1 Introduction

Let n be a natural number. Let \(\Omega _n\) be a finite set with n elements, say \(\Omega _n=\{1,2,\ldots ,n\}\). We denote by \(\mathscr {PT}_n\) the monoid (under composition) of all partial transformations of \(\Omega _n\). The submonoids of \(\mathscr {PT}_n\) of all full transformations and of all partial permutations are denoted by \(\mathscr {T}_n\) and \(\mathscr {I}_n\), respectively. Also, denote by \(\mathscr {S}_n\) the symmetric group on \(\Omega _n\), i.e., the subgroup of \(\mathscr {PT}_n\) of all permutations of \(\Omega _n\). For \(s\in \mathscr {PT}_n\), we denote the domain of s by \(\mathop {\mathrm {Dom}}\nolimits (s)\), the image of s by \(\mathop {\mathrm {Im}}\nolimits (s)\), the kernel of s by \(\ker (s)\), and the set of fix points of s by \(\mathop {\mathrm {Fix}}\nolimits (s)\), i.e., \(\mathop {\mathrm {Fix}}\nolimits (s)=\{x\in \mathop {\mathrm {Dom}}\nolimits (s)\mid xs=x\}\).

Let us consider \(\Omega _n\) endowed with the usual (linear) order. An element \(s\in \mathscr {PT}_n\) is said to be order-preserving if \(x\le y\) implies \(xs\le ys\), for all \(x,y \in \mathop {\mathrm {Dom}}\nolimits (s)\). Denote by \(\mathscr {PO}_n\) the submonoid of \(\mathscr {PT}_n\) of all order-preserving partial transformations. As usual, we denote by \(\mathscr {O}_n\) the monoid \(\mathscr {PO}_n\cap \mathscr {T}_n\) of all full transformations that preserve the order and by \(\mathscr {POI}_n\) its injective counterpart, i.e., the inverse monoid \(\mathscr {PO}_n\cap \mathscr {I}_n\) of all order-preserving partial permutations of \(\Omega _n\).

Semigroups of order-preserving transformations have been considered in the literature for over than fifty years. Starting in 1962, Aǐzenštat [1, 2] gave a presentation for \(\mathscr {O}_n\), from which it can be deduced that, for \(n>1\), \(\mathscr {O}_n\) only has one non-trivial automorphism, and characterized the congruences of \(\mathscr {O}_n\). Also in 1962, Popova [26] exhibited a presentation for \(\mathscr {PO}_n\). In 1971, Howie [20] calculated the cardinal and the number of idempotents of \(\mathscr {O}_n\) and later (1992), jointly with Gomes [18], determined the ranks and the idempotent ranks of \(\mathscr {O}_n\) and \(\mathscr {PO}_n\). More recently, Laradji and Umar [22, 23] presented more combinatorial properties of these two monoids. Certain classes of divisors of the monoid \(\mathscr {O}_n\) were determined by Higgins [19] and by Vernitskii and Volkov [30] in 1995, by Fernandes [8] in 1997 and by Fernandes and Volkov [15] in 2010. On the other hand, the monoid \(\mathscr {POI}_n\) has been object of study by the first author in several papers [8,9,10,11,12], by Derech [6], by Garba [17], by Cowan and Reilly [4], by Delgado and Fernandes [5], by Ganyushkin and Mazorchuk [16], by Dimitrova and Koppitz [7], among other authors and papers.

For general background on semigroups, we refer the reader to Howie’s book [21].

Let \(n\ge 2\).

Let \(S\in \{\mathscr {O}_n,\mathscr {POI}_n,\mathscr {PO}_n\}\). We have the following descriptions of the Green relations in the semigroups S:

  • \(s\mathcal {L}t\) if and only if \(\mathop {\mathrm {Im}}\nolimits (s)=\mathop {\mathrm {Im}}\nolimits (t)\),

  • \(s\mathcal {R}t\) if and only if \(\ker (s)=\ker (t)\),

  • \(s\mathcal {J}t\) if and only if \(|\mathop {\mathrm {Im}}\nolimits (s)|=|\mathop {\mathrm {Im}}\nolimits (t)|\), and

  • \(s\mathcal {H}t\) if and only if \(s=t\),

for all \(s,t\in S\). If \(S=\mathscr {POI}_n\), for the Green relation \(\mathcal {R}\), we have, even more simply,

  • \(s\mathcal {R}t\) if and only if \(\mathop {\mathrm {Dom}}\nolimits (s)=\mathop {\mathrm {Dom}}\nolimits (t)\),

for all \(s,t\in S\). Let

$$\begin{aligned} J_k=J_k^S=\{s\in S\mid |\mathop {\mathrm {Im}}\nolimits (s)|=k\}\quad \;\text {and}\;\quad I_k=I_k^S=\{s\in S\mid |\mathop {\mathrm {Im}}\nolimits (s)|\le k\}, \end{aligned}$$

for \(0\le k\le n\). If \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\) then

$$\begin{aligned} S/\mathcal {J}=\{J_0<_\mathcal {J}J_1<_\mathcal {J}\cdots <_\mathcal {J}J_n\} \end{aligned}$$

and \(\{\emptyset \}=I_0\subset I_2\subset \cdots \subset I_n=S\) are all the ideals of S. On the other hand, if \(S=\mathscr {O}_n\) then

$$\begin{aligned} S/\mathcal {J}=\{J_1<_\mathcal {J}J_2<_\mathcal {J}\cdots <_\mathcal {J}J_n\} \end{aligned}$$

and \(I_1\subset I_2\subset \cdots \subset I_n=S\) are all the ideals of S. See [10, 18].

Recall that a Rees congruence\(\rho \) of a semigroup S is a congruence associated to an ideal I of S: \(s\rho t\) if and only if \(s=t\) or \(s,t\in I\), for all \(s,t\in S\).

Observe that Aǐzenštat [2] proved the congruences of \(\mathscr {O}_n\) are exactly the identity and its n Rees congruences. See [24] for another proof. Analogously, the congruences of \(\mathscr {POI}_n\) and \(\mathscr {PO}_n\) are exactly their \(n+1\) Rees congruences. This has been shown, for \(\mathscr {POI}_n\), by Derech [6] and, independently, by Fernandes [10] and, for \(\mathscr {PO}_n\), by Fernandes et al. [13].

