1 Introduction

Let G be a group and let A be a set. Consider the set \(A^G\) of all functions from G to A equipped with the shift action of G, defined by

$$\begin{aligned} (g \cdot x) (h) := x(g^{-1}h), \end{aligned}$$

for all \(g, h \in G\) and \(x \in A^G\). Although we shall not focus on this, the set \(A^G\) is usually seen as a topological space with the product topology of the discrete topology on A.

The G-space \(A^G\) is a fundamental object in areas such as symbolic dynamics and the theory of cellular automata (e.g. see [4, 10]). Following [4], we call the elements of \(A^G\) configurations. For any \(x \in A^G\), the stabiliser \(G_x\) of x and the G-orbit Gx of x are defined as follows:

$$\begin{aligned} G_x := \{ g \in G : g \cdot x = x \} \quad \text { and } \quad Gx := \{ g \cdot x \in A^G : g \in G \}. \end{aligned}$$

For a subgroup H of G, a configuration \(x \in A^G\) has period H, or is H-periodic, if \(h \cdot x = x\) for all \(h \in H\), or, equivalently, if \(H \le G_x\). Denote by \(\mathrm {Fix}(H)\) the subset of \(A^G\) consisting of all H-periodic configurations. It is known (see [4, Proposition 1.3.3]) that \(\mathrm {Fix}(H)\) is in bijection with \(A^{H{\setminus } G}\), where \(H{\setminus } G = \{ Hg :g \in G \}\) is the set of rights cosets of H in G. Hence, it follows that \(\vert \mathrm {Fix}(H) \vert = \vert A \vert ^{[G:H]}\), where \([G:H] := \vert H{\setminus } G \vert \) is the index of H in G. In particular, the configurations whose period is the trivial subgroup of G are known as aperiodic points, and have been used in [6] as powerful tools to study the dynamics in \(A^G\) and its subshifts, or subflows (i.e. closed G-equivariant subsets of \(A^G\)).

We say that \(x \in A^G\) has least period, or fundamental period, H if \(G_x = H\) (c.f. [10, Definition 1.1.3.]). In this paper, we are interested in the number \(\psi _H(G;A)\) of configurations with least period H:

$$\begin{aligned} \psi _H(G;A) := \vert \{ x \in A^G : G_x = H \} \vert . \end{aligned}$$

If \(x,y \in A^G\) satisfy that \(y=g\cdot x\), then \(G_y = g G_x g^{-1}\); hence, it is sometimes convenient to consider the G-invariant set \(\{ x \in A^G : [G_x] = [H] \}\), where \([H] := \{gHg^{-1} : g \in G \}\) is the conjugacy class of H, and its cardinality

$$\begin{aligned} \psi _{[H]}(G;A) := \vert \{ x \in A^G : [G_x] = [H] \} \vert . \end{aligned}$$

As \(\psi _{H}(G;A) = \psi _{gHg^{-1}}(G;A)\) for all \(g \in G\), we have

$$\begin{aligned} \psi _{[H]}(G;A) = \vert [H] \vert \; \psi _H(G;A). \end{aligned}$$

Finally, we also consider the number of G-orbits whose stabiliser is conjugate to H:

$$\begin{aligned} \alpha _{[H]}(G;A) := \vert \{ Gx : [G_x] = [H] \} \vert . \end{aligned}$$

By the Orbit-Stabiliser Theorem [14, Theorem 7.2.1], all G-orbits inside \(\{ x \in A^G : [G_x] = [H]\}\) have size [G : H]; therefore, we have

$$\begin{aligned} \alpha _{[H]}(G;A) \; [G:H] = \psi _{[H]}(G;A). \end{aligned}$$

