1}(x)(g)=0 and τm+1(x)(sg)=1\tau^{m+1}(x)(sg)=1 for all s∈S∖{e}s\in S\setminus\{e\}. By (6), we have that g=sℓ⋯s1s0g=s_{\ell}\cdots s_{1}s_{0} for some ℓ∈{0,…,n−(m+1)}\ell\in\{0,\ldots,n-(m+1)\} and sj⋯s1s0∉(S∖{e})gs_{j}\cdots s_{1}s_{0}\notin(S\setminus\{e\})g for all j∈{0,…,n−(m+1)}j\in\{0,\ldots,n-(m+1)\}. If ℓ<n−(m+1)\ell<n-(m+1), then
sℓ+1=(sℓ+1⋯s1s0)(sℓ⋯s1s0)−1∉S∖{e},s_{\ell+1}=(s_{\ell+1}\cdots s_{1}s_{0})(s_{\ell}\cdots s_{1}s_{0})^{-1}\notin S\setminus\{e\}, |
---|
which contradicts that sℓ+1∈S∖{e}s_{\ell+1}\in S\setminus\{e\}. Therefore, g=sn−(m+1)⋯s1s0g=s_{n-(m+1)}\cdots s_{1}s_{0}, and the claim follows. ∎
The previous claim and Corollary 1 imply that τj−1≠τj\tau^{j-1}\neq\tau^{j} for all j∈{1,…,n}j\in\{1,\ldots,n\}. The result follows by Corollary 2. ∎
Finally, Theorem 1 (3.) follows as a direct consequence of Proposition 4.
4 Open problems
In this section, we propose two open problems related to the study of lazy CAs.
In this paper, Corollary 3 gives a sufficient condition for the idempotency of a lazy cellular automaton. In [4], Conjecture 4.1 proposes that a lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} with a unique active transition p∈ASp\in A^{S} is not idempotent if and only if pp satisfies a self-overlapping condition. The conjecture is true when S⊆ℤS\subseteq\mathbb{Z} is an interval, but it turns out that the direct implication fails even for general one-dimensional lazy cellular automata. Hence, we propose our first problem.
Problem 1.
Characterize the idempotency of a lazy cellular automaton τ:AG→AG\tau:A^{G}\to A^{G} in terms of its unique active transition p∈ASp\in A^{S} and writing symbol a∈A∖{p(e)}a\in A\setminus\{p(e)\}.
The second problem we propose comes from the theory of monoids. Since the composition of two cellular automata over AGA^{G} is again a cellular automaton over AGA^{G}, the set of all cellular automata over AGA^{G}, denoted by CA(G,A)\text{CA}(G,A) or End(AG)\text{End}(A^{G}), is a monoid equipped with the composition of functions. The group of all invertible cellular automata over AGA^{G} is denoted by ICA(G,A)\text{ICA}(G,A) or Aut(AG)\text{Aut}(A^{G}). For any subset 𝒞\mathcal{C} of cellular automata over AGA^{G}, denote by ⟨𝒞⟩\langle\mathcal{C}\rangle the submonoid of CA(G,A)\text{CA}(G,A) generated by 𝒞\mathcal{C}.
It is well-known that the full transformation monoid Tran(A)\text{Tran}(A) of a finite set AA is generated by the idempotents of defect 11 (i.e., self-maps of AA whose image has size |A|−1|A|-1) together with the group of invertible transformations Sym(A)\text{Sym}(A) [10]. As the minimal local maps of lazy cellular automata over AGA^{G} with minimal neighborhood S={e}S=\{e\} are precisely the idempotents of defect 11 of AA, the previous result inspired the following problem.
Problem 2.
If ℒ(G,A)\mathcal{L}(G,A) is the set of all lazy cellular automata over AGA^{G}, prove or disprove the following:
CA(G,A)=⟨ICA(G,A)∪ℒ(G,A)⟩.\text{CA}(G,A)=\langle\text{ICA}(G,A)\cup\mathcal{L}(G,A)\rangle. |
---|
If the above does not hold, what can we say of the submonoids ⟨ICA(G,A)∪ℒ(G,A)⟩\langle\text{ICA}(G,A)\cup\mathcal{L}(G,A)\rangle and ⟨ℒ(G,A)⟩\langle\mathcal{L}(G,A)\rangle?
