1 Introduction

Introduced in [11] by Houghton, the Houghton groups have since attracted attention for their finiteness properties [2, 13], their growth [3, 10], their many interesting combinatorial features [1, 9, 14, 15] as well as other properties.

Definition 1 For \(X\ne \emptyset \), let \({{\,\mathrm{Sym}\,}}(X)\) denote the group of all bijections on X. For \(g\in {{\,\mathrm{Sym}\,}}(X)\), let \({{\,\mathrm{supp}\,}}(g):=\{x\in X\;:\;(x)g\ne x\}\), called the support of g. Then \({{\,\mathrm{FSym}\,}}(X)=\{g\in {{\,\mathrm{Sym}\,}}(X)\;:\;|{{\,\mathrm{supp}\,}}(g)|<\infty \}\) and \({{\,\mathrm{Alt}\,}}(X)\le {{\,\mathrm{FSym}\,}}(X)\) consists of only the even permutations, meaning \([{{\,\mathrm{FSym}\,}}(X)\;:\;{{\,\mathrm{Alt}\,}}(X)]=2\).

We give a brief overview of these groups for our purposes; more detailed introductions can be found, for example, in [1, 4]. We will use right actions throughout. We define \(\mathbb {N}:=\{1, 2,\ldots \}\), let \(G\le _f H\) denote that G is a finite index subgroup of H, and for a group G and \(g, h\in G\), let \([g, h]:=g^{-1}h^{-1}gh\) and \(g^h:=h^{-1}gh\).

Definition 2 Let \(n\in \{3, 4, \ldots \}\). Then the nth Houghton group, denoted \(H_n\), is generated by \(g_2, \ldots , g_n\in {{\,\mathrm{Sym}\,}}(X_n)\) where \(X_n=\{1, \ldots , n\}\times \mathbb {N}\) and for \(k\in \{2, \ldots , n\}\),

$$\begin{aligned} (i,m)g_k=\left\{ \begin{array}{ll}(1, m+1) &{} \mathrm {if}\ i=1\ \mathrm {and}\ m\in \mathbb {N},\\ (1,1) &{} \mathrm {if}\ i=k\ \mathrm {and}\ m=1,\\ (k, m-1) &{} \mathrm {if}\ i=k\ \mathrm {and}\ m\in \{2, 3, \ldots \},\\ (i, m) &{} \mathrm {otherwise}. \end{array}\right. \end{aligned}$$
(1)

For each \(n\in \{3, 4, \ldots \}\), we have that \({{\,\mathrm{FSym}\,}}(X_n)\le H_n\). One way to see this is to first compute that \([g_2, g_3]=((1, 1), (1, 2))\), and then observe that any 2-cycle with support in \(X_n\) can be conjugated, using the elements \(g_2, \ldots , g_n\), to ((1, 1), (1, 2)). Furthermore, as observed in [15], we have a short exact sequence of groups

$$\begin{aligned} 1\longrightarrow {{\,\mathrm{FSym}\,}}(X_n){\mathop {\longrightarrow }\limits ^{}} H_n{\mathop {\longrightarrow }\limits ^{\pi }} \mathbb {Z}^{n-1}\longrightarrow 1. \end{aligned}$$

Here \(\pi \) is induced by defining \(\pi (g_i):=e_{i-1}\) for \(i=2, \ldots , n\), where \(e_i\) denotes the vector in \(\mathbb {Z}^{n-1}\) with ith entry 1 and other entries 0.

Definition 3 The second Houghton group is generated by the two cycle ((1, 1), (1, 2)) together with the element \(g_2\), defined analogously to (1) above. This is isomorphic to \({{\,\mathrm{FSym}\,}}(\mathbb {Z})\rtimes \langle t\rangle \) where \(t\in {{\,\mathrm{Sym}\,}}(\mathbb {Z})\) sends each \(z\in \mathbb {Z}\) to \(z+1\).

There have been many papers with questions and results relating to the finite index subgroups of this family of groups, e.g. the questions on invariable generation in [15] and subsequent answers in [6], showing they all have solvable conjugacy problem in [4], and also all have the \(R_\infty \) property [5, 8]. Some of these use the partial description of the finite index subgroups from [3]. We start by giving the first full description of them.

Theorem 1

Let \(n\in \{3, 4, \ldots \}\) and \(U\le _f H_n\). Then there exist \(c_2, \ldots , c_n\in \mathbb {N}\) such that \(\pi (U)=\langle c_2e_1, \ldots , c_ne_{n-1}\rangle \) and either U is:

  1. (i)

    equal to \(\langle {{\,\mathrm{FSym}\,}}(X_n), g_i^{c_i}\;:\;i=2, \ldots , n\rangle \); or

  2. (ii)

    isomorphic to \(\langle {{\,\mathrm{Alt}\,}}(X_n), g_i^{c_i}\;:\;i=2, \ldots , n\rangle \).

If \(U\le _fH_2\), then there exists \(c_2\in \mathbb {N}\) such that \(\pi (U)=\langle c_2\rangle \le \mathbb {Z}\) and either (i) or (ii) above occurs or U is equal to \(\langle {{\,\mathrm{Alt}\,}}(X_2), ((1, 1), (1, 2)g_2)\rangle \).