Besides being interesting for their own right, the above numbers have connections with the structure of the automorphism group of \(A^G\). Recall that a map \(\tau : A^G \rightarrow A^G\) is G-equivariant if \(\tau (g \cdot x) = g \cdot \tau (x)\), for all \(g \in G\), \(x \in A^G\). Let \(\mathrm {Aut}(A^G)\) the group of all G-equivariant homeomorphisms of \(A^G\). By the Curtis–Heldund Theorem [4, Theorem 1.8.1], \(\mathrm {Aut}(A^G)\) is the same as the group of invertible cellular automata of \(A^G\). It follows by G-equivariance that for every \(\tau \in \mathrm {Aut}(A^G)\), \(x \in A^G\), we have \(G_{x} = G_{\tau (x)}\). Thus, \(\psi _{G_x}(G;A)\) is an upper bound for the cardinality of the \(\mathrm {Aut}(A^G)\)-orbit of x. Moreover, if the group G is finite, the structure of \(\mathrm {Aut}(A^G)\) was described in [3, Theorem 3] as

$$\begin{aligned} \mathrm {Aut}(A^G) \cong \prod _{i=1}^{r} ( (N_G(H_i)/H_i) \wr \mathrm {Sym}_{\alpha _i}), \end{aligned}$$
(1.1)

where \([H_1], \dots , [H_r]\) is the list of all different conjugacy classes of subgroups of G, and \(\alpha _i = \alpha _{[H_i]}(G;A)\), as defined above. Hence, the structure of \(\mathrm {Aut}(A^G)\) completely depends on the quotient groups \(N_G(H_i)/H_i\), which may be easily calculated by knowing the group G, and the integers \(\alpha _{[H_i]}(G;A)\), which depend on \(\psi _H(G;A)\). Finally, in [1, 2], the sets of points of a given least period were a fundamental tool in the study of automorphism groups of shifts of finite type, which include the group \(\mathrm {Aut}(A^{{\mathbb {Z}}})\).

As \(\psi _H(G;A)\) is finite if and only if [G : H] is finite (see Lemma 2.1 below), we shall focus on finite index subgroups of G. In the first part of this paper, we prove that, when G is finitely generated, the poset L(G) of finite index subgroups of G is a locally finite lattice, so we use Möbius inversion to show that

$$\begin{aligned} \psi _H(G;A) = \sum _{H \le K \le G} \mu (H,K) \vert A \vert ^{[G:K]}, \end{aligned}$$
(1.2)

where \(\mu \) is the Möbius function of L(G). In the second part of this paper, we note that if H is a normal subgroup, then \(\psi _H(G;A) = \psi _1(G/H;A)\) and \(\alpha _{[H]}(G;A) = \alpha _{[1]}(G/H;A)\). Hence, by computing the Möbius function of the subgroup lattice of all finite groups of size up to 7, we classify under which situations we have \(\alpha _{[H]}(G;A) \le 10\).

Our work generalises previous results known in the literature. When \(G = {\mathbb {Z}}_n\) is a cyclic group and \(H=1\) is the trivial subgroup, \(\alpha _{[1]}({\mathbb {Z}}_n;A)\) is equivalent to the number of aperiodic necklaces of length n, and equation (1.2) gives the so-called Moreau’s necklace-counting function [12]. Moreover, \(\alpha _{[1]}({\mathbb {Z}}_n;A)\) is also equivalent to the number of Lyndon words of length n [11, Sect. 5.1] . For a finite group G, this equation may be derived using the result of [9, Sect. 4]. However, as far as we know, Eq. (1.2) had not been derived when G is an arbitrary finitely generated group.

2 Periodic Configurations when G is Finitely Generated

For the rest of the paper, let A be a set with at least two elements and assume that \(\{0,1 \} \subseteq A\). We begin by justifying our claim that \(\psi _H(G;A)\) is finite if and only if [G : H] is finite.

Lemma 2.1

Let G be a group and let H be a subgroup of G. Then \(\psi _H(G;A)\) is finite if and only if [G : H] is finite.

Proof

If [G : H] is finite, then \(\psi _H(G;A)\) is clearly finite, as every configuration with least period H is contained in \(\mathrm {Fix}(H)\) and \(\vert \mathrm {Fix}(H) \vert = \vert A \vert ^{[G:H]} < \infty \).