Let S be a monoid such that \(S{\setminus }\{1\}\) is an ideal of S. Let e and f be two idempotents of S such that \(ef=fe=f\). Then, clearly, the mapping \(\phi :S\longrightarrow S\) defined by \(1\phi =e\) and \(s\phi =f\), for all \(s\in S{\setminus }\{1\}\), is an endomorphism (of semigroups) of S. This applies to any \(S\in \{\mathscr {O}_n,\mathscr {POI}_n,\mathscr {PO}_n\}\).

Let M be a monoid, let S be a subsemigroup of M and let g be a unit of M such that \(g^{-1}Sg=S\). Then, it is easy to check that the mapping \(\phi ^g:S\longrightarrow S\) defined by \(s\phi ^g=g^{-1}sg\), for all \(s\in S\), is an automorphism of S.

Consider the following permutation of \(\Omega _n\):

$$\begin{aligned} \sigma =\left( \begin{array}{lllll} 1&{}\quad 2&{}\quad \cdots &{}\quad n-1&{}\quad n\\ n&{}\quad n-1&{}\quad \cdots &{}\quad 2&{}\quad 1 \end{array}\right) . \end{aligned}$$

Let \(S\in \{\mathscr {O}_n,\mathscr {POI}_n,\mathscr {PO}_n\}\). It is easy to verify that \(\sigma ^{-1}S\sigma =S\). Therefore, the permutation \(\sigma \) induces a non-trivial automorphism \(\phi ^{\sigma }\) of S. In fact, this is the unique non-trivial automorphism of S. See [3, Corollary 5.2].

Finding the automorphisms and endomorphisms of transformation semigroups is a classical problem. They have been determined for several transformation semigroups, for instance, \(\mathscr {T}_n\) [28], \(\mathscr {I}_n\) [27], and the Brauer-type semigroups [25]. Furthermore, regarding semigroups of order-preserving transformations, Fernandes et al. [14] proved the following description of the endomorphisms of \(\mathscr {O}_n\):

Theorem 1.1

[14, Theorem 1.1] Let \(\phi : \mathscr {O}_n\rightarrow \mathscr {O}_n\) be any mapping. Then \(\phi \) is an endomorphism of the semigroup \(\mathscr {O}_n\) if and only if one of the following holds:

  1. (a)

    \(\phi \) is an automorphism and so \(\phi \) is the identity or \(\phi =\phi ^{\sigma }\);

  2. (b)

    there exist idempotents \(e,f\in \mathscr {O}_n\) with \(e\not =f\) and \(ef=fe=f\) such that \(1\phi =e\) and \((\mathscr {O}_n{\setminus }\{1\})\phi =\{f\}\);

  3. (c)

    \(\phi \) is a constant mapping with idempotent value.

And, as a corollary:

Theorem 1.2

[14, Theorem 1.2] The semigroup \(\mathscr {O}_n\) has \(2+\sum _{i=0}^{n-1}\left( {\begin{array}{c}n+i\\ 2i+1\end{array}}\right) F_{2i+2}\) endomorphisms, where \(F_{2i+2}\) denotes the \((2i+2)\)th Fibonacci number.

In this paper we describe the monoids of the endomorphisms of the semigroups \(\mathscr {POI}_n\) and \(\mathscr {PO}_n\). As an application of these descriptions, we compute the number of such endomorphisms. This paper is organized as follows. After the current section, we give a miscellaneous of auxiliary results in Sect. 2. Finally, in Sect. 3 we present and prove our main results.

2 Preliminary results

In this section we present a series of auxiliary results. We also construct a certain type of endomorphisms of \(\mathscr {POI}_n\) and \(\mathscr {PO}_n\).

Our first two lemmas have some general nature.

Lemma 2.1

Let S be a regular semigroup. Let I be an ideal of S and \(\phi \) be an endomorphism of S such that the kernel of \(\phi \) is the Rees congruence associated to I. Let \(s,t\in S{\setminus }I\). Then:

  1. 1.

    \(s\mathcal {L}t\) if and only if \(s\phi \mathcal {L}t\phi \);

  2. 2.

    \(s\mathcal {R}t\) if and only if \(s\phi \mathcal {R}t\phi \).

Proof

We prove the lemma for the Green relation \(\mathcal {L}\). The proof for \(\mathcal {R}\) is similar.

First, notice that \(I\phi \phi ^{-1}=I\) and the restriction of \(\phi \) to \(S{\setminus } I\) is injective.

Let \(s,t\in S{\setminus } I\). If \(s\mathcal {L}t\), then \(s\phi \mathcal {L}t\phi \), since any homomorphism of semigroups preserve Green relations (i.e., images of related elements are related). Conversely, suppose that \(s\phi \mathcal {L}t\phi \). As S is regular, then \(S\phi \) is also regular, whence \(s\phi \) and \(t\phi \) are also \(\mathcal {L}\)-related in \(S\phi \) and so, for some \(u,v\in S\), we have \(s\phi =(u\phi )(t\phi )\) and \(t\phi =(v\phi )(s\phi )\). Thus, \(s\phi =(ut)\phi \) and so \(ut\in S{\setminus } I\), since \(s\in S{\setminus } I\) and \(I\phi \phi ^{-1}=I\). Moreover, as \(\phi \) is injective in \(S{\setminus } I\), it follows that \(s=ut\). Analogously, \(t=vs\) and so \(s\mathcal {L}t\), as required. \(\square \)

Lemma 2.2

Let S be any subsemigroup of \(\mathscr {PT}_n\) which contains transformations with arbitrary images of size less than n. Let \(s\in S\) and \(k\in \mathbb N\) be such that \(1\le \mathop {\mathrm {rank}}\nolimits (s)\le k\le n-2\). Then, there exists \(t\in S\) such that \(\mathop {\mathrm {rank}}\nolimits (t)=k+1\) and \(st\ne s\).

