1 Introduction and Preliminaries

Let \(X_n\) be a chain with n elements, say \(X_n=\{1<2<\cdots <n\}\). As usual, we denote by \(\mathcal {T}_n\) the monoid of all full transformations of a finite set \(X_n\). We say that a transformation \(\alpha \in \mathcal {T}_n\) is order preserving if \(x\le y\) implies \(x\alpha \le y\alpha \), for all \(x,y\in X_n\). Denote by \(\mathcal {O}_n\) the submonoid of \(\mathcal {T}_n\) of all full order-preserving transformations of \(X_n\).

Let \(c=(c_1,c_2,\ldots ,c_t)\) be a sequence of t (\(t\ge 0\)) elements from the chain \(X_n\). We say that c is cyclic if there exists not more than one index \(i\in \{1,\ldots ,t\}\) such that \(c_i>c_{i+1}\), where \(c_{t+1}\) denotes \(c_1\). We say that \(\alpha \in \mathcal {T}_n\) is orientation preserving if the sequence of its image \((1\alpha ,2\alpha ,\ldots ,n\alpha )\) is cyclic. Denote by \(\mathcal {OP}_n\) the submonoid of \(\mathcal {T}_n\) of all full orientation-preserving transformations of \(X_n\).

Let S be a semigroup. Denote by \(S^1\) the monoid obtained from S through the adjoining of an identity if S has none and exactly S otherwise. Recall the definition of Green’s equivalence relations \({\mathscr {R}}\), \({\mathscr {L}}\), \({\mathscr {H}}\), and \({\mathscr {J}}\): for all \(u, v\in S\),

$$\begin{aligned} \begin{array}{rcl} u{\mathscr {R}} v &{}\text{ if } \text{ and } \text{ only } \text{ if }&{} uS^1=vS^1;\\ u{\mathscr {L}} v &{}\text{ if } \text{ and } \text{ only } \text{ if }&{} S^1u=S^1v;\\ u{\mathscr {H}} v &{}\text{ if } \text{ and } \text{ only } \text{ if }&{} u{\mathscr {R}} v \text{ and } u{\mathscr {L}} v;\\ u{\mathscr {J}} v &{}\text{ if } \text{ and } \text{ only } \text{ if }&{} S^1uS^1=S^1vS^1. \end{array} \end{aligned}$$

Associated to Green’s relation \({\mathscr {J}}\), there is a quasi-order \(\le _{\mathscr {J}}\) on S defined by

$$\begin{aligned} u\le _{\mathscr {J}} v \quad \text{ if } \text{ and } \text{ only } \text{ if }\quad S^1uS^1\subseteq S^1vS^1, \end{aligned}$$

for all \(u, v\in S\). Notice that, for every \(u, v\in S\), we have \(u\,{\mathscr {J}}\,v\) if and only if \(u\le _{\mathscr {J}}v\) and \(v\le _{\mathscr {J}}u\). Denote by \(J_{u}^S\) the \({\mathscr {J}}\)-class of the element \(u\in S\). As usual, a partial order relation \(\le _{\mathscr {J}}\) is defined on the quotient set \(S/{\mathscr {J}}\) by putting \(J_{u}^S\le _{\mathscr {J}}J_{v}^S\) if and only if \(u\le _{\mathscr {J}}v\), for all \(u, v\in S\). Given a subset A of S and \(u\in S\), we denote by E(A) set of idempotents of S belonging to A and by \(L^S_u\), \(R^S_u\), and \(H^S_u\) the \({\mathscr {L}}\)-class, \({\mathscr {R}}\)-class, and \({\mathscr {H}}\)-class of u, respectively. For general background on Semigroup Theory, we refer the reader to Howie’s book [8].