Conversely, suppose that [G : H] is infinite. Let \(T \subseteq G\) be a transversal for the set of right cosets of H in G, i.e., T contains exactly one element from each right coset of H in G. It is clear that \(\vert T \vert = [G:H]\). For each \(s \in T\), consider the configuration \(x_s \in A^G\) defined by

$$\begin{aligned} x_s(g) = {\left\{ \begin{array}{ll} 1 &{} \text { if } g \in Hs \\ 0 &{} \text { otherwise} \end{array}\right. } ,\end{aligned}$$

for any \(g \in G\). Given \(h \in H\), then \(h \cdot x_s(g) = x_s(h^{-1}g) = x_s(g)\), as \(h^{-1}g \in Hs\) if and only if \(g \in Hs\). Hence, \(H \le G_{x_s}\). On the other hand, if \(k \in G_{x_s}\), then \(k \cdot x_s = x_s\); in particular we have \((k \cdot x_s)(s) =x_s(k^{-1}s) = x_s(s) = 1\), which implies that \(k^{-1}s \in Hs\). Therefore, \(k \in H\), which shows that \(G_{x_s}=H\). As \(\vert T \vert = [G:H]\) is infinite, we have constructed infinitely many different configurations with least period H, which establishes that \(\psi _{H}(G;A)\) is infinite. \(\square \)

We shall recall some basic definitions on posets; for further details see [15, Ch. 3]. Recall that a partially ordered set, or a poset, is a set P equipped with a partial order relation \(\le \). Given \(s, t \in P\) with \(s \le t\), define the closed interval \([s,t] := \{ u \in P : s \le u \le t \}\). We say that P is locally finite if every closed interval of P is finite. A chain of P is a subposet S of P that is totally ordered, i.e. any two elements of S are comparable. For \(t \in P\), the principal order ideal generated by t is \(\Lambda _t := \{ s \in P : s \le t \}\), and the principal dual order ideal generated by t is \(V_t := \{s \in P : s \ge t \}\).

A lattice is a poset L for which every pair of elements \(s,t \in L\) has a lest upper bound, denoted by \(s \vee t\) and read s join t, and a greatest lower bound, denoted by \(s \wedge t\) and read s meet t.

The Möbius function of a locally finite poset P is the map \(\mu : P \times P \rightarrow {\mathbb {Z}}\) defined inductively by the following equations:

$$\begin{aligned} \mu (a,a)&= 1, \ \ \forall a \in P, \\ \mu (a,b)&= 0, \ \ \forall a \not \le b, \\ \sum _{a \le c \le b} \mu (a,c)&= 0, \ \ \forall a < b. \end{aligned}$$

The Möbius function is the inverse of the zeta function of a locally finite poset, and it importantly satisfies the so-called Möbius inversion formula (see [15, Sect. 3.7]). In this section we shall use the dual form of the Möbius inversion formula [15, Proposition 3.7.2].

Theorem 2.2

(Möbius inversion formula, dual form). Let P be a poset for which every principal dual order ideal \(V_t\) is finite. Consider functions \(f,g : P \rightarrow K\), where K is a field. Then

$$\begin{aligned} g(t) = \sum _{s \ge t} f(s), \quad \forall t \in P, \end{aligned}$$

if and only if

$$\begin{aligned} f(t) = \sum _{s \ge t} g(s) \mu (t,s), \quad \forall t \in P. \end{aligned}$$

For any group G, it is standard to consider the poset of all subgroups of G ordered by inclusion. Here, we shall consider the poset L(G) of all subgroups of G of finite index ordered by inclusion. The following is a key observation for this section.

Lemma 2.3

The poset L(G) is a lattice. Furthermore, if G is finitely generated, then for every \(H \in L(G)\), the principal dual order ideal \(V_H = \{ K \le G : H \le K \}\) is finite, so L(G) is a locally finite lattice.