Given \(U, U'\le _fH_n\) such that \(\pi (U)=\langle c_2e_1, \ldots , c_ne_{n-1}\rangle \) for \(c_2, \ldots , c_n\in \mathbb {N}\) and \(\pi (U')=\langle d_2e_1, \ldots , d_ne_{n-1}\rangle \) for \(d_2, \ldots , d_n\in \mathbb {N}\), one might wonder when \(U\cong U'\). Clearly any permutation of the constants \(c_2, \ldots , c_n\) produces an isomorphism. By considering \({{\,\mathrm{Aut}\,}}(U)\) and \({{\,\mathrm{Aut}\,}}(U')\), it seems that this is the only way for the groups to be isomorphic. Our methods do allow us to obtain the following.

Corollary 2

Let \(n \in \{2, 3, \ldots \}\) and \(c_2, \ldots , c_n\in \mathbb {N}\). If \(U\le _fH_n\) and \(\pi (U)=\langle c_2e_1, \ldots , c_ne_{n-1}\rangle \), then either

  • at least two of \(c_2, \ldots , c_n\) are odd and \(U=\langle g_2^{c_2}, \ldots , g_n^{c_n}, {{\,\mathrm{FSym}\,}}(X_n)\rangle \); or

  • U is one of exactly \(2^{n-1}+1\) specific subgroups of \(H_n\).

We then extend the work in [7], where the groups \(\langle {{\,\mathrm{Alt}\,}}(X_2), g_2^c\rangle \) were shown to be 2-generated for each \(c\in \mathbb {N}\), by investigating the generation properties of each of these groups.

Notation 1 For a finitely generated group G, let \(d(G):=\min \{|S|\;:\;\langle S\rangle =G\}\).

Every finite index subgroup of a finitely generated group is itself finitely generated. In the case of \(\mathbb {Z}^n\), where \(n\in \mathbb {N}\), we have that each \(A\le _f \mathbb {Z}^n\) satisfies \(d(A)=d(\mathbb {Z}^n)\). There has been much work related to this notion, e.g. [12] and papers verifying the Nielsen–Schreier formula for families of groups other than free groups. Our second theorem states that there is a close connection between the minimal number of generators of a Houghton group and its finite index subgroups. Note that \(d(H_2)=2\) and \(d(H_n)=n-1\) for \(n\in \{3, 4, \ldots \}\).

Theorem 3

If \(U\le _f H_2\), then \(d(U)=d(H_2)\). For \(n\in \{3, 4 \ldots \}\) and \(U\le _fH_n\), we have that \(d(U)\in \{d(H_n), d(H_n)+1\}\). Furthermore, let \(\pi (U)=\langle c_2e_1, \ldots , c_ne_{n-1}\rangle \). Then \(d(U)=d(H_n)+1\) occurs exactly when both of the following conditions are met:

  1. i)

    that \({{\,\mathrm{FSym}\,}}(X_n)\le U\); and

  2. ii)

    either one or zero elements in \(\{c_2, \ldots , c_n\}\) are odd.

Our proof involves providing a generating set. Theorem 1 allows us to replace U with either \(\langle {{\,\mathrm{FSym}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \) or \(\langle {{\,\mathrm{Alt}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \). Proposition states that, in the second case, \(d(U)=n-1\). Lemma 3.5 and Lemma 3.6 combine to tell us that \(\langle {{\,\mathrm{FSym}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle =\langle {{\,\mathrm{Alt}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \) when there are distinct \(i, j \in \{2, \ldots , n\}\) that are both odd. Thus the only remaining possibility is that \(U=\langle {{\,\mathrm{FSym}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \ne \langle {{\,\mathrm{Alt}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \). Lemma 3.9 shows that in this case \(d(U)=d(H_n)+1\), by determining that the abelianization of U, with these conditions, is \(C_2\times \mathbb {Z}^{n-1}\). This complete categorisation provides us with subgroups of the Houghton groups with constant minimal number of generators on finite index subgroups.

Corollary 4

Let \(n\in \{3, 4, \ldots \}\) and define \(G_\mathbf{2}:=\langle {{\,\mathrm{Alt}\,}}(X_n), g_2^2, \ldots , g_n^2\rangle \le _f H_n\). If \(U\le _f G_\mathbf{2}\), then \(d(U)=d(G_\mathbf{2})=d(H_n)\).

2 The structure of finite index subgroups of \(H_n\)

In this section, we prove Theorem 1. The following is well known.

Lemma 2.1

Given \(U\le _fG\), there exists \(N\le _fU\) which is normal in G.

Some structure regarding finite index subgroups of \(H_n\) is known; the following is particularly useful to us.

Lemma 2.2

([5, Prop. 2.5]). Let X be a non-empty set and \({{\,\mathrm{Alt}\,}}(X)\le G\le {{\,\mathrm{Sym}\,}}(X)\). Then G has \({{\,\mathrm{Alt}\,}}(X)\) as a unique minimal normal subgroup.