Proof

Let A be any subset of \(\Omega _n\) such that \(\mathop {\mathrm {Im}}\nolimits (s)\subset A\) and \(|A|=k+2\). Let t be any element of S such that \(\mathop {\mathrm {Im}}\nolimits (t)=A{\setminus }\{\min (\mathop {\mathrm {Im}}\nolimits (s))\}\) (notice that \(\mathop {\mathrm {Im}}\nolimits (s)\ne \emptyset \)). Then, \(\mathop {\mathrm {rank}}\nolimits (t)=k+1\) and \(\mathop {\mathrm {Im}}\nolimits (st)\subseteq \mathop {\mathrm {Im}}\nolimits (t)\). Since \(\min (\mathop {\mathrm {Im}}\nolimits (s))\not \in \mathop {\mathrm {Im}}\nolimits (t)\), then also \(\min (\mathop {\mathrm {Im}}\nolimits (s))\not \in \mathop {\mathrm {Im}}\nolimits (st)\), whence \(\mathop {\mathrm {Im}}\nolimits (s)\ne \mathop {\mathrm {Im}}\nolimits (st)\), and so \(s\ne st\), as required. \(\square \)

Observe that the previous lemma applies to any \(S\in \{\mathscr {O}_n,\mathscr {POI}_n,\mathscr {PO}_n\}\).

Next, we give two particular properties of our semigroups.

Lemma 2.3

Let \(k\in \{0,1,\ldots ,n-2\}\). Then, there exist idempotents \(h_1,\ldots ,h_n\in J_{k+1}^{\mathscr {POI}_n}\) such that \(h_ih_j\in I_k^{\mathscr {POI}_n}\), for all \(1\le i<j\le n\).

Proof

We have \(\left( {\begin{array}{c}n\\ k+1\end{array}}\right) \) distinct subsets of \(\Omega _n\) with \(k+1\) elements. As \(0\le k\le n-2\), then \(1\le k+1\le n-1\), and so \(\left( {\begin{array}{c}n\\ k+1\end{array}}\right) \ge n\). Therefore, we may take (at least) n distinct subsets \(Y_1,\ldots , Y_n\) of \(\Omega _n\) with \(k+1\) elements. Let \(h_i\) be the partial identity on \(Y_i\), for \(1\le i\le n\). Then, for \(1\le i<j\le n\), the transformation \(h_ih_j\) is the partial identity on \(Y_i\cap Y_j\) and, since \(Y_i\ne Y_j\), we obtain \(|Y_i\cap Y_j|<|Y_j|=k+1\), whence \(h_ih_j\in I_k^{\mathscr {POI}_n}\), as required. \(\square \)

For \(k\in \{2,3,\ldots ,n\}\), consider the following two transformations of \(\mathscr {O}_n\) with rank \(n-1\):

$$\begin{aligned} f_k= & {} \left( \begin{array}{lllllll} 1&{}\quad \cdots &{}\quad k-1&{}\quad k&{}\quad k+1&{}\quad \cdots &{}\quad n\\ 1&{}\quad \cdots &{}\quad k-1&{}\quad k-1&{}\quad k+1&{}\quad \cdots &{}\quad n \end{array}\right) \quad \text {and}\quad \\ g_k= & {} \left( \begin{array}{lllllll} 1&{}\quad \cdots &{}\quad k-2&{}\quad k-1&{}\quad k&{}\quad \cdots &{}\quad n\\ 1&{}\quad \cdots &{}\quad k-2&{}\quad k&{}\quad k&{}\quad \cdots &{}\quad n \end{array}\right) . \end{aligned}$$

We have:

Lemma 2.4

Each of the \(n-1\)\(\mathcal {R}\)-classes of transformations of rank \(n-1\) of \(\mathscr {O}_n\) has exactly two idempotents, namely \(f_k\) and \(g_k\), for some \(k\in \{2,3,\ldots ,n\}\).

Proof

Let R be an \(\mathcal {R}\)-class of \(\mathscr {O}_n\) contained in \(J_{n-1}^{\mathscr {O}_n}\). Then all elements of R have the same kernel, which is associated to a partition of \(\Omega _n\) of the form \(\{\{i\}\mid i\in \{1,2,\ldots ,k-2,k+1,\ldots ,n\}\}\cup \{\{k-1,k\}\}\), for some \(k\in \{2,3,\ldots ,n\}\). Clearly, \(f_k\) and \(g_k\) are two distinct idempotents of R. Moreover, let

$$\begin{aligned} e=\left( \begin{array}{lllllll} 1&{}\quad \cdots &{}\quad k-1&{}\quad k&{}\quad k+1&{}\quad \cdots &{}\quad n\\ i_1&{}\quad \cdots &{}\quad i_{k-1}&{}\quad i_{k-1}&{}\quad i_k&{}\quad \cdots &{}\quad i_{n-1} \end{array}\right) , \end{aligned}$$

with \(1\le i_1< \cdots <i_{n-1}\le n\), be an (arbitrary) idempotent of R. Then \(\mathop {\mathrm {Fix}}\nolimits (e)=\mathop {\mathrm {Im}}\nolimits (e)\) and so \(|\mathop {\mathrm {Fix}}\nolimits (e)|=n-1\), whence there exists a unique \(i\in \Omega _n\) such that \((i)e\ne i\). Since \(((i)e)e=(i)e\), we have \(((i)e,i)\in \ker (e)\), and so \(\{(i)e,i\}=\{k-1,k\}\). If \(i=k\) then, clearly, \(e=f_k\). Otherwise, \(i=k-1\) and then, clearly, \(e=g_k\). This proves the lemma. \(\square \)

For \(i\in \{1,2,\ldots ,n\}\), let

$$\begin{aligned} e_i=\left( \begin{array}{llllll} 1&{}\quad \cdots &{}\quad i-1&{}\quad i+1&{}\quad \cdots &{}\quad n\\ 1&{}\quad \cdots &{}\quad i-1&{}\quad i+1&{}\quad \cdots &{}\quad n \end{array}\right) . \end{aligned}$$

It is clear that each of the n\(\mathcal {R}\)-classes of transformations of rank \(n-1\) of \(\mathscr {POI}_n\) has exactly one idempotent (notice that \(\mathscr {POI}_n\) is an inverse semigroup), namely \(e_i\), for some \(i\in \{1,2,\ldots ,n\}\).