Proof

We shall show that L(G) is a sublattice of the subgroup lattice of G by showing that it is closed under the join, given by \(H \vee J = \langle H \cup J \rangle \), and the meet, given by \(H \wedge J = H \cap J\).

Let H and K be subgroups of G such that \(H \le K\). It is well-known (see, for instance [14, Theorem 3.1.3]) that the indices of H and K in G satisfy, as cardinal numbers, that

$$\begin{aligned}{}[G:H] = [G:K] [K:H]. \end{aligned}$$

Hence, if [G : H] is finite, then [G : K] must be finite. This implies that for any \(H, J \in L(G)\), then \(\langle H \cup J \rangle \in L(G)\). On the other hand, it is also well-known (see, for instance [14, Theorem 3.1.6]) that the intersection of subgroups of finite index has finite index, so \(H \cap J \in L(G)\), and the first part of the lemma follows.

For the second part, for any \(H \in L(G)\) and \(K \in V_H\), the index of K in G must be a divisor of [G : H]. The result follows as in a finitely generated group there are only finitely many subgroups of a given finite index (this is a well-known theorem by M. Hall [7]; see also [14, Theorem 4.20]). \(\square \)

The previous lemma allows us to use the Möbius inversion formula for the poset L(G) when G is finitely generated. Let \(\mu \) be the Möbius function of L(G).

Theorem 2.4

Let G be a finitely generated group, let H be a subgroup of G of finite index, and let A be a finite set. Then,

$$\begin{aligned} \psi _H(G;A) = \sum _{H \le K \le G} \mu (H,K) \vert A \vert ^{[G:K]}. \end{aligned}$$

Proof

It follows from the definitions that

$$\begin{aligned} \vert \mathrm {Fix}(H) \vert = \sum _{K \ge H} \psi _K(G;A) . \end{aligned}$$

By Lemma 2.3 this summation is finite and we may use Theorem 2.2, with \(g(H) = \vert \mathrm {Fix}(H) \vert \) and \(f(K) = \psi _K(G;A)\). Therefore, we obtain

$$\begin{aligned} \psi _H(G;A) = \sum _{K \ge H} \mu (H,K) \vert \mathrm {Fix}(K) \vert . \end{aligned}$$

The result follows as \(\vert \mathrm {Fix}(K) \vert = \vert A \vert ^{[G:K]}\) by [4, Proposition 1.3.3]. \(\square \)

Remark 2.5

Note that, for any \(H,J \in L(G)\), the value of \(\mu (H,J)\) only depends on the on the interval [HJ]. Hence, \(\psi _H(G;A)\) may be calculated by only knowing the subposet [HG].

Corollary 2.6

With the notation of Theorem 2.4, suppose that the interval from H to G consists of a chain \(H = H_0< H_1< \dots < H_k=G\). Then,

$$\begin{aligned} \psi _H(G;A) = \vert A \vert ^{[G:H]} - \vert A \vert ^{[G:H_1]}. \end{aligned}$$

In particular, if H is a maximal subgroup of G, then

$$\begin{aligned} \psi _H(G;A) = \vert A \vert ^{[G:H]} - \vert A \vert . \end{aligned}$$

Proof

By Theorem 2.4,

$$\begin{aligned} \psi _H(G;A) = \sum _{i=0}^k \mu (H,H_i) \vert A \vert ^{[G:H_i]}. \end{aligned}$$

Now, by the definition of the Möbius function,

$$\begin{aligned} \mu (H,H_0)&= 1, \\ \mu (H,H_1)&= -1, \\ \mu (H,H_i)&= 0, \quad \quad \forall i=2,3,\dots , k. \end{aligned}$$

The result follows. \(\square \)

Corollary 2.7

With the notation of Theorem 2.4,

$$\begin{aligned} \psi _{[H]}(G;A)&= \vert [H] \vert \sum _{H \le K \le G} \mu (H,K) \vert A \vert ^{[G:K]}, \\ \alpha _{[H]}(G;A)&= \frac{\vert [H] \vert }{[G:H]} \sum _{H \le K \le G} \mu (H,K) \vert A \vert ^{[G:K]}. \end{aligned}$$

3 Configurations with Normal Period

In this section, we shall specialise on the case when H is a normal subgroup of G of finite index. In this case, the conjugacy class of H just contains H itself, so

$$\begin{aligned} \psi _H(G;A) = \psi _{[H]}(G;A). \end{aligned}$$

Denote by 1 the trivial subgroup. The following result has been noted in [3, Lemma 6].