Let \(n\in \{2, 3, \ldots \}\) and \(U\le _f H_n\). By Lemma 2.1, U contains a normal subgroup of \(H_n\) and so, by Lemma 2.2, \({{\,\mathrm{Alt}\,}}(X_n)\le U\). Furthermore, \(U\le _f H_n\) and so \(\pi (U)\le _f\pi (H_n)\) where \(\pi : H_n\rightarrow \mathbb {Z}^{n-1}\) sends \(g_i\) to \(e_{i-1}\) for \(i=2, \ldots , n\). Hence there exist minimal \(k_2, \ldots , k_n\in \mathbb {N}\) such that \(\pi (g_i^{k_i})\in \pi (U)\) for each \(i\in \{2, \ldots , n\}\). But \(\pi (g_i^{k_i})=k_ie_{i-1}\), and so the preimage in \(H_n\) of \(\pi (g_i^{k_i})\) is \(\{\sigma g_i^{k_i}\;:\;\sigma \in {{\,\mathrm{FSym}\,}}(X_n)\}\). If \({{\,\mathrm{FSym}\,}}(X_n)\le U\), then \(g_2^{k_2}, \ldots , g_n^{k_n} \in U\). Otherwise \({{\,\mathrm{FSym}\,}}(X_n)\cap U={{\,\mathrm{Alt}\,}}(X_n)\) and, for each \(i\in \{2, \ldots , n\}\), either \(g_i^{k_i}\in U\) or there exists \(\omega _i\in {{\,\mathrm{FSym}\,}}(X_n)\setminus {{\,\mathrm{Alt}\,}}(X_n)\) such that \(\omega _i g_i^{k_i}\in U\); since \({{\,\mathrm{Alt}\,}}(X_n)\le U\), we can specify that \(\omega _i=((1, 1)\;(1, 2))\). Then, using the minimality of \(k_2, \ldots , k_n\),

$$\begin{aligned} U=\langle g_2^{k_2}, \ldots , g_n^{k_n}, {{\,\mathrm{FSym}\,}}(X_n)\rangle \text { or }U=\langle \epsilon _2g_2^{k_2}, \ldots , \epsilon _ng_n^{k_n}, {{\,\mathrm{Alt}\,}}(X_n)\rangle \end{aligned}$$
(2)

where each \(\epsilon _i\) is either trivial or \(((1, 1)\;(1, 2))\). We now describe the isomorphism type for these subgroups. To do this, we introduce two families of finite index subgroups in \(H_n\) where \(n\in \{2, 3, \ldots \}\).

Notation 2 For any given \(n\in \{2, 3, \ldots \}\) and any \(\mathbf{c}=(c_2, \ldots , c_n)\in \mathbb {N}^{n-1}\), let \(F_\mathbf{c}:=\langle {{\,\mathrm{FSym}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \) and \(G_\mathbf{c}:=\langle {{\,\mathrm{Alt}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \).

There are some \(\mathbf{c}\in \mathbb {N}^n\) such that \(G_\mathbf{c}=F_\mathbf{c}\), e.g. if \(n\ne 2\) and \(c_2=\cdots =c_n=1\).

Notation 3 Let \(I_n:=\{\mathbf{c}=(c_2, \ldots , c_n)\in \mathbb {N}^n\;:\;G_\mathbf{c}\ne F_\mathbf{c}\}\).

Lemma 3.5 and Lemma 3.6 together describe \(I_n\). This description, together with (2), yields Corollary 2.

Notation 4 For \(n\in \{2, 3, \ldots \}\) and \(i\in \{1, \ldots , n\}\), let \(R_i:=\{(i,m)\;:\;m\in \mathbb {N}\}\subset X_n\).

Proposition 2.3

Let \(n\in \{3, 4, \ldots \}\), \(\mathbf{c}=(c_2, \ldots , c_n)\in I_n\), each \(\epsilon _2, \ldots , \epsilon _n\) be either trivial or \(((i, 1)\;(i, 2))\), and \(U=\langle {{\,\mathrm{Alt}\,}}(X_n), \epsilon _ig_i^{c_i}\;:\; i=2, \ldots , n, \rangle \le _f H_n\). Then U is isomorphic to \(G_\mathbf{c}\).

Proof

If there are \(i\ne j\) such that \(c_i=c_j=1\), then \({{\,\mathrm{FSym}\,}}(X_n)\le G_\mathbf{c}\) and so \(G_\mathbf{c}= F_\mathbf{c}\), i.e. \(\mathbf{c}\not \in I_n\).

Now consider if \(c_i=1\) for some i and that \(\epsilon _i =((i, 1)\;(i, 2))\). Then \(\epsilon _ig_i\) fixes (i, 1). Moreover, we have \((i, m)\epsilon _ig_i=(i,m-1)\) for all \(m\ge 3\), \((i,2)\epsilon _ig_i=(1, 1)\), \((1, m)\epsilon _i=(1,m+1)\) for all \(m\ge 1\), and that \(\epsilon _ig_i\) fixes all other points in \(X_n\). This allows us to relabel \(X_n\) so that the point (i, 1) becomes part of another ray \(j\in \{2, \ldots , n\}\).