Notice that \(J_{n-1}^{\mathscr {PO}_n}=J_{n-1}^{\mathscr {O}_n}\cup J_{n-1}^{\mathscr {POI}_n}\) (a disjoint union). Hence, by the above observations, \(J_{n-1}^{\mathscr {PO}_n}\) contains n\(\mathcal {R}\)-classes exactly with one idempotent and \(n-1\)\(\mathcal {R}\)-classes exactly with two idempotents. Moreover, each \(\mathcal {R}\)-class contained in \(J_{n-1}^{\mathscr {PO}_n}\) has precisely n transformations (one for each possible image with \(n-1\) elements). Additionally, it is clear that each \(\mathcal {R}\)-class contained in \(J_1^{\mathscr {PO}_n}\) is determined by the domain of the transformations (as it happens in general in \(\mathscr {POI}_n\)), and so we have \(2^n-1\)\(\mathcal {R}\)-classes of transformations of rank 1 of \(\mathscr {PO}_n\), each with exactly as many idempotents as elements in the corresponding domain. Such as for \(J_{n-1}^{\mathscr {PO}_n}\), each \(\mathcal {R}\)-class contained in \(J_1^{\mathscr {PO}_n}\) has precisely n transformations (one for each possible image with 1 element).

Next, we aim to construct an endomorphism of \(\mathscr {PO}_n\).

Let us define a mapping \(\phi _1:\mathscr {PO}_n\longrightarrow \mathscr {PO}_n\) by:

  1. 1.

    \(1\phi _1=1\);

  2. 2.

    For \(s\in J_{n-1}^{\mathscr {POI}_n}\), let \(s\phi _1=\left( {\begin{array}{c}i\\ j\end{array}}\right) \), where \(i,j\in \{1,2,\ldots ,n\}\) are the unique indices such that \(e_i\mathcal {R}s\mathcal {L}e_j\);

  3. 3.

    For \(s\in J_{n-1}^{\mathscr {O}_n}\), let \(s\phi _1=\left( \begin{array}{ll} k-1&{}k\\ k_s&{}k_s \end{array}\right) \), where \(\{k_s\}=\Omega _n{\setminus }\mathop {\mathrm {Im}}\nolimits (s)\) and \(k\in \{2,3,\ldots ,n\}\) is the unique index such that \(s\mathcal {R}f_k\) (and \(s\mathcal {R}g_k\));

  4. 4.

    \(I_{n-2}^{\mathscr {PO}_n}\phi _1=\{\emptyset \}\).

Clearly, \(\phi _1\) is well defined mapping. Moreover, since \(\mathscr {PO}_n\) (such as \(\mathscr {POI}_n\)) is an \(\mathcal {H}\)-trivial semigroup (and so each element of \(\mathscr {PO}_n\) is perfectly defined by its \(\mathcal {L}\)-class and \(\mathcal {R}\)-class, i.e., its image and kernel), the restriction of \(\phi _1\) to \(\mathscr {PO}_n{\setminus } I_{n-2}^{\mathscr {PO}_n}\) is injective. Furthermore, it is a routine matter to prove the following lemma.

Lemma 2.5

The mapping \(\phi _1\) is an endomorphism of \(\mathscr {PO}_n\) such that \(\mathscr {POI}_n\phi _1\subseteq \mathscr {POI}_n\). Consequently, the restriction of \(\phi _1\) to \(\mathscr {POI}_n\) may also be seen as an endomorphism of \(\mathscr {POI}_n\).

All endomorphisms similar to \(\phi _1\) have the following property.

Lemma 2.6

Let \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\). Let \(\phi \) be an endomorphism of S such that \(1\phi =1\), \(J_{n-1}\phi \subseteq J_1\) and \(I_{n-2}\phi =\{\emptyset \}\). Then \(\phi \) is perfectly defined by the images of the idempotents \(e_1,\ldots , e_n\). Moreover, \(|\mathop {\mathrm {Dom}}\nolimits (s\phi )|=1\), for all \(s\in J_{n-1}^{\mathscr {POI}_n}\), and \(|\mathop {\mathrm {Dom}}\nolimits (s\phi )|=2\), for all \(s\in J_{n-1}^{\mathscr {O}_n}\).

Proof

We begin by observing that the kernel of \(\phi \) must be the Rees congruence associated to \(I_{n-2}\) and so \(\phi \) is injective in \(S{\setminus } I_{n-2}\). Next, as \(\phi \) applies \(\mathcal {R}\)-related transformations of \(J_{n-1}\) in \(\mathcal {R}\)-related transformations of \(J_1\), \(\phi \) injective in \(J_{n-1}\) and the \(\mathcal {R}\)-classes contained in \(J_{n-1}\) and in \(J_1\) have the same number of elements, namely n, the endomorphism \(\phi \) applies each \(\mathcal {R}\)-class of \(J_{n-1}\) bijectively in a \(\mathcal {R}\)-class of \(J_1\). It follows that, for each \(\mathcal {R}\)-class R of \(J_{n-1}\), the number of idempotents of R and of the \(\mathcal {R}\)-class \(R\phi \) of \(J_1\) must be the same. Thus, in particular, \(J_{n-1}^{\mathscr {POI}_n}\phi =J_1^{\mathscr {POI}_n}\).

For each \(i\in \{1,2,\ldots ,n\}\), let \(k_i\in \Omega _n\) be such that \(e_i\phi =\left( {\begin{array}{c}k_i\\ k_i\end{array}}\right) \). Notice that \(\left( \begin{array}{llll} 1&{}2&{}\cdots &{}n\\ k_1&{}k_2&{}\cdots &{}k_n \end{array}\right) \) is a permutation of \(\Omega _n\).

Let \(s\in J_{n-1}^{\mathscr {POI}_n}\). Then, there exists a unique pair \((i,j)\in \{1,2,\ldots ,n\}\times \{1,2,\ldots ,n\}\) such that \(s\mathcal {R}e_i\) and \(s\mathcal {L}e_j\). Then, \(s\phi \mathcal {R}e_i\phi \) and \(s\phi \mathcal {L}e_j\phi \), and so \(s\phi =\left( {\begin{array}{c}k_i\\ k_j\end{array}}\right) \).

If \(S=\mathscr {POI}_n\) then the proof is completed. Thus, from now on, we suppose that \(S=\mathscr {PO}_n\).