Lemma 3.1

Let G be any group and let H be a normal subgroup of G of finite index. Then,

$$\begin{aligned}\psi _H(G;A) = \psi _1(G/H;A) \ \text { and } \ \alpha _{[H]}(G;A) = \alpha _{[1]}(G/H; A). \end{aligned}$$

Proof

By [4, Proposition 1.3.7.], there is a G/H-equivariant bijection between \(A^{G/H}\) and \(\mathrm {Fix}(H)\). Hence, configurations in \(A^{G/H}\) with trivial stabiliser are in bijection with the configurations in \(A^G\) with stabiliser equal to H. \(\square \)

The previous lemma allows to apply the machinery of Möbius functions of subgroup lattices which has been developed for a variety of finite groups (e.g. see [5, 8, 13]).

Recall that the classical Möbius function \({\tilde{\mu }}\) of the poset of natural numbers \({\mathbb {N}}\) ordered by divisibility is given by

$$\begin{aligned} {\tilde{\mu }}(d) = {\left\{ \begin{array}{ll} 0 &{} \text { if }d\text { has a squared prime factor} \\ 1 &{} \text { if }d\text { is square-free with an even number of prime factors} \\ -1 &{} \text { if }d\text { is square-free with an odd number of prime factors}. \end{array}\right. }\end{aligned}$$

Using Lemma 3.1, the following result gives the values of \(\psi _H(G;A)\) in some particular cases when H is a normal subgroup of G.

Lemma 3.2

Let G be a finitely generated group, let H be a normal subgroup of G of finite index, and let A be a finite set. Let \(n \in {\mathbb {N}}\), and let p and \(p^\prime \) be two distinct primes.

  1. (1)

    If \(G/H \cong {\mathbb {Z}}_n\), then \(\psi _{H}(G;A) = \sum _{d \mid n} {\tilde{\mu }}(d) \vert A \vert ^{n/d}\).

  2. (2)

    If \(G/H \cong {\mathbb {Z}}_{p^k}\), then \(\psi _{H}(G;A)= \vert A \vert ^{p^k} - \vert A \vert ^{p^{k-1}}\).

  3. (3)

    If \(G/H \cong {\mathbb {Z}}_{pp^\prime }\), then \(\psi _{H}(G;A)= \vert A \vert ^{pp^\prime } - \vert A \vert ^{p} - \vert A \vert ^{p^\prime }+ \vert A \vert \).

  4. (4)

    If \(G/H \cong {\mathbb {Z}}_{p} \oplus {\mathbb {Z}}_p\), then \(\psi _{H}(G;A) = \vert A \vert ^{p^2} - (p+1)\vert A \vert ^p + p\vert A \vert \).

Proof

Parts (1), (2) and (3) follow as it is well-known that \(\mu (1, {\mathbb {Z}}_n) = {\tilde{\mu }}(n)\) (as the subgroup lattice of \({\mathbb {Z}}_n\) is isomorphic to the divisibility lattice of n). For part (4), just observe that the group \({\mathbb {Z}}_{p} \oplus {\mathbb {Z}}_p\) has \(\frac{p^2-1}{p-1}=p+1\) subgroups isomorphic to \({\mathbb {Z}}_p\) (as each of the \(p^2-1\) nontrivial elements of \({\mathbb {Z}}_{p} \oplus {\mathbb {Z}}_p\) generates a subgroup with \(p-1\) nontrivial elements), which account for all its proper nontrivial subgroups. \(\square \)