In other words, Problem 2 asks if every cellular automaton over AGA^{G} may be written as a composition of lazy and invertible CAs.
Acknowledgments.
The first author was supported by SECIHTI Becas nacionales para estudios de posgrados. We sincerely thank Nazim Fatès for suggesting the name lazy for the class of cellular automata studied in this paper during the 31st International Workshop on Cellular Automata and Discrete Complex Systems AUTOMATA 2025, at the University of Lille, France.
References
- [1] Balbi, P. P., Mattos, T., Ruivo, E.: Characterisation of the elementary cellular automata with neighbourhood priority based deterministic updates. Comun. Nonlinear Sci. Numer. Simulat. 104 (2022) 106018. https://doi.org/10.1016/j.cnsns.2021.106018
- [2] Balbi, P. P., Mattos, T., Ruivo, E.: From Multiple to Single Updates Per Cell in Elementary Cellular Automata with Neighbourhood Based Priority. In: Das, S., Roy, S., Bhattacharjee, K. (eds) The Mathematical Artist. Emergence, Complexity and Computation 45, Springer, Cham., 2022. https://doi.org/10.1007/978-3-031-03986-7_6
- [3] Castillo-Ramirez, A., Magaña-Chavez, M. G., Veliz-Quintero, E.: Idempotent cellular automata and their natural order, Theoretical Computer Science, 1009, https://doi.org/10.1016/j.tcs.2024.114698.
- [4] Castillo-Ramirez, A., Magaña-Chavez, M. G., Santos Baños, L. de los.: One-dimensional cellular automata with a unique active transition, Dynamical Systems, 1-15, https://doi.org/10.1080/14689367.2025.2487698.
- [5] Castillo-Ramirez, A., Veliz-Quintero, E.: On the Minimal Memory Set of Cellular Automata. In: Gadouleau, M., Castillo-Ramirez, A. (eds) Cellular Automata and Discrete Complex Systems. AUTOMATA 2024. Lecture Notes in Computer Science, vol 14782. Springer, Cham, 2024. https://doi.org/10.1007/978-3-031-65887-7_7
- [6] Ceccherini-Silberstein, T., Coornaert, M.: Cellular Automata and Groups. 2nd Ed. Springer Monographs in Mathematics, Springer Cham, 2023. https://doi.org/10.1007/978-3-031-43328-3
- [7] Concha-Vega, P., Goles, E., Montealegre, P., Ríos-Wilson, M., & Santivañez, J.: Introducing the activity parameter for elementary cellular automata. Int. J. Mod. Phys. C 33(9) (2022) 2250121. https://doi.org/10.1142/S0129183122501212
- [8] Fates, N.: A tutorial on elementary cellular automata with fully asynchronous updating. Nat. Comput. 19 (1) (2020) 179–197. https://doi.org/10.1007/s11047-020-09782-7
- [9] Fates, N.: Asynchronous Cellular Automata. In: Meyers, R. (eds) Encyclopedia of Complexity and Systems Science. Springer, Berlin, Heidelberg, 2018. https://doi.org/10.1007/978-3-642-27737-5_671-2
- [10] Howie, J. M.: The Subsemigroup Generated By the Idempotents of a Full Transformation Semigroup, J. London Math. Soc. s1-41 (1966) 707-716. https://doi.org/10.1112/jlms/s1-41.1.707
- [11] Kari, J.: Theory of cellular automata: a survey. Theoretical Computer Science 334 (2005) 3 – 33. https://doi.org/10.1016/j.tcs.2004.11.021
- [12] Kůrka, P. (1997): Languages, equicontinuity and attractors in cellular automata. Ergodic Theory and Dynamical Systems 17(2), 417–433. https://doi.org/10.1017/S014338579706985X
- [13] Rotman, J.: An Introduction to the Theory of Groups. 4th Ed. Springer New York, 1995.
- [14] Wolfram, S. (1983): Statistical mechanics of cellular automata. Reviews of Modern Physics, 55(3), 601–644.