If \(c_i\ne 1\) and \(\epsilon _i=((i,1)\;(i,2))\), then we can relabel \(R_i\) as \(R_i'\) by swapping the labels on the points (i, 1) and (i, 2) so that \(\epsilon _ig_i^{c_i}\) acts on \(R_i'\cup R_1\) in the same way as \(g_i^{c_i}\) acts on \(R_i\cup R_1\). \(\square \)

Lemma 2.4

Let \(k\in \{2, 3, \ldots \}\), \(\epsilon =((1, 1)\;(1, 2))\), and \(U=\langle {{\,\mathrm{Alt}\,}}(X_n), \epsilon g_2^k \rangle \le _f H_2\). Then U is isomorphic to \(\langle {{\,\mathrm{Alt}\,}}(X_2), g_2^k\rangle \).

Proof

In a similar way to the preceding proof, we can relabel \(R_2\) as \(R_2'\) (by swapping the labels on the points (2, 1) and (2, 2)) so that \(\epsilon g_2^k\) acts on \(R_1\cup R_2'\) in the same way that \(g_2^k\) acts on \(R_1\cup R_2\). \(\square \)

Remark 2.5

The preceding lemma may hold for \(k=1\). Such an isomorphism would not be induced by a permutation of \(X_2\), however: any element g in \(\langle {{\,\mathrm{Alt}\,}}(X_2), ((1, 1)\;(1, 2)) t\rangle \) with \(\pi (g)=1\) cannot have an infinite orbit equal to \(X_2\).

3 Generation of the groups \(F_\mathbf{c}\) and \(G_\mathbf{c}\)

We start with the case of \(H_2\), first dealing with the ‘exceptional’ case.

Lemma 3.1

The group \(\langle {{\,\mathrm{Alt}\,}}(X_2), ((1, 1)\;(1, 2))g_2\rangle \) is 2-generated.

Proof

Note that \(\langle {{\,\mathrm{Alt}\,}}(X_2), ((1, 1)\;(1, 2))g_2\rangle \cong \langle {{\,\mathrm{Alt}\,}}(\mathbb {Z}), (1\;2)t\rangle \le {{\,\mathrm{FSym}\,}}(\mathbb {Z})\rtimes \langle t\rangle \) where \(t: z\rightarrow z+1\) for all \(z\in \mathbb {Z}\). Our aim will be to show that \(S=\{(0\;1\;2), (0\;1)t\}\) generates \(\langle {{\,\mathrm{Alt}\,}}(\mathbb {Z}), (1\;2)t\rangle \). To do this, we will use \(\langle S\rangle \) to construct all 3-cycles \((0\;a\;b)\) with \(0<a<b\). With these elements we can then also produce every 3-cycle of the form \((a\;b\;c)\) where \(0<a<b<c\), by conjugating \((0\;a\;b)\) by \((0\;c\;2c)\). Then every 3-cycle in \({{\,\mathrm{Alt}\,}}(\mathbb {N}\cup \{0\})\), or its inverse, is accounted for by such elements. Conjugation by some power of \((0\;1)t\) yields every 3-cycle in \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\), and we recall that \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\) is generated by the set of all 3-cycles with support in \(\mathbb {Z}\).

In addition to our first simplification, note that it is sufficient to show that \((0, 1, k+1)\in \langle (0\;1\;2), (0\;1)t\rangle \) for every \(k\in \mathbb {N}\). This is because any 3-cycle \((0\;a\;b)\) with \(0<a<b\) will be conjugate, by \((0\;1)t\), to an element of the form \((0, 1, k+1)\). We start with \(\sigma _1:=(0, k, k+1)\). If \(k=1\), then we are done. Otherwise, conjugate \(\sigma _1^{-1}\) by \((0, k-1, k)\) to obtain \(\sigma _2=(0, k-1, k+1)\). If \(k>2\), conjugate \(\sigma _2^{-1}\) by \((0, k-2, k-1)\) to obtain \(\sigma _3\). Continuing in this way yields the result. \(\square \)

The following results conclude the \(H_2\) case. We include the proof here as we will adapt it for \(F_\mathbf{c}\) and \(G_\mathbf{c}\). The notation \(\Omega _{k}:=\{1, \ldots , k\}\) is helpful for these results.

Lemma 3.2

([5, Lem. 3.6]). Let \(k\in \{3, 4, \ldots \}\) and \(t: z\rightarrow z+1\) for all \(z\in \mathbb {Z}\). Then \(\langle {{\,\mathrm{Alt}\,}}(\mathbb {Z}), t^k\rangle \) is generated by \({{\,\mathrm{Alt}\,}}(\Omega _{2k}) \cup \{t^k\}\).

Lemma 3.3

([5, Lem. 3.7]). Let \(k\in \mathbb {N}\) and \(t: z\rightarrow z+1\) for all \(z\in \mathbb {Z}\). Then \(G_k:=\langle {{\,\mathrm{Alt}\,}}(\mathbb {Z}), t^k\rangle \) and \(F_k:=\langle {{\,\mathrm{FSym}\,}}(\mathbb {Z}), t^k\rangle \) are 2-generated.