In the rest of this section, we shall focus on the exact determination of the small values of \(\alpha _{[H]}(G;A)\). The inspiration for this question is [3, Lemma 5], which established, without using the Möbius function, that \(\alpha _{[H]}(G;A) = 1\) if and only if \([G:H]=2\) and \(\vert A \vert = 2\). In general, the classification of small values of \(\alpha _{[H]}(G;A)\) is relevant as it classifies configurations with small \(\mathrm {Aut}(A^G)\)-orbits, and, when G is finite, it classifies the small degrees of the symmetric groups appearing in the decomposition (1.1) of \(\mathrm {Aut}(A^G)\).

For \(x \in A^G\), we have \(G_x = G\) if and only if x is a constant configuration. As we have precisely \(\vert A \vert \) constant configurations in \(A^G\), then \(\alpha _{[G]}(G;A) = \vert A \vert \). Hence, we shall exclude the case \(H=G\) in the following theorem. Moreover, we exclude the degenerate case \(\vert A \vert = 1\).

Table 1 Small values for \(\alpha _{[H]}(G;A)\) with H normal in G

Theorem 3.3

Let G be a finitely generated group, let H be a proper normal subgroup of G of finite index, and let A a finite set with at least two elements.

  1. (1)

    \(\alpha _{[H]}(G;A) = 1\) if and only if \(\vert A \vert =2\) and \([G:H] = 2\).

  2. (2)

    \(\alpha _{[H]}(G;A) = 2\) if and only if \(\vert A \vert =2\) and \([G:H] = 3\), or \(\vert A \vert =2\) and \(G/H \cong {\mathbb {Z}}_2 \oplus {\mathbb {Z}}_2\).

  3. (3)

    \(\alpha _{[H]}(G;A) = 3\) if and only if \(\vert A \vert =3\) and \([G:H]=2\), or \(\vert A \vert =2\) and \(G/H \cong {\mathbb {Z}}_4\).

  4. (4)

    \(\alpha _{[H]}(G;A) = 6\) if and only if \(\vert A \vert =2\) and \([G:H] =5\), or \(\vert A \vert =4\) and \([G:H]=2\).

  5. (5)

    \(\alpha _{[H]}(G;A) = 7\) if and only if \(\vert A \vert =2\) and \(G/H \cong S_3\).

  6. (6)

    \(\alpha _{[H]}(G;A) = 8\) if and only if \(\vert A \vert =3\) and \([G:H] = 3\).

  7. (7)

    \(\alpha _{[H]}(G;A) = 9\) if and only if \(\vert A \vert =2\) and \(G/H \cong {\mathbb {Z}}_6\).

  8. (8)

    \(\alpha _{[H]} (G;A)= 10\) if and only if \(\vert A \vert =5\) and \([G:H]=2\).

  9. (9)

    \(\alpha _{[H]}(G;A) \ne 4\) and \(\alpha _{[H]}(G;A) \ne 5\).

Proof

By [6, Corollary 1.7.2],

$$\begin{aligned} \vert A \vert ^{[G:H]} - \vert A \vert ^{[G:H]-1} \le \alpha _{[1]}(G/H,A) = \alpha _{[H]}(G;A). \end{aligned}$$

(This lower bound has been improved in [3, Theorem 5], but the above is enough for this proof). Hence, we see that \(\alpha _{[1]}(G/H,A)\) is a strictly increasing function on both [G : H] and \(\vert A \vert \). Table 1 shows all values of \(\alpha _{[1]}(G/H,A)\) with \([G:H] \le 7\) and \(\vert A \vert \le 5\). Most of these values may be calculated using the formulas of Lemma 3.2; the only exception is the case \(G/H \cong S_3\), which may be directly computed using the Möbius function of the subgroup lattice of \(S_3\) (see Fig. 1). The result follows by inspection of Table 1. \(\square \)

Fig. 1
figure 1

Subgroup lattice of \(S_3\)