Let \(i\in \{2,3,\ldots ,n\}\). Then \(f_i\mathcal {L}e_i\) and \(g_i\mathcal {L}e_{i-1}\), whence \(f_i\phi \mathcal {L}e_i\phi \) and \(g_i\phi \mathcal {L}e_{i-1}\phi \), and so \(\mathop {\mathrm {Im}}\nolimits (f_i\phi )=\{k_i\}\) and \(\mathop {\mathrm {Im}}\nolimits (g_i\phi )=\{k_{i-1}\}\). Hence, \(k_i\in \mathop {\mathrm {Im}}\nolimits (f_i\phi )=\mathop {\mathrm {Fix}}\nolimits (f_i\phi )\subseteq \mathop {\mathrm {Dom}}\nolimits (f_i\phi )\) and \(k_{i-1}\in \mathop {\mathrm {Im}}\nolimits (g_i\phi )=\mathop {\mathrm {Fix}}\nolimits (g_i\phi )\subseteq \mathop {\mathrm {Dom}}\nolimits (g_i\phi )\). Besides, as \(f_i\mathcal {R}g_i\), we have \(f_i\phi \mathcal {R}g_i\phi \) and so \(\mathop {\mathrm {Dom}}\nolimits (f_i\phi )=\mathop {\mathrm {Dom}}\nolimits (g_i\phi )\). Since \(|\mathop {\mathrm {Dom}}\nolimits (f_i\phi )|=2\) and \(k_{i-1}\ne k_i\), it follows that \(\mathop {\mathrm {Dom}}\nolimits (f_i\phi )=\{k_{i-1},k_i\}\).

Now, let \(s\in J_{n-1}^{\mathscr {O}_n}\). Then, there exists a unique pair \((i,j)\in \{2,3,\ldots ,n\}\times \{1,2,\ldots ,n\}\) such that \(s\mathcal {R}f_i\) and \(s\mathcal {L}e_j\). Thus, \(s\phi \mathcal {R}f_i\phi \) and \(s\phi \mathcal {L}e_j\phi \), and so \(s\phi =\left( \begin{array}{ll} k_{i-1}&{}k_i\\ k_j&{}k_j \end{array}\right) \), which finishes the proof of this lemma. \(\square \)

Observe that we may also conclude that, within the conditions of the previous lemma, we have at most n! endomorphisms.

Now, we recall that a subsemigroup T of \(\mathscr {PT}_n\) is said to be \(\mathscr {S}_n\)-normal if \(g^{-1}Tg\subseteq T\), for all \(g\in \mathscr {S}_n\). In 1975, Sullivan [29, Theorem 2] proved that \(\text {Aut}(T)\simeq \mathscr {S}_n\), for any \(\mathscr {S}_n\)-normal subsemigroup T of \(\mathscr {PT}_n\) containing a constant mapping. Moreover, we obtain an isomorphism \(\Theta :\mathscr {S}_n\longrightarrow \text {Aut}(T)\) by defining \(g\Theta =\theta _g\), where \(\theta _g\) denotes the inner automorphism of T associated to g (i.e., \(t\theta _g=g^{-1}tg\), for all \(t\in T\)), for all \(g\in \mathscr {S}_n\).

Let \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\). Let \(I_1^1=I_1\cup \{1\}\). It is clear that \(I_1^1\) is an \(\mathscr {S}_n\)-normal subsemigroup of \(\mathscr {PT}_n\) containing a constant mapping. Therefore, by Sullivan’s Theorem, we have

$$\begin{aligned} \text {Aut}(I_1^1)=\{\theta _g\mid g\in \mathscr {S}_n\}\simeq \mathscr {S}_n, \end{aligned}$$

where \(\theta _g\) denotes the inner automorphism of \(I_1^1\) associated to g, for all \(g\in \mathscr {S}_n\).

Let \(\phi _g={\phi _1}_{\displaystyle |_S}\theta _g\) (where \({\phi _1}_{\displaystyle |_S}\) denotes the restriction of \(\phi _1\) to S), considered as a mapping from S to S, for all \(g\in \mathscr {S}_n\). Clearly, \(\{\phi _g\mid g\in \mathscr {S}_n\}\) is a set of n! distinct endomorphisms of S. Moreover, it is easy to conclude the following result, with which we finish this section.

Lemma 2.7

Let \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\). Let \(\phi \) be an endomorphism of S such that \(1\phi =1\), \(J_{n-1}\phi \subseteq J_1\) and \(I_{n-2}\phi =\{\emptyset \}\). Then, \(\phi =\phi _g\), for some \(g\in \mathscr {S}_n\).

3 Main results

Let \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\). Let \(\phi \) be an endomorphism of the semigroup S. Then, \(\ker (\phi )\) is the Rees congruence of S associated to \(I_k\), for some \(k\in \{0,1,\ldots ,n\}\). Observe that the restriction of \(\phi \) to \(S{\setminus } I_k\) is an injective mapping and \(|I_k\phi |=1\). Let \(f\in S\) be such that \(I_k\phi =\{f\}\). Clearly, f is an idempotent of S. Notice also that \(\{f\}\phi ^{-1}=I_k\).

Suppose that \(\phi \) is neither a constant mapping nor an automorphism. Then, since \(\ker (\phi )\) is not trivial and not universal, we have \(1\le k\le n-1\).

Let us admit that \(k=n-1\) and take \(e=1\phi \). Then e is also an idempotent, \(e\ne f\) and, for \(s\in I_{n-1}\), we have \(ef=(1\phi )(s\phi )=(1\cdot s)\phi =s\phi =f=s\phi =(s\cdot 1)\phi =(s\phi )(1\phi )=fe\).

From now on, suppose that \(1\le k\le n-2\). Since any homomorphism of semigroups preserve Green relations, there exists \(\ell \in \{0,1,\ldots ,n\}\) such that \(J_{k+1}\phi \subseteq J_{\ell }\). Under these conditions, we prove two lemmas.

Lemma 3.1

Under the above conditions, one has \(\mathop {\mathrm {rank}}\nolimits (f)<\ell \).

Proof

First, notice that \((\cup _{i=k+1}^{n}J_i)\phi \subseteq \cup _{i=\ell }^{n}J_i\), since \(J_{k+1}\phi \subseteq J_{\ell }\) and any homomorphism preserves the quasi-order \(\le _\mathcal {J}\).

Next, we show that \(\mathop {\mathrm {rank}}\nolimits (f)\le \ell \). Take \(s\in J_k\) and \(t\in J_{k+1}\). Then, \(s<_\mathcal {J}t\), and so \(s\phi \le _\mathcal {J}t\phi \). Since \(J_k\phi =\{f\}\) and \(J_{k+1}\phi \subseteq J_{\ell }\), we have \(s\phi =f\) and \(t\phi \in J_{\ell }\), whence \(\mathop {\mathrm {rank}}\nolimits (f)\le \ell \).