Let \(\mathcal {M}_n\) denotes any of the monoids \(\mathcal {T}_n\), \(\mathcal {O}_n\), or \(\mathcal {OP}_n\). Then \(\mathcal {M}_n\) is regular. The Green relations \({\mathscr {L}}\) and \({\mathscr {R}}\) of \(\mathcal {M}_n\) can be characterized by \(\alpha {\mathscr {L}} \beta \) if and only if \(\mathop {\mathrm {im}}\nolimits (\alpha )=\mathop {\mathrm {im}}\nolimits (\beta )\), for all \(\alpha ,\beta \in \mathcal {M}_n\), and \(\alpha {\mathscr {R}} \beta \) if and only if \(\mathop {\mathrm {ker}}\nolimits (\alpha )=\mathop {\mathrm {ker}}\nolimits (\beta )\), for all \(\alpha ,\beta \in \mathcal {M}_n\). Regarding the Green relation \({\mathscr {J}}\), we have \(\alpha \le _{{\mathscr {J}}} \beta \) if and only if \(|\mathop {\mathrm {im}}\nolimits (\alpha )|\le |\mathop {\mathrm {im}}\nolimits (\beta )|\) and so \(\alpha {{\mathscr {J}}} {\beta }\) if and only if \(|\mathop {\mathrm {im}}\nolimits (\alpha )|=|\mathop {\mathrm {im}}\nolimits (\beta )|\), for all \(\alpha ,\beta \in \mathcal {M}_n\). It follows that the partial order \(\le _{\mathscr {J}}\) on the quotient \(\mathcal {M}_n/{\mathscr {J}}\) is linear. More precisely, letting

$$\begin{aligned} J_r^{\mathcal {M}_n}=\{\alpha \in {\mathcal {M}_n}: |\mathop {\mathrm {im}}\nolimits (\alpha )|=r\}, \end{aligned}$$

i.e., the \({\mathscr {J}}\)-class of the transformations of image size r (called the rank of the transformations) of \(\mathcal {M}_n\), for \(1\le r\le n\), we have

$$\begin{aligned} \mathcal {M}_n/{\mathscr {J}}=\{J_1^{\mathcal {M}_n}\le _{\mathscr {J}}J_2^{\mathcal {M}_n}\le _{\mathscr {J}}\dots \le _{\mathscr {J}}J_n^{\mathcal {M}_n}\}. \end{aligned}$$

See [1, 2, 7, 8] for more details.

Since \({\mathcal {M}_n}/ {\mathscr {J}}\) is a chain, the sets

$$\begin{aligned} \mathcal {M}(n,r)=\{\alpha \in \mathcal {M}_n: |\mathop {\mathrm {im}}\nolimits (\alpha )|\le r\}=J_1^{\mathcal {M}_n}\cup J_2^{\mathcal {M}_n}\cup \cdots \cup J_r^{\mathcal {M}_n}, \end{aligned}$$

with \(1\le r\le n\), constitute all the non-null ideals of \(\mathcal {M}_n\) (see [4, Note of p. 181]). Notice that \(\mathcal {M}(n,r)=\mathcal {T}(n,r)\) if \(\mathcal {M}_n=\mathcal {T}_n\); \(\mathcal {M}(n,r)=\mathcal {O}(n,r)\) if \(\mathcal {M}_n=\mathcal {O}_n\); \(\mathcal {M}(n,r)=\mathcal {OP}(n,r)\) if \(\mathcal {M}_n=\mathcal {OP}_n\).

Notice that every principal factor of \(\mathcal {M}_n\) associated with the maximum \({\mathscr {J}}\)-class \(J_r^{\mathcal {M}_n}\) is the Rees quotient \(\mathcal {M}(n,r)/\mathcal {M}(n,r-1)\) (\(2\le r\le n\)), which we denote by \(\mathcal {P}^{\mathcal {M}_n}_r\). It is usually convenient to think of \(\mathcal {P}^{\mathcal {M}_n}_r\) as \(J_r^{\mathcal {M}_n}\cup \{0\}\), and the product of two elements of \(\mathcal {P}^{\mathcal {M}_n}_r\) is taken to be zero if it falls in \(\mathcal {M}(n,r-1)\). \(\mathcal {P}^{\mathcal {M}_n}_r\) is a completely 0-simple semigroup.

As usual, the rank of a finite semigroup S is defined by \(\mathop {\mathrm {rank}}\nolimits S=\min \{|A|: A\subseteq S, \langle A \rangle =S \}\). If S is generated by its set E of idempotents, then the idempotent rank of S is defined by \(\mathop {\mathrm {idrank}}\nolimits S=\min \{|A|: A\subseteq E, \langle A \rangle =S \}\). If S is generated by its set N of nilpotents, then the nilpotent rank of S is defined by \(\mathop {\mathrm {nilrank}}\nolimits S=\min \{|A|: A\subseteq N, \langle A \rangle =S \}\). Clearly, \(\mathop {\mathrm {rank}}\nolimits S\le \mathop {\mathrm {idrank}}\nolimits S \) and \(\mathop {\mathrm {rank}}\nolimits S\le \mathop {\mathrm {nilrank}}\nolimits S \).