Proof

We start with \(G_k\). We will show that we can find, for each \(k\in \{3, 4, \ldots \}\), an \(\alpha _k\in {{\,\mathrm{Alt}\,}}(\mathbb {Z})\) such that \(\langle t^k, \alpha _k\rangle \) contains \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\). Then, for \(d\in \{1, 2, 3\}\), \(\langle t^d, \alpha _6\rangle \) contains \(\langle t^6, \alpha _6\rangle \) and so contains \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\). We can therefore fix some \(k\in \{3, 4, \ldots \}\), meaning all 3-cycles in \({{\,\mathrm{Alt}\,}}(\Omega _{2k})\) are conjugate.

Let \(r=\left( {\begin{array}{c}2k\\ 3\end{array}}\right) \) and let \(\omega _1, \ldots , \omega _r\) be a choice of distinct 3-cycles in \({{\,\mathrm{Alt}\,}}(\Omega _{2k})\) with \(\omega _i\ne \omega _j^{-1}\) for every \(i,j \in \{1, \ldots , r\}\). Thus \(\langle \omega _1, \ldots , \omega _r\rangle ={{\,\mathrm{Alt}\,}}(\Omega _{2k})\). Set \(\sigma _0=(1\;3)\) and \(\sigma _{r+1}=(2\;3)\) and note that \((\sigma _{r+1}^{-1}\sigma _0^{-1}\sigma _{r+1})\sigma _0=(1\;2\;3)\). Now choose \(\sigma _1, \ldots , \sigma _{r}\in {{\,\mathrm{Alt}\,}}(\Omega _{2k})\) so that for each \(m\in \{1,\ldots , r\}\), we have \(\sigma _m^{-1}(1\;2\;3)\sigma _m=\omega _m\).

Let \(\alpha _k:=\prod _{i=0}^{r+1}t^{-2ik}\sigma _it^{2ik}\) and, for \(m\in \{1,\ldots , r+1\}\), let \(\beta _m:=t^{2mk}\alpha _k t^{-2mk}\). Then \(\beta _{r+1}^{-1}\alpha _k^{-1}\beta _{r+1}\alpha _k=\sigma _{r+1}^{-1}\sigma _0^{-1}\sigma _{r+1}\sigma _0=(1\;2\;3)\) and, for \(m\in \{1, \ldots , r\}\), we have that \(\beta _m^{-1}(1\;2\;3)\beta _m=\omega _m\). Hence \(\langle \alpha _k, t^k\rangle =G_k\).

We can now adapt the element \(\alpha _k\) to an element \(\gamma _k\) so, for each \(k\in \{3, 4, \ldots \}\), we have that \(\langle \gamma _k, t^k\rangle =F_k\). One way to do so is to set \(\gamma _k:=(\sigma _0')(\prod _{i=1}^{r+1}t^{-2ik}\sigma _it^{2ik})\) where \(\sigma _0'\!=\!(1\;3)(4\;5)\). This means that \([\sigma _{r+1}, \sigma _0']\) still equals \((1\;2\;3)\), and so \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\le \langle \gamma _k, t^k\rangle \), but that \({{\,\mathrm{Alt}\,}}(\mathbb {Z})\cup \gamma _k({{\,\mathrm{Alt}\,}}(\mathbb {Z}))={{\,\mathrm{FSym}\,}}(\mathbb {Z})\le \langle \gamma _k, t^k\rangle \) as well. \(\square \)

We now work with a fixed \(n\in \{3, 4, \ldots \}\). The following well-known commutator identities will be helpful for a few of our remaining proofs.

$$\begin{aligned}{}[a, bc]=a^{-1}(bc)^{-1}a(bc)=a^{-1}c^{-1}acc^{-1}a^{-1}b^{-1}abc=[a, c][a, b]^c \end{aligned}$$
(3)
$$\begin{aligned}{}[ab, c]=(ab)^{-1}c^{-1}(ab)c=b^{-1}(a^{-1}c^{-1}ac)bb^{-1}c^{-1}bc=[a, c]^b[b, c] \end{aligned}$$
(4)

Remark 3.4

Let \(a, b, c \in H_n\) and \(\mathrm {sgn}\) denote the sign function on \({{\,\mathrm{FSym}\,}}(X_n)\). From the identities (3) and (4) above, we have that \(\mathrm {sgn}([a, bc])=\mathrm {sgn}([a, b])\mathrm {sgn}([a, c])\) and \(\mathrm {sgn}([ab, c])=\mathrm {sgn}([a, c])\mathrm {sgn}([b, c])\).

Recall that \(\mathbf{c}\in \mathbb {N}^{n-1}\) is in \(I_n\) if and only if \(G_\mathbf{c}\ne F_\mathbf{c}\).

Lemma 3.5

Let \(\mathbf{c}=(c_2, \ldots , c_n)\in \mathbb {N}^n\). We have that \(\mathbf{c}\in I_n\) if and only if \([g_i^{c_i}, g_j^{c_j}]\in {{\,\mathrm{Alt}\,}}(X_n)\) for all \(i, j \in \{2, \ldots , n\}\).