Now, we prove that \(\mathop {\mathrm {rank}}\nolimits (f)\le k\). By contradiction, suppose that \(\mathop {\mathrm {rank}}\nolimits (f)\ge k+1\). Then, \(I_k\subseteq I_{\mathop {\mathrm {rank}}\nolimits (f)-1}\), and so

$$\begin{aligned} (S{\setminus } I_k)\phi =(\cup _{i=k+1}^{n}J_i)\phi \subseteq \cup _{i=\ell }^{n}J_i \subseteq \cup _{i=\mathop {\mathrm {rank}}\nolimits (f)}^{n}J_i =S{\setminus } I_{\mathop {\mathrm {rank}}\nolimits (f)-1}\subseteq S{\setminus } I_k. \end{aligned}$$

It follows that \((S{\setminus } I_k)\phi =S{\setminus } I_k\), since the restriction of \(\phi \) to \(S{\setminus } I_k\) is injective. As \(\mathop {\mathrm {rank}}\nolimits (f)\ge k+1\) also implies that \(f\in S{\setminus } I_k\), we deduce that \(I_k\cap S{\setminus } I_k=\{f\}\phi ^{-1}\cap S{\setminus } I_k\ne \emptyset \), which is a contradiction. Thus, we must have \(\mathop {\mathrm {rank}}\nolimits (f)\le k\).

Finally, we prove that \(\mathop {\mathrm {rank}}\nolimits (f)<\ell \). Suppose, by contradiction, that \(\mathop {\mathrm {rank}}\nolimits (f)=\ell \). Take \(s\in J_{k+1}\). Since \(J_{k+1}\phi \subseteq J_{\ell }\), we have \(s\phi \mathcal {J}f\). Moreover, as \(f\in I_k\), we also have \(sf,fs\in I_k\), whence \((sf)\phi =(fs)\phi =f\phi =f\). From \(f=(fs)\phi =f\phi s\phi =f(s\phi )\), it follows that \(\mathop {\mathrm {Im}}\nolimits (f)=\mathop {\mathrm {Im}}\nolimits (f(s\phi ))\subseteq \mathop {\mathrm {Im}}\nolimits (s\phi )\). From \(f=(sf)\phi =s\phi f\phi =(s\phi )f\), it follows that \(\ker (f)=\ker ((s\phi )f)\supseteq \ker (s\phi )\). Since \(s\phi \mathcal {J}f\), we have \(|\ker (f)|=|\mathop {\mathrm {Im}}\nolimits (f)|=|\mathop {\mathrm {Im}}\nolimits (s\phi )|=|\ker (s\phi )|\), and so \(\mathop {\mathrm {Im}}\nolimits (f)=\mathop {\mathrm {Im}}\nolimits (s\phi )\) and \(\ker (f)=\ker (s\phi )\), whence \(f\mathcal {L}s\phi \) and \(f\mathcal {R}s\phi \), i.e., \(f\mathcal {H}s\phi \). Thus, \(f=s\phi \), which is a contradiction (being the case that \(s\in J_{k+1}\) and \(\{f\}\phi ^{-1}=I_k\)). Therefore, \(\mathop {\mathrm {rank}}\nolimits (f)<\ell \), as required. \(\square \)

Lemma 3.2

Under the above conditions, \(1\phi =1\), \(J_{n-1}\phi \subseteq J_1\), and \(I_{n-2}\phi =\{\emptyset \}\).

Proof

Since \(k\in \{1,2,\ldots ,n-2\}\), by Lemma 2.3, we may take idempotents \(h_1,\ldots ,h_n\in J_{k+1}^{\mathscr {POI}_n}\) such that \(h_ih_j\in I_k^{\mathscr {POI}_n}\), for all \(1\le i<j\le n\).

Recall that \(f\in I_k\), and so \(f\phi =f\). Let \(i\in \{1,2,\ldots ,n\}\). Then, \(f (h_i\phi )=f\phi h_i\phi =(fh_i)\phi =f\), since \(fh_i\in I_k\). Hence, \(\mathop {\mathrm {Im}}\nolimits (f)\subseteq \mathop {\mathrm {Im}}\nolimits (h_i\phi )\).

Next, let \(1\le i<j\le n\). By the previous paragraph, we have \(\mathop {\mathrm {Im}}\nolimits (f)\subseteq \mathop {\mathrm {Im}}\nolimits (h_i\phi )\cap \mathop {\mathrm {Im}}\nolimits (h_j\phi )\). Conversely, let \(a\in \mathop {\mathrm {Im}}\nolimits (h_i\phi )\cap \mathop {\mathrm {Im}}\nolimits (h_j\phi )\). Since \(h_i\phi \) and \(h_j\phi \) are idempotents and \(h_ih_j\in I_k\), we have \(\mathop {\mathrm {Im}}\nolimits (h_i\phi )=\mathop {\mathrm {Fix}}\nolimits (h_i\phi )\), \(\mathop {\mathrm {Im}}\nolimits (h_j\phi )=\mathop {\mathrm {Fix}}\nolimits (h_j\phi )\) and \((h_ih_j)\phi =f\), whence \(af=a(h_ih_j)\phi =(a(h_i\phi ))(h_j\phi )=a(h_j\phi )=a\), and so \(a\in \mathop {\mathrm {Im}}\nolimits (f)\). Then, \(\mathop {\mathrm {Im}}\nolimits (h_i\phi )\cap \mathop {\mathrm {Im}}\nolimits (h_j\phi )\subseteq \mathop {\mathrm {Im}}\nolimits (f)\), and thus \(\mathop {\mathrm {Im}}\nolimits (h_i\phi )\cap \mathop {\mathrm {Im}}\nolimits (h_j\phi )=\mathop {\mathrm {Im}}\nolimits (f)\).