In [6], Gomes and Howie showed that both the rank and the idempotent rank of \(\mathcal {P}^{\mathcal {T}_n}_{n-1}\) are equal to \(n(n-1)/2\). This result was later generalized by Howie and McFadden [9] who showed that the rank and idempotent rank of \(\mathcal {P}^{\mathcal {T}_n}_{r}\) are both equal to S(nr), the Stirling number of the second kind, for \(2\le r\le n-1\). Yang [10] showed that the nilpotent rank of \(\mathcal {P}^{\mathcal {T}_n}_{r}\) is also S(nr), \(2\le r\le n-1\).

The rank and idempotent rank of \(\mathcal {P}^{\mathcal {O}_n}_{n-1}\) were shown to be n and \(2n-2\), respectively, by Gomes and Howie [7]. Garba [5] generalized this result by showing that both the rank and the idempotent rank of the principal factor \(\mathcal {P}^{\mathcal {O}_n}_{r}\) are equal to \({n\atopwithdelims ()r}\), for \(2\le r\le n-2\). Yang [11] showed that the nilpotent rank of the principal factor \(\mathcal {P}^{\mathcal {O}_n}_{r}\) are also equal to \({n\atopwithdelims ()r}\), for \(3\le r\le n-1\).

Regarding the semigroup \(\mathcal {OP}_n\), Zhao [12] showed that both the rank and the idempotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{n-1}\) are equal to \({n\atopwithdelims ()2}\). Recently, Zhao and Fernandes [13] showed that the rank and the idempotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{r}\) are equal to \({n\atopwithdelims ()r}\), for \(2\le r\le n-1\). In this paper, we investigate the nilpotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{r}\), for \(2\le r\le n-1\). In Sect. 2, we characterize the structure of the minimal generating sets of \(\mathcal {OP}(n,r)\). As applications, we prove that the number of distinct minimal generating sets is \(r^n n!\). In Sect. 3, we show that the nilpotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{r}\) is equal to \({n\atopwithdelims ()r}\), for \(2\le r\le n-1\).

Remark 1

In this paper, it will always be clear from context when additions are taken modulo n (or modulo t where t is the number of elements of any sequence).

Throughout this paper, for simplicity, we always assume that \(n\ge 3\).

2 The Minimal Generating Sets of \(\mathcal {OP}(n,r)\)

Let \(\alpha \in \mathcal {T}_n\). As usual, we write \(\mathop {\mathrm {im}}\nolimits (\alpha )\) for the image of \(\alpha \). The kernel of \(\alpha \) is the equivalence \(\mathop {\mathrm {ker}}\nolimits (\alpha )=\{(x,y)\in X_n \times X_n: x\alpha =y\alpha \}\). The equivalence classes of \(X_n\) with respect to \(\mathop {\mathrm {ker}}\nolimits (\alpha )\) are called the kernel classes of \(\alpha \).

We denote by [ik] the set \(\{i,i+1, \dots , k-1,k\}\) for \(i,k\in X_n\). A set \(K\subseteq X_n\) is convex if K has the form \([i,i+t]\), for some \(i\in X_n\) and \(0\le t\le n-1\). We shall refer to an equivalence \(\pi \) on \(X_n\) as convex if its classes are convex subsets of \(X_n\), and we shall say that \(\pi \) is of weight r if \(|X_n/\pi |=r\).

It is known that that every kernel \(\mathop {\mathrm {ker}}\nolimits (\alpha )\) of \(\alpha \in \mathcal {OP}_n\) is convex (see [2]). Let \(2\le r\le n-1\). Then, given a transformation \(\alpha \in \mathcal {OP}_n\) of rank r, the kernel \(\mathop {\mathrm {ker}}\nolimits (\alpha )\) is convex. Therefore, we may establish a one-to-one correspondence between the collection of all subsets of \(X_n\) of cardinality r (which consists of all possible images of elements of rank r of \(\mathcal {OP}_n\)) and the collection of all possible convex kernels of elements of rank r of \(\mathcal {OP}_n\) as follows: associate to each r-set \(\{a_1, a_2,\ldots , a_r\}\), with \(1\le a_1<a_2<\cdots <a_r\le n\), the r-partition \(\{A_1,\ldots , A_r\}\) of \(X_n\) defined by

$$\begin{aligned} A_i=\{a_i,a_i+1,\ldots ,a_{i+1}-1\},~\hbox { for}\ 1\le i\le r \end{aligned}$$
(1)

(notice that \(a_{r+1}=a_1\) by Remark 1). Thus, the \({\mathscr {J}}\)-class \(J^{\mathcal {OP}_n}_r\) of \(\mathcal {OP}_n\) (and of \(\mathcal {OP}(n,r)\)) contains \(\left( {\begin{array}{c}n\\ r\end{array}}\right) \) \({\mathscr {R}}\)-classes and \(\left( {\begin{array}{c}n\\ r\end{array}}\right) \) \({\mathscr {L}}\)-classes. See [2] for more details.