Proof

If there exist \(i, j\in \{2, \ldots , n\}\) such that \([g_i^{c_i}, g_j^{c_j}]\not \in {{\,\mathrm{Alt}\,}}(X_n)\), then \(G_\mathbf{c}\) contains an odd permutation and so \(G_\mathbf{c}= F_\mathbf{c}\). Recall that \(G_\mathbf{c}=\langle {{\,\mathrm{Alt}\,}}(X_n), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \) and that we have a homomorphism \(\pi : H_n\rightarrow \mathbb {Z}^{n-1}\) induced by sending \(g_i\) to \(e_{i-1}\) for \(i=2, \ldots , n\). Then \(\pi (G_\mathbf{c})\le _f\pi (H_n)=\mathbb {Z}^{n-1}\), and so \(\pi (G_\mathbf{c})\) is free abelian of rank \(n-1\). Define \(S:=\{g_2^{c_2}, \ldots , g_n^{c_n}\}\) so that \(\langle S\cup {{\,\mathrm{Alt}\,}}(X_n)\rangle = G_\mathbf{c}\) and \(\pi (S)\) is a linearly independent set. Let \(t_i:=\pi (g_i^{c_i})\) for \(i=2, \ldots , n\). Thus

$$\begin{aligned} \pi (\langle S\rangle )=\langle t_2, \ldots , t_n\mid R\rangle \text { where }R=\{[t_i, t_j]\;:\;i\ne j\}. \end{aligned}$$
(5)

Now let \(\alpha \in \langle S\rangle \cap {{\,\mathrm{FSym}\,}}(X_n)\). We will show that \(\alpha \in {{\,\mathrm{Alt}\,}}(X_n)\), meaning that \(\mathbf{c}\in I_n\). First, express \(\alpha \) as a word w in \(S^{\pm 1}\). Thus \(\alpha =w(g_2^{c_2}, \ldots , g_n^{c_n})\) and \(\pi (\alpha )=w(t_2, \ldots , t_n)\) for the same word w. But also \(\pi (\alpha )=\underline{0}\) which, from (5), means that \(w(t_2, \ldots , t_n)\) is in the normal closure of R. Thus \(w(t_2, \ldots , t_n)\) is a product of conjugates of powers of commutators in terms of \(t_2, \ldots , t_n\). Then \(\alpha =w(g_2^{c_2}, \ldots , g_n^{c_n})\) must also be a product of conjugates of powers of commutators in terms of \(g_2^{c_2}, \ldots , g_n^{c_n}\). Our assumption that \([g_i^{c_i}, g_j^{c_j}]\in {{\,\mathrm{Alt}\,}}(X_n)\) for all \(i, j \in \{2, \ldots , n\}\) together with Remark 3.4 means that \(\alpha \in {{\,\mathrm{Alt}\,}}(X_n)\). \(\square \)

The following gives a clearer description of the set \(I_n\).

Lemma 3.6

Let \(i, j\in \{2, \ldots , n\}\) and \(c_i, c_j\in \mathbb {N}\). Then \([g_i^{c_i}, g_j^{c_j}]\not \in {{\,\mathrm{Alt}\,}}(X_n)\) if and only if \(c_i\) and \(c_j\) are both odd.

Proof

Note that \([g_i, g_j]\in {{\,\mathrm{FSym}\,}}(X_n)\setminus {{\,\mathrm{Alt}\,}}(X_n)\) whereas \([g_i^2, g_j^2], [g_i, g_j^2], [g_i^2, g_j] \in {{\,\mathrm{Alt}\,}}(X_n)\). Now, by repeatedly applying Remark 3.4, we can reduce \(c_i\) and \(c_j\) to be either 1 or 2, depending on whether they are odd or even respectively. \(\square \)

Combining Lemma 3.5 and Lemma 3.6, we see that \(\mathbf{c}=(c_2, \ldots , c_n)\in \mathbb {N}^{n-1}\) is in \(I_n\) if and only if at least two of \(c_2, \ldots , c_n\) are odd.

Proposition 3.7

Given \(\mathbf{c}=(c_2, \ldots , c_n)\in \mathbb {N}^{n-1}\), we have that \(d(G_\mathbf{c})=d(H_n)\).