Let \(E_i=\mathop {\mathrm {Im}}\nolimits (h_i\phi ){\setminus }\mathop {\mathrm {Im}}\nolimits (f)\), for \(1\le i\le n\). Clearly, \( \mathop {\mathrm {Im}}\nolimits (f)\cap (\cup _{i=1}^{n}E_i)=\emptyset \) and, for \(1\le i<j\le n\), the equality \(\mathop {\mathrm {Im}}\nolimits (h_i\phi )\cap \mathop {\mathrm {Im}}\nolimits (h_j\phi )=\mathop {\mathrm {Im}}\nolimits (f)\) implies that \(E_i\cap E_j=\emptyset \). Moreover, since \(h_i\in J_{k+1}\), then \(\mathop {\mathrm {rank}}\nolimits (h_i\phi )=\ell >\mathop {\mathrm {rank}}\nolimits (f)\), by Lemma 3.1, and so \(|E_i|\ge 1\), for \(1\le i\le n\). Thus \(|E_1|=|E_2|=\cdots =|E_n|=1\) and \(\cup _{i=1}^{n}E_i=\Omega _n\), from which follows that \(\mathop {\mathrm {Im}}\nolimits (f)=\emptyset \).

Now, observe that \(\ell =\mathop {\mathrm {rank}}\nolimits (h_1\phi )=|\mathop {\mathrm {Im}}\nolimits (h_1\phi )|=|E_1\cup \mathop {\mathrm {Im}}\nolimits (f)|=|E_1|=1\). Then, \(J_{k+1}\phi \subseteq J_1\), and so we cannot have more than n elements of \(J_{k+1}\phi \) in distinct \(\mathcal {L}\)-classes. Thus, by Lemma 2.1, we cannot have more than n elements of \(J_{k+1}\) in distinct \(\mathcal {L}\)-classes, i.e. \(\left( {\begin{array}{c}n\\ k+1\end{array}}\right) \le n\), and so \(k+1\in \{0,1,n-1,n\}\). Since \(1\le k\le n-2\), it follows that \(k=n-2\).

It remains to show that \(e=1\). Since \(J_{n-1}\phi \subseteq J_1\) and \(I_{n-2}\phi =\{\emptyset \}\), the reasoning of the first paragraph of the proof of Lemma 2.6 applies here and we can conclude that \(J_{n-1}^{\mathscr {POI}_n}\phi =J_1^{\mathscr {POI}_n}\).

Suppose, by contradiction, that \(e\ne 1\). Then, \(|\mathop {\mathrm {Im}}\nolimits (e)|<n\) and so we may take \(i\in \Omega _n{\setminus }\mathop {\mathrm {Im}}\nolimits (e)\). Let \(h\in J_{n-1}^{\mathscr {POI}_n}\) be such that \(h\phi =\left( {\begin{array}{c}i\\ i\end{array}}\right) \). Hence, \( \emptyset =e\left( {\begin{array}{c}i\\ i\end{array}}\right) =(1\phi )(h\phi )=(1\cdot h)\phi =h\phi =\left( {\begin{array}{c}i\\ i\end{array}}\right) , \) a contradiction. Thus, \(1\phi =1\), as required. \(\square \)

Now, by Lemmas 3.2 and 2.7, we can deduce that, if \(1\le k\le n-2\), then (\(k=n-2\) and) \(\phi =\phi _g\), for some \(g\in \mathscr {S}_n\). This concludes the proof of the following description of the monoid of endomorphisms of S.

Theorem 3.3

Let \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\). Let \(\phi : S\longrightarrow S\) be any mapping. Then, \(\phi \) is an endomorphism of the semigroup S if and only if one of the following holds:

  1. (a)

    \(\phi \) is an automorphism and so \(\phi \) is the identity or \(\phi =\phi ^{\sigma }\);

  2. (b)

    there exist idempotents \(e,f\in S\) with \(e\not =f\) and \(ef=fe=f\) such that \(1\phi =e\) and \((S{\setminus }\{1\})\phi =\{f\}\);

  3. (c)

    \(\phi =\phi _g\), for some \(g\in \mathscr {S}_n\);

  4. (d)

    \(\phi \) is a constant mapping with idempotent value.

As a corollary of the this theorem, we finish this paper by counting the number of endomorphisms of \(\mathscr {POI}_n\) and of \(\mathscr {PO}_n\).

For \(S\in \{\mathscr {POI}_n,\mathscr {PO}_n\}\) and for each idempotent \(e\in S\), let

$$\begin{aligned} E_S(e)=\{f\in S\mid f^2=f\, {\text {and}}\, fe=ef=f\}. \end{aligned}$$

Then, according to Theorem 3.3, we have \( \sum _{e^2=e\in S}|E_S(e)| \) endomorphisms of S of type (b) and (d).

We start by considering \(\mathscr {POI}_n\).

Theorem 3.4

The semigroup \(\mathscr {POI}_n\) has \(2+n!+3^n\) endomorphisms.

Proof

First, recall that the idempotents of \(\mathscr {POI}_n\) are all the partial identities of \(\Omega _n\), and so we precisely have \(\left( {\begin{array}{c}n\\ k\end{array}}\right) \) idempotents of \(\mathscr {POI}_n\) with rank k, for all \(0\le k\le n\).

Let e be an idempotent of \(\mathscr {POI}_n\). Then, it is easy to show that \(f\in E_{\mathscr {POI}_n}(e)\) if and only if \(\mathop {\mathrm {Im}}\nolimits (f)\subseteq \mathop {\mathrm {Im}}\nolimits (e)\), for any idempotent f of \(\mathscr {POI}_n\). Hence, \(|E_{\mathscr {POI}_n}(e)|=2^{|\mathop {\mathrm {Im}}\nolimits (e)|}\).

It follows that the number of endomorphisms of \(\mathscr {POI}_n\) of type (b) and (d) is equal to

$$\begin{aligned} \sum _{e^2=e\in \mathscr {POI}_n}|E_{\mathscr {POI}_n}(e)| = \sum _{e^2=e\in \mathscr {POI}_n}2^{|\mathop {\mathrm {Im}}\nolimits (e)|} = \sum _{k=0}^{n}\left( {\begin{array}{c}n\\ k\end{array}}\right) 2^k=3^n, \end{aligned}$$

and so, since we have n! endomorphisms of type (c), as observed at the end of Sect. 2, and two automorphisms, we obtain a total of \(2+n!+3^n\) endomorphisms of \(\mathscr {POI}_n\), as required. \(\square \)

Counting the number of endomorphisms of \(\mathscr {PO}_n\) is more elaborated than for \(\mathscr {POI}_n\). First, we prove two lemmas.