Now, for \(1\le a_1<a_2<\cdots <a_r\le n\), define

$$\begin{aligned} \varepsilon _{a_1, a_2,\ldots , a_r} = \begin{pmatrix} A_{1} &{} A_{2} &{} \cdots &{} A_{r} \\ a_{1} &{} a_{2} &{} \cdots &{} a_{r} \end{pmatrix}, \end{aligned}$$

where \(\{A_1,\ldots , A_r\}\) is the r-partition of \(X_n\) associated to \(\{a_1, a_2,\ldots , a_r\}\) as in (1). Clearly, \(\varepsilon _{a_1, a_2,\ldots , a_r}\in E(J^{\mathcal {OP}_n}_r)\). Moreover, the set

$$\begin{aligned} E_r=\left\{ \varepsilon _{a_1, a_2,\ldots , a_r}: 1\le a_1<a_2<\cdots <a_r\le n\right\} \end{aligned}$$

contains exactly one (idempotent) element from each \({\mathscr {R}}\)-class and from each \({\mathscr {L}}\)-class of \(\mathcal {OP}_n\) of rank r.

The following lemma was proved by Zhao and Fernandes [13, Proposition 2.4].

Lemma 2.1

For \(2\le r \le n-1\), the set \(E_r\) generates \(\mathcal {OP}(n,r)\).

Now, we record a well-known result, due to Miller and Clifford ([3, Theorem 2.17]).

Lemma 2.2

For any two elements ab in a semigroup S, \(ab\in R_a\cap L_b\) if and only if \(E(R_b\cap L_a)\ne \emptyset \).

Let U be a subset of \(J^{\mathcal {OP}_n}_r\). We say that U satisfies Condition \((\mathbf {R\sim L})\) if U contains exactly one element from each \({\mathscr {R}}\)-class and from each \({\mathscr {L}}\)-class of \(J^{\mathcal {OP}_n}_r\). Notice that the set \(E_r\) satisfies Condition \((\mathbf {R\sim L})\). From Lemma 2.1, we know that \(E_r\) is a generating set of \(\mathcal {OP}(n,r)\). In fact, we have the following.

Lemma 2.3

Let G be a subset of \(J^{\mathcal {OP}_n}_r\). If G satisfies Condition \((\mathbf {R\sim L})\), then \(\mathcal {OP}(n,r)=\langle G \rangle \).