Proof

We start with the case that \(\mathbf{c}\in I_n\). By possibly relabelling the branches of \(X_n\), set \(c_3:=\max \{c_2, \ldots , c_n\}\). Thus \(c_3\ne 1\), as otherwise \(c_2=c_3=1\) and \(\mathbf{c}\not \in I_n\). Let \(k:=2c_3\) and \(\Omega ^*:=\{(1, 1), \ldots , (1, 2k)\}\). Thus \(2k\ge 8\) and all 3-cycles in \({{\,\mathrm{Alt}\,}}(\Omega ^*)\) are conjugate in \({{\,\mathrm{Alt}\,}}(\Omega ^*)\). We claim that \({{\,\mathrm{Alt}\,}}(X_n)\le \langle {{\,\mathrm{Alt}\,}}(\Omega ^*), g_2^{c_2}, \ldots , g_n^{c_n}\rangle \). By Lemma 3.2, \(\langle {{\,\mathrm{Alt}\,}}(\Omega ^*), g_3^{c_3}\rangle \) is a subgroup of \(H_n\) which generates \(\langle {{\,\mathrm{Alt}\,}}(R_1\cup R_3), g_3^{c_3}\rangle \cong \langle {{\,\mathrm{Alt}\,}}(\mathbb {Z}), t^{c_3}\rangle \). Most importantly, \(\langle {{\,\mathrm{Alt}\,}}(\Omega ^*), g_3^{c_3}\rangle \) contains \({{\,\mathrm{Alt}\,}}(R_1)\). Now, given any \(\sigma \in {{\,\mathrm{FSym}\,}}(X_n)\), there exists a word of the form \(g_2^{d_2}\ldots g_n^{d_n}\) for some \(d_i\in c_i\mathbb {N}\) which conjugates \(\sigma \) to \(\sigma '\), where \({{\,\mathrm{supp}\,}}(\sigma ')\subset R_1\). Hence \(\langle {{\,\mathrm{Alt}\,}}(\Omega ^*), g_2^{c_2}, \ldots , g_n^{c_n}\rangle =G_\mathbf{c}\). Our next aim is to adapt the proof of Lemma 3.3. In particular, we will show that there exists a \(\sigma \in {{\,\mathrm{Alt}\,}}(X_n)\) such that \(\langle \sigma g_2^{c_2}, g_3^{c_3}, \ldots , g_n^{c_n}\rangle =G_\mathbf{c}\). For any \(\omega \in {{\,\mathrm{FSym}\,}}(X_n)\), let \(\sigma _\omega :=[\omega g_2^{c_2}, g_3^{2k}]\), which we can calculate using (4) as

$$\begin{aligned}{}[\omega , g_3^{2k}]^{g_2^{c_2}}[g_2^{c_2}, g_3^{2k}]=g_2^{-c_2}(\omega ^{-1}g_3^{-2k}\omega g_3^{2k})g_2^{c_2}[g_2^{c_2}, g_3^{2k}]. \end{aligned}$$
(6)

Observe that \({{\,\mathrm{supp}\,}}([g_2^{c_2}, g_3^{2k}])\subseteq \{(1, 1), \ldots , (1, c_2+2k)\}\) and, since \(k=2c_3>2c_2\), that \((1, 4k+1), (1, 4k+3)\not \in {{\,\mathrm{supp}\,}}([g_2^{c_2}, g_3^{k}])\). Let \(\delta _0:=((1, 2k-c_2+1), (1, 2k-c_2+3))\) and note, using (6), that \({{\,\mathrm{supp}\,}}(\sigma _{\delta _0})\cap \{(1, m)\;:\;m=4k+2 \text { or }m\ge 4k+4\}=\emptyset \) and \(\sigma _{\delta _0}\) swaps \((1, 4k+1)\) and \((1, 4k+3)\). We now restrict to those \(\omega \in {{\,\mathrm{Alt}\,}}(X_n)\) with

$${{\,\mathrm{supp}\,}}(\omega \delta _0)\subset \bigcup _{j\in \mathbb {N}}(\Omega ^*)g_3^{-4jk}.$$

The following implications of this restriction are all important in relation to (6):

  • \({{\,\mathrm{supp}\,}}(\sigma _\omega )\subset R_1\cup R_3\);

  • \({{\,\mathrm{supp}\,}}(\omega ^{-1})\cap {{\,\mathrm{supp}\,}}(g_3^{-2k}\omega g_3^{2k})=\emptyset \);

  • \((x)g_2^{-c_2}\omega ^{-1}g_2^{c_2}=(x)\omega ^{-1}\) for all \(x\in R_3\); and

  • \((x)g_2^{-c_2}(g_3^{-2k}\omega g_3^{2k})g_2^{c_2}=(x)g_3^{-2k}\omega g_3^{2k}\) for all \(x\in R_3\).

Now, by mimicking the proof of Lemma 3.3, we can choose one such \(\omega \), which we will denote by \(\alpha \), so that \({{\,\mathrm{Alt}\,}}(R_1)\le \langle g_3^k, \sigma _{\alpha }\rangle \). Let \(r=\left( {\begin{array}{c}2k\\ 3\end{array}}\right) \) and \(\omega _1, \ldots , \omega _r\) be a set of 3-cycles such that \(\langle \omega _1, \ldots , \omega _r\rangle ={{\,\mathrm{Alt}\,}}(\Omega ^*)\). Choose \(\sigma _1, \ldots , \sigma _r\in {{\,\mathrm{Alt}\,}}(\Omega ^*)\) such that \(\sigma _i^{-1}((1, 1)\,(1, 2)\,(1, 3))\sigma _i=\omega _i\) for \(i=1, \ldots , r\). Also set \(\sigma _{r+1}:=((1, 2)\,(1, 3))\) and \(\sigma _0:=((1, 1)\,(1, 3))\). Define \(\delta _i:=g_3^{4ki}\sigma _ig_3^{-4ki}\) for \(i=1, \ldots , r+1\) and \(\alpha :=\delta _0\ldots \delta _{r+1}\). Define \(\beta _m:=g_3^{-4mk}\sigma _\alpha g_3^{4mk}\) for \(m=-1, \ldots , r+1\). Then, as in Lemma 3.3, we have that \([\beta _{r+1}, \beta _{-1}]=[\sigma _{d+1}, \sigma _0]\) and \(\beta _i[\sigma _{d+1}, \sigma _0]\beta _i^{-1}=\omega _i\) for \(i=1, \ldots , r\).