Lemma 3.5

Let \(1\le k\le n\) and let e be an idempotent of \(\mathscr {PO}_n\) with rank k. Then, the set \(E_{\mathscr {PO}_n}(e)\) has as many elements as the number of idempotents of \(\mathscr {PO}_k\), i.e., \(E_{\mathscr {PO}_n}(e)\) has \(1+(\sqrt{5})^{k-1}((\frac{\sqrt{5}+1}{2})^k-(\frac{\sqrt{5}-1}{2})^k)\) elements.

Proof

Let X be any finite chain and denote by \(\mathscr {PO}(X)\) the monoid of all order-preserving partial transformations on X. Observe that \(\mathscr {PO}(X)\) and \(\mathscr {PO}_{|X|}\) are isomorphic monoids.

Let \(1\le k\le n\) and let e be an idempotent of \(\mathscr {PO}_n\) with rank k. Given an idempotent \(g\in \mathscr {PO}(\mathop {\mathrm {Im}}\nolimits (e))\), it is a routine matter to check that \(eg\in E_{\mathscr {PO}_n}(e)\) and, moreover, that the map \(\{f\in \mathscr {PO}(\mathop {\mathrm {Im}}\nolimits (e))\mid f^2=f\}\longrightarrow E_{\mathscr {PO}_n}(e)\), \(g\longmapsto eg\), is a bijection. Therefore, \(|E_{\mathscr {PO}_n}(e)|=|\{f\in \mathscr {PO}_k\mid f^2=f\}|\).

Now, it remains to recall that Laradji and Umar proved that the number of idempotents of \(\mathscr {PO}_k\) is equal to \(1+(\sqrt{5})^{k-1}((\frac{\sqrt{5}+1}{2})^k-(\frac{\sqrt{5}-1}{2})^k)\) (see [22, Theorem 3.8]).

\(\square \)

Lemma 3.6

The number of idempotents of \(\mathscr {PO}_n\) with rank k is \(\sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) \), for \(1\le k\le n\).

Proof

Let X be any finite chain and denote by \(\mathscr {O}(X)\) the monoid of all order-preserving full transformations on X. As for order-preserving partial transformations, we have that \(\mathscr {O}(X)\) and \(\mathscr {O}_{|X|}\) are isomorphic monoids.

Let \(1\le k\le n\). Since an idempotent fixes its image, given an idempotent e of \(\mathscr {PO}_n\), it is clear that \(\mathop {\mathrm {Im}}\nolimits (e)\subseteq \mathop {\mathrm {Dom}}\nolimits (e)\), and so \(e\in \mathscr {O}(\mathop {\mathrm {Dom}}\nolimits (e))\). Hence, an element e of \(\mathscr {PO}_n\) is an idempotent with rank k if and only if e is an idempotent of \(\mathscr {O}(X)\) with rank k, for some subset X of \(\Omega _n\) such that \(|X|\ge k\). Therefore, the number of idempotents of \(\mathscr {PO}_n\) with rank k is

$$\begin{aligned} \sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) |\{e\in J_k^{\mathscr {O}_i}\mid e^2=e\}|, \end{aligned}$$

and so, since \(|\{e\in J_k^{\mathscr {O}_i}\mid e^2=e\}|=\left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) \), by [23, Corollary 4.4], for \(i\ge k\), the lemma is proved. \(\square \)

Observe that, by the previous lemma and [22, Theorem 3.8], we obtain the following equality:

$$\begin{aligned} \sum _{k=1}^{n}\sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) = (\sqrt{5})^{n-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^n-\left( \frac{\sqrt{5}-1}{2}\right) ^n\right) . \end{aligned}$$
(1)

Now, we can calculate the number of endomorphisms of \(\mathscr {PO}_n\).

Theorem 3.7

The semigroup \(\mathscr {PO}_n\) has

$$\begin{aligned}&3 + n! + (\sqrt{5})^{n-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^n \!\!\! -\left( \frac{\sqrt{5}-1}{2}\right) ^n\right) \\&\quad + \sum _{k=1}^{n} (\sqrt{5})^{k-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^k \!\!\! -\left( \frac{\sqrt{5}-1}{2}\right) ^k\right) \sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) \end{aligned}$$

endomorphisms.

Proof

The number of endomorphisms of \(\mathscr {PO}_n\) of type (b) and (d) is equal to

$$\begin{aligned} \sum _{e^2=e\in \mathscr {PO}_n}|E_{\mathscr {PO}_n}(e)|= & {} 1+\sum _{e^2=e\in \mathscr {PO}_n{\setminus }\{\emptyset \}}|\{f\in \mathscr {PO}_{|\mathop {\mathrm {Im}}\nolimits (e)|}\mid f^2=f\}| \\= & {} 1 + \sum _{k=1}^{n} |\{f\in \mathscr {PO}_k\mid f^2=f\}| |\{e\in J_k^{\mathscr {PO}_n}\mid e^2=e\}| \\= & {} 1 + \sum _{k=1}^{n} \left( 1+(\sqrt{5})^{k-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^k\right. \right. \\&\left. \left. -\left( \frac{\sqrt{5}-1}{2}\right) ^k\right) \right) \sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) \\= & {} 1 + \sum _{k=1}^{n}\sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) + \sum _{k=1}^{n} (\sqrt{5})^{k-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^k\right. \\&\left. -\left( \frac{\sqrt{5}-1}{2}\right) ^k\right) \sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) , \end{aligned}$$

by Lemmas 3.5 and 3.6, i.e. equal to

$$\begin{aligned} 1+ & {} (\sqrt{5})^{n-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^n-\left( \frac{\sqrt{5}-1}{2}\right) ^n\right) \\+ & {} \sum _{k=1}^{n} (\sqrt{5})^{k-1}\left( \left( \frac{\sqrt{5}+1}{2}\right) ^k-\left( \frac{\sqrt{5}-1}{2}\right) ^k\right) \sum _{i=k}^{n}\left( {\begin{array}{c}n\\ i\end{array}}\right) \left( {\begin{array}{c}i+k-1\\ 2k-1\end{array}}\right) , \end{aligned}$$

by the equality (1). Once again we have n! endomorphisms of type (c) and two automorphisms, so the result follows. \(\square \)