Proof

We shall show that \(E_r\subseteq \langle G \rangle \) and so \(\mathcal {OP}(n,r)=\langle E_r\rangle \) by Lemma 2.1. Let \(\varepsilon \in E_r\). Since G satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\alpha \in G\) such that \(\alpha {{\mathscr {R}}} \varepsilon \). If \((\alpha , \varepsilon )\in {{\mathscr {L}}} \), then \(\alpha {{\mathscr {H}}} \varepsilon \). Notice that each \({\mathscr {H}}\)-class of \(J^{\mathcal {OP}_n}_r\) that contains an idempotent is a cyclic group of order r (by [2, Corollary 3.6]). Thus \(\varepsilon =\alpha ^r\in \langle G \rangle \). If \((\alpha , \varepsilon )\notin {{\mathscr {L}}}\), then since \(E_r\) satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\varepsilon _1\in E_r\backslash \{\varepsilon \}\) such that \((\alpha ,\varepsilon _1)\in {\mathscr {L}}\). Since G satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\alpha _1\in G\backslash \{\alpha \}\) such that \((\alpha _1, \varepsilon _{1})\in {{\mathscr {R}}}\). Notice that \(\varepsilon _1\in E(L_\alpha \cap R_{\alpha _1})\). Then, by Lemma 2.2, \(\alpha \alpha _1\in R_{\alpha }\cap L_{\alpha _1}=R_{\varepsilon }\cap L_{\alpha _1}\). If \((\alpha _1, \varepsilon )\in {\mathscr {L}}\), then \((\alpha \alpha _1) {\mathscr {H}} \varepsilon \) and so \(\varepsilon =(\alpha \alpha _1) ^r\in \langle G \rangle \). If \((\alpha _1, \varepsilon )\notin {{\mathscr {L}}}\), then since \(E_r\) satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\varepsilon _2\in E_r\backslash \{\varepsilon , \varepsilon _1 \}\) such that \((\alpha _1,\varepsilon _2)\in {\mathscr {L}}\). Since G satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\alpha _2 \in G\backslash \{\alpha , \alpha _1\}\) such that \((\alpha _2, \varepsilon _{2})\in {{\mathscr {R}}}\). Notice that \(\varepsilon _2\in E(L_{\alpha _1}\cap R_{\alpha _2})=E(L_{\alpha \alpha _1}\cap R_{\alpha _2})\). Then, by Lemma 2.2, \((\alpha \alpha _1)\alpha _2\in R_{\alpha \alpha _1}\cap L_{\alpha _2}=R_{\varepsilon }\cap L_{\alpha _2}\). If \((\alpha _2, \varepsilon )\in {\mathscr {L}}\), then \((\alpha \alpha _1\alpha _2) {\mathscr {H}} \varepsilon \) and so \(\varepsilon =(\alpha \alpha _1\alpha _2) ^r\in \langle G \rangle \). If \((\alpha _2, \varepsilon )\notin {{\mathscr {L}}}\), then since \(E_r\) satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\varepsilon _3\in E_r\backslash \{\varepsilon , \varepsilon _1, \varepsilon _2 \}\) such that \((\alpha _2,\varepsilon _3)\in {\mathscr {L}}\). Since G satisfies Condition \((\mathbf {R\sim L})\), there exists unique \(\alpha _3 \in G\backslash \{\alpha , \alpha _1,\alpha _2\}\) such that \((\alpha _3,\varepsilon _{3})\in {{\mathscr {R}}}\). Notice that \(\varepsilon _3\in E(L_{\alpha _2}\cap R_{\alpha _3})=E(L_{\alpha \alpha _1\alpha _2}\cap R_{\alpha _3})\). Then, by Lemma 2.2, \((\alpha \alpha _1\alpha _2)\alpha _3\in R_{\alpha \alpha _1\alpha _2}\cap L_{\alpha _3}=R_{\varepsilon }\cap L_{\alpha _3}\). Continuing this demonstration, since G and \(E_r\) satisfy Condition \((\mathbf {R\sim L})\), there must exist \(k\le m\) (\(m={n\atopwithdelims ()r}\)) such that \(\alpha _k\in G\backslash \{\alpha , \alpha _1, \alpha _{k-1}\}\), \((\alpha \dots \alpha _{k-1})\alpha _k\in R_{\varepsilon }\cap L_{\alpha _k}\) and \(\alpha _k {\mathscr {L}} \varepsilon \). Then \((\alpha \alpha _1\dots \alpha _k) {\mathscr {H}} \varepsilon \) and so \(\varepsilon =(\alpha \alpha _1\dots \alpha _k) ^r\in \langle G \rangle \). \(\square \)

Since \(\mathcal {OP}(n,r)\) has rank \({n\atopwithdelims ()r}\) (see [13, Theorem 2.7]), a generating set of \(\mathcal {OP}(n,r)\) with \({n\atopwithdelims ()r}\) elements is a minimal generating set. Moreover, if \(\alpha \) is an element of \(\mathcal {OP}(n,r)\) of rank r and \(\beta \) and \(\gamma \) are two elements of \(\mathcal {OP}(n,r)\) such that \(\alpha =\beta \gamma \), then \(\mathop {\mathrm {ker}}\nolimits (\alpha )=\mathop {\mathrm {ker}}\nolimits (\beta )\) and \(\mathop {\mathrm {im}}\nolimits (\alpha )=\mathop {\mathrm {im}}\nolimits (\gamma )\). Then any generating set of \(\mathcal {OP}(n,r)\) with \({n\atopwithdelims ()r}\) elements be the subset having exactly one element from each \({\mathscr {R}}\)-class and from each \({\mathscr {L}}\)-class of rank r. These observations, together with the Lemma 2.3, prove the following result:

Theorem 2.4

Let M be a subset of \(\mathcal {OP}(n,r)\) with \({n\atopwithdelims ()r}\) elements. Then M is a minimal generating set of \(\mathcal {OP}(n,r)\) if and only if M be the subset having exactly one element from each \({\mathscr {R}}\)-class and from each \({\mathscr {L}}\)-class of \(\mathcal {OP}_n\) of rank r.