We end our proof with the case where \(\mathbf{c}\not \in I_n\). From Lemma 3.5 and Lemma 3.6, there then exist \(i, j \in \{2, \ldots , n\}\) such that \([g_i^{c_i}, g_j^{c_j}]\in {{\,\mathrm{FSym}\,}}(X_n)\setminus {{\,\mathrm{Alt}\,}}(X_n)\). Using the element \(\sigma _\alpha \in {{\,\mathrm{Alt}\,}}(X_n)\) defined above, we claim that \(\langle \sigma _\alpha g_2^{c_2}, g_3^{c_3}, \ldots , g_n^{c_n}\rangle = F_\mathbf{c}\). We note that if \([g_i^{c_i}, g_j^{c_j}]\in {{\,\mathrm{FSym}\,}}(X_n)\setminus {{\,\mathrm{Alt}\,}}(X_n)\), then, by Remark 3.4, so is \([\sigma g_i^{c_i}, g_j^{c_j}]\) for any \(\sigma \in {{\,\mathrm{FSym}\,}}(X_n)\). Hence \(\langle \sigma _\alpha g_2^{c_2}, g_3^{c_3}, \ldots , g_n^{c_n}\rangle \) contains both \({{\,\mathrm{Alt}\,}}(X_n)\) and an odd permutation, and so \(d(G_\mathbf{c})=d(H_n)\). \(\square \)

We now look at the final case of determining \(d(F_\mathbf{c})\) where \(\mathbf{c}\in I_n\).

Lemma 3.8

Let \(\mathbf{c}\in I_n\). Then the commutator subgroup of \(F_\mathbf{c}\) equals \({{\,\mathrm{Alt}\,}}(X_n)\).

Proof

Note that \([{{\,\mathrm{Alt}\,}}(X_n), {{\,\mathrm{Alt}\,}}(X_n)]={{\,\mathrm{Alt}\,}}(X_n)\). Let \(\sigma \prod _{i\in I} g_i^{p_i}\) and \(\omega \prod _{j\in J} g_j^{q_j}\) be in \(F_\mathbf{c}\), where \(I, J\subseteq \{2, \ldots , n\}\), \(\{p_i\;:\;i\in I\}\subset \mathbb {Z}\), \(\{q_j\;:\;j\in J\}\subset \mathbb {Z}\), and \(\sigma , \omega \in {{\,\mathrm{FSym}\,}}(X_n)\). Then their commutator, using (3), will be a product of conjugates of elements of the form \([\sigma \prod _{i\in I} g_i^{p_i}, \omega ]\) or \([\sigma \prod _{i\in I} g_i^{p_i}, g_{k}^{q_{k}}]\) for some \(k\in J\). Applying (4) to these elements will produce a product of conjugates of elements of the form

$$\begin{aligned}{}[\sigma , \omega ], [g_l^{p_l}, \omega ], [\sigma , g_k^{q_k}],\text { or }[g_l^{p_l}, g_k^{q_k}] \end{aligned}$$

for some \(l\in I\) and \(k\in J\). The first 3 of these are clearly in \({{\,\mathrm{Alt}\,}}(X_n)\). The 4th is also in \({{\,\mathrm{Alt}\,}}(X_n)\) from our assumption that \(\mathbf{c}\in I_n\) together with Lemma 3.5. Hence, from Remark 3.4, a commutator of any two elements in \(F_\mathbf{c}\) is in \({{\,\mathrm{Alt}\,}}(X_n)\). \(\square \)

Lemma 3.9

Let \({ \mathbf{c}}=(c_2, \ldots , c_n)\in I_n\). Then \(d(F_\mathbf{{c}})=d(H_n)+1\).

Proof

Recall that \(d(G_\mathbf{c})=n-1\). Including an element from \({{\,\mathrm{FSym}\,}}(X_n)\setminus {{\,\mathrm{Alt}\,}}(X_n)\) to a minimal generating set for \(G_\mathbf{c}\) therefore yields a generating set of \(F_\mathbf{c}\) of size n. So we need only show that, with the given hypothesis, no generating set of \(F_\mathbf{c}\) of size \(n-1\) exists. From Lemma , the abelianization of \(F_\mathbf{c}\) is \(\mathbb {Z}^{n-1}\times C_2\). Thus \(F_\mathbf{c}\) cannot be generated by \(n-1\) elements. \(\square \)