Notice that each \({\mathscr {H}}\)-class of \(J^{\mathcal {OP}_n}_r\) that contains an idempotent is a cyclic group of order r (by [2, Corollary 3.6]). Thus, we have the following corollary from Theorem 2.4:

Corollary 2.5

Let M be a minimal generating set of \(\mathcal {OP}(n,r)\). Then the number of distinct sets M is \(r^nn!\).

3 The Nilpotent Rank of \(\mathcal {P}^{\mathcal {OP}_n}_{r}\)

Recall that Zhao [12] showed that both the rank and the idempotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{n-1}\) are equal to \({n\atopwithdelims ()2}\). Recently, Zhao and Fernandes [13] showed that the rank and the idempotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{r}\) are equal to \({n\atopwithdelims ()r}\), for \(2\le r\le n-1\). In this section, we show that the nilpotent rank of the principal factor \(\mathcal {P}^{\mathcal {OP}_n}_{r}\) are also equal to \({n\atopwithdelims ()r}\), for \(2\le r\le n-1\).

Let A be a subset of \(X_n\) of cardinality r and let \(\pi \) be a convex equivalence of weight r on \(X_n\). We may write \(H_{(\pi ,A)}\) for the \({\mathscr {H}}\)-class of \(J^{\mathcal {OP}_n}_{r}\), which is the intersection of \(R_{\pi }=\{\alpha \in J^{\mathcal {OP}_n}_{r}: \mathop {\mathrm {ker}}\nolimits (\alpha )=\pi \}\) and \(L_A=\{\alpha \in J^{\mathcal {OP}_n}_{r}: \mathop {\mathrm {im}}\nolimits (\alpha )=A\}\). The subset A of \(X_n\) of cardinality r is said to be a transversal of the convex equivalence \(\pi \) of weight r on \(X_n\) if each convex equivalence \(\pi \)-class contains exactly one element of A. The following lemma is obvious:

Lemma 3.1

Let \(\alpha \in \mathcal {P}^{\mathcal {OP}_n}_{r}\). Then \(\alpha \) is nilpotent if and only if \(\mathop {\mathrm {im}}\nolimits (\alpha )\) is not a transversal of \(\mathop {\mathrm {ker}}\nolimits (\alpha )\).

Our main result of this section is as follows:

Theorem 3.2

Let \(n\ge 3\) and \(2\le r\le n-1\). Then

$$\begin{aligned}\mathop {\mathrm {nilrank}}\nolimits \mathcal {P}^{\mathcal {OP}_n}_{r}={n\atopwithdelims ()r}. \end{aligned}$$

The proof depends on the following lemma:

Lemma 3.3

Let \(A_1, A_2,\dots , A_m\) (where \(m={n\atopwithdelims ()r}\)) be a list of all the subsets of \(X_n\) with cardinality r. Suppose that there exist distinct convex equivalences \(\pi _1,\pi _2,\dots , \pi _m\) of weight r on \(X_n\) with the property that \(A_i\) is not a transversal of \(\pi _i\), for \(1\le i\le m\). Then there exist nilpotent \(\gamma _i\) in the \({\mathscr {H}}\)-class \(H_{(\pi _i, A_i)}\) (\(1\le i\le m\)) such that the set \(\{ \gamma _1,\gamma _2,\dots , \gamma _m \}\) is a minimal generating set of \(\mathcal {P}^{\mathcal {OP}_n}_{r}\).

Proof

From Lemma 3.1, we know that the \({\mathscr {H}}\)-classes \(H_{(\pi _1, A_1)}, \dots , H_{(\pi _m,A_m)}\) are non-group \({\mathscr {H}}\)-classes, whose elements are nilpotents of \(\mathcal {P}^{\mathcal {OP}_n}_{r}\). Put

$$\begin{aligned} \gamma _i\in H_{(\pi _i,A_i)}, \ {\text{ for }} \ 1\le i\le m. \end{aligned}$$

Then the set \(\{ \gamma _1,\gamma _2,\dots , \gamma _m \}\) is the subset having exactly one element from each \({\mathscr {R}}\)-class and from each \({\mathscr {L}}\)-class of \(\mathcal {OP}_n\) of rank r. It follows immediately from Theorem 2.4 that the set \(\{ \gamma _1,\gamma _2,\dots , \gamma _m \}\) is a minimal generating set of \(\mathcal {P}^{\mathcal {OP}_n}_{r}\). \(\square \)

It remains to prove that the listing of images and convex equivalences postulated in the statement of Lemma 3.3 can actually be carried out. Let \(n\ge 3\) and \(2\le r\le n-1\), and consider the statement:

\(\mathbf{P}(n,r)\): There is a way of listing all the subsets of \(X_n\) of cardinality r as \(A_1,A_2,\dots , A_m\) (with \(m={n\atopwithdelims ()r}\)) so that there exist distinct convex equivalences \(\pi _1, \dots , \pi _m\) of weight r on \(X_n\) with the properties that \(A_i\) is not a transversal of \(\pi _i\) (\(i=1,\dots , m\)).

We shall prove this by a double induction on n and r, the key step being a kind of Pascal’s Triangle implication

$$\begin{aligned} \mathbf{P}(n-1,r-1) \ and \ \mathbf{P}(n-1,r)\Rightarrow \mathbf{P}(n,r). \end{aligned}$$

First, however, we anchor the induction with two lemmas:

Lemma 3.4

\(\mathbf{P}(n,n-1)\) holds for every \(n\ge 3\).

Proof

Consider the list \(A_1, A_2, \dots , A_{n}\) of \(X_n\) of cardinality \(n-1\), where \(A_i=X_n\backslash \{i\}\). For \(1\le i\le n-2\), let \(\pi _i\) be the convex equivalence with a unique non-singleton class \(\{i+1,i+2\}\) and all other classes being singletons. Let \(\pi _{n-1}\) have classes \(\{2\}, \{3\}, \dots , \{n-1\}, \{n,1\}\); and let \(\pi _{n}\) have classes \(\{1,2\}, \{3\}, \dots , \{n-1\}, \{n\}\). It is easy to verify that \(A_1, A_2, \dots , A_{n}\) and \(\pi _1, \pi _2, \dots , \pi _{n}\) have the required property. \(\square \)

Lemma 3.5

\(\mathbf{P}(n,2)\) holds for every \(n\ge 3\).

Proof

We prove this by the induction n. Notice first that P(3, 2) holds by Lemma 3.4. For \(n=4\), we arrange the subsets and convex equivalences as follows:

$$\begin{aligned}&\{1,2\}, \quad \{1,3\}, \quad \{1,4\},\quad \{2,3\},\quad \{2,4\},\quad \{3,4\}\\&12/34, \quad 123/4, \quad 3/412, \quad 23/41, \quad 1/234, \quad 2/341. \end{aligned}$$

Then, it is easy to verify that the subsets and convex equivalences as arranged above satisfy P(4, 2).

Suppose inductively that \(P(n-1,2)\) holds \((n\ge 5)\). Thus, we have a list \(A_1, A_2,\dots , A_t\) (with \(t={n-1\atopwithdelims ()2}\)) of all the subsets of \(X_{n-1}\) of cardinality 2, and a list \(\pi _1,\pi _2,\dots , \pi _t\) of distinct convex equivalences of weight 2 on \(X_{n-1}\) such that \(A_i\) is not a transversal of \(\pi _i\) (\(i=1,2,\dots , t\)). Notice that \(t+n-1={n\atopwithdelims ()2}\). All subsets of \(X_{n}\) of cardinality 2 are

$$\begin{aligned} B_{1},B_{2}, \dots , B_{n-1}; A_1, A_2, \dots , A_t, \end{aligned}$$

where \(B_{j}=\{j,n\}\), \(1\le j\le n-1\). Let \(\pi _i'\) (\(i=1,2, \dots , t\)) be the convex equivalence on \(X_n\) obtained from \(\pi _i\) by adjoining n to the \(\pi _i\)-class containing \(n-1\). Then \(\pi _1', \pi _i', \dots , \pi _t'\) are all distinct, and \(A_i\) is not a transversal of \(\pi _i'\) (\(i=1,2,\dots , t\)).

Let \(\sigma _{1}, \dots , \sigma _{n-1}\) be the list of convex equivalences of weight 2 on \(X_n\), where

$$\begin{aligned}&\sigma _{j} \ \text{ has } \text{ classes } \ \{n,1,\dots ,j\}, \{j+1, \dots , n-1\}, 1\le j\le n-2, \\&\quad \sigma _{n-1} \ \text{ has } \text{ classes } \ \{n\}, \{1, 2, \dots ,n-1\}. \end{aligned}$$

Then \(\sigma _{1}, \dots , \sigma _{n-1}\) are all distinct, \(B_j\) is not a transversal of \(\sigma _j\), for \(1\le j\le n-2\), and each \(\sigma _{j}\) is distinct from every \(\pi '_i\), since \((n-1,n)\in \pi '_i\) and \((n-1,n)\notin \sigma _{j}\). Notice that \(t+n-1={n\atopwithdelims ()2}\). Hence \(\{ \pi _1',\dots ,\pi _t', \sigma _1,\dots , \sigma _{n-1},\}\) is a complete list of all the convex equivalences of weight 2 on \(X_n\). Notice that \(A_1\) is a subset of \(X_{n-1}\) and \((n,n-1)\in \pi _1'\). Then \(A_1\) is not a transversal of \(\sigma _{n-1}\) and \(B_{n-1}=\{n-1,n\}\) is not a transversal of \(\pi _1'\). Arrange the subsets and the convex equivalence as follows:

$$\begin{aligned}&A_{2}, \quad \dots , \quad A_{t}; B_1, \quad B_2, \quad \dots , \quad B_{n-2}, \quad A_{1}, \quad B_{n-1}, \\&\quad \pi _{2}',\quad \dots , \quad \pi _{t}'; \quad \sigma _1,\quad \sigma _2, \quad \dots ,\quad \sigma _{n-2},\quad \sigma _{n-1},\quad \pi _{1}'. \end{aligned}$$

Then, it is easy to verify that the subsets and convex equivalences as arranged above satisfy P(n, 2). \(\square \)

Lemma 3.6

Let \(n\ge 5\) and \(3\le r\le n-2\). Then \(\mathbf{P}(n-1,r-1)\) and \(\mathbf{P}(n-1,r)\) together imply \(\mathbf{P}(n,r)\).

Proof

From the assumption \(P(n-1,r)\), we have a list \(A_1,\dots , A_m\) (where \(m={n-1\atopwithdelims ()r}\)) of the subsets of \(X_{n-1}\) with cardinality r and a list \(\sigma _1,\dots ,\sigma _m\) of the convex equivalences of weight r on \(X_{n-1}\) such that \(A_i\) is not a transversal of \(\sigma _i\), for \(1\le i\le m\).

From the assumption \(P(n-1,r-1)\), we have a list \(B_1,\dots , B_t\) (where \(t={n-1\atopwithdelims ()r-1}\)) of the subsets of \(X_{n-1}\) with cardinality \(r-1\) and a list \(\pi _1,\dots ,\pi _t\) of the convex equivalences of weight \(r-1\) on \(X_{n-1}\) such that \(B_i\) is not a transversal of \(\pi _i\), for \(1\le i\le t\).

Let \(\sigma _i'\) be the convex equivalence obtained from \(\sigma _i\) by adjoining n to the \(\sigma _i\)-class containing \(n-1\), and define \(\pi _i'=\pi _i\cup \{(n,n)\}\). Then \(\sigma _1',\dots , \sigma _m', \pi _1',\dots ,\pi _t'\) are all distinct. Notice that \(m+t={n\atopwithdelims ()r}\). Hence \(\{\sigma _1',\dots , \sigma _m', \pi _1',\dots ,\pi _t'\}\) is a complete list of all the convex equivalences of weight r on \(X_n\). Next we define

$$\begin{aligned} B_i'=B_i\cup \{n\}, \ \text {for} \ 1\le i\le t. \end{aligned}$$

Then \(A_1,\dots , A_m, B_1',B_2',\dots ,B_t'\) are all distinct. Moreover, \(A_i\) is not a transversal of \(\sigma _i'\), for \(1\le i\le m\) and \(B_i'\) is not a transversal of \(\pi _i'\), for \(1\le i\le t\). Arrange the subsets and convex equivalences as follows:

$$\begin{aligned}&A_1,\dots , A_m, \ \ B_1', \ \ B_2',\dots ,B_t' \\&\quad \sigma _1', \ \ \dots , \sigma _m', \ \ \pi _1', \ \ \pi _2', \dots ,\pi _t' \end{aligned}$$

They satisfy all the properties necessary and the inductive step is complete. We have shown that P(nr) holds for all \(n\ge 5\) and \(3\le r\le n-2\). \(\square \)

The pattern of deductions is