Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-11T12:13:58.300Z Has data issue: false hasContentIssue false

Proper Lie automorphisms of incidence algebras

Published online by Cambridge University Press:  07 February 2022

Érica Z. Fornaroli
Affiliation:
Departamento de Matemática, Universidade Estadual de Maringá, Maringá, PR, CEP: 87020–900, Brazilezancanella@uem.br, easjunior@uem.br
Mykola Khrypchenko
Affiliation:
Departamento de Matemática, Universidade Federal de Santa Catarina, Campus Reitor João David Ferreira Lima, Florianópolis, SC, CEP: 88040–900, Brazilnskhripchenko@gmail.com
Ednei A. Santulo Jr.
Affiliation:
Departamento de Matemática, Universidade Estadual de Maringá, Maringá, PR, CEP: 87020–900, Brazilezancanella@uem.br, easjunior@uem.br
Rights & Permissions [Opens in a new window]

Abstract

Let X be a finite connected poset and K a field. We study the question, when all Lie automorphisms of the incidence algebra I(X, K) are proper. Without any restriction on the length of X, we find only a sufficient condition involving certain equivalence relation on the set of maximal chains of X. For some classes of posets of length one, such as finite connected crownless posets (i.e., without weak crown subposets), crowns, and ordinal sums of two anti-chains, we give a complete answer.

Type
Research Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Glasgow Mathematical Journal Trust

Introduction

A Lie isomorphism of associative rings $(R,\cdot)$ and $(S,\cdot)$ is an isomorphism of the corresponding Lie rings $\left(R,[\,,]\right)$ and $\left(S,[\, ,]\right)$ , where $[a,b]=a\cdot b-b\cdot a$ . If $R=S$ , a Lie isomorphism $R\to S$ is called a Lie automorphism of R. If, moreover, R and S are algebras, it is natural to require Lie isomorphisms $R\to S$ to be linear. Any bijective map of the form $\phi+\nu$ , where $\phi$ is either an isomorphism $R\to S$ or the negative of an anti-isomorphism $R\to S$ , and $\nu$ is an additive map on R with values in the center of S whose kernel contains [R, R], is always a Lie isomorphism. Such Lie isomorphisms are called proper. In most of the cases studied in the literature, these are the only examples of Lie isomorphisms. Indeed, this is true for Lie automorphisms of full matrix rings $M_n(R)$ over division rings R with ${\rm{char}}(R)\not \in \{ 2,3\}$ as proved in [Reference Hua14], for Lie isomorphisms of primitive [Reference Martindale16], simple [Reference Martindale18] and prime rings [Reference Martindale17], for Lie automorphisms of upper triangular matrix algebras $T_n(R)$ over commutative rings [Reference Cao7, Reference Doković9], and for Lie isomorphisms of block-triangular matrix algebras over a UFD [Reference Cecil8].

In [Reference Fornaroli, Khrypchenko and Santulo13], we described Lie automorphisms of the incidence algebra I(X, K) of a finite connected poset X over a field K. In general, they are not proper as shown in [Reference Fornaroli, Khrypchenko and Santulo13, Example 5.20], but, for some classes of posets, every Lie automorphism of I(X, K) is proper. For instance, if X is a chain of cardinality n, then $I(X,K)\cong T_n(K)$ , and thus every Lie automorphism of I(X, K) is proper in view of [Reference Fornaroli, Khrypchenko and Santulo13, Corollary 5.19] (see also [Reference Doković9, Theorem 6]). So the following question arises.

Question. What are the necessary and sufficient conditions on a finite connected poset X such that all Lie automorphisms of I(X,K) are proper?

In this paper, we give a partial answer to this question. Namely, for a general X we find only a sufficient condition (see Corollary 3.12), and for some particular classes of posets of length one X we give a complete answer (see Corollaries 4.7 and 4.12 and Proposition 4.14).

More precisely, our work is organized as follows. Section 1 serves as a background on posets, incidence algebras, and maps on them. In particular, we recall all the necessary definitions from [Reference Fornaroli, Khrypchenko and Santulo13] and introduce some new notations. In Section 2, we reduce the question of when all Lie automorphisms of I(X, K) are proper to a purely combinatorial property of X dealing with a certain group $\mathcal{AM}(X)$ of bijections on maximal chains of X (see Theorem 2.10). In our terminology, we prove that every Lie automorphism of I(X, K) is proper if and only if every bijection $\theta\in\mathcal{AM}(X)$ is proper. In Section 3, we introduce an equivalence relation $\sim$ on maximal chains of X and show that any $\theta\in\mathcal{AM}(X)$ induces isomorphisms or anti-isomorphisms between certain subsets of X (the so-called supports of $\sim$ -classes), as proved in Theorem 3.10. Consequently, if all the maximal chains of X are equivalent, then all Lie automorphisms of I(X, K) are proper (see Corollary 3.12). The equivalence $\sim$ is just the equality relation whenever X is of length one, hence this situation is treated separately. This is done in Section 4. We first consider the case when X has no crown subset and give a full description of those X for which any $\theta\in\mathcal{AM}(X)$ is proper (see Corollary 4.7). We then pass to two specific classes of X: n-crowns ${\rm{Cr}}_n$ and ordinal sums of two anti-chains ${\rm{K}}_{m,n}$ . If $X={\rm{Cr}}_n$ , then we explicitly describe the group $\mathcal{AM}(X)$ (see Proposition 4.10) and its subgroup of proper bijections (see Proposition 4.11). It follows that all $\theta\in\mathcal{AM}({\rm{Cr}}_n)$ are proper exactly when $n=2$ (see Corollary 4.12). If $X={\rm{K}}_{m,n}$ , then there are only proper $\theta\in\mathcal{AM}(X)$ as proved in Proposition 4.14.

1. Preliminaries

1.1. Automorphisms and anti-automorphisms

Let A be an algebra. We denote by Aut(A) the group of (linear) automorphisms of A, by ${\rm{Aut}}^-(A)$ the set of (linear) anti-automorphisms of A and by ${\rm{Aut}}^\pm(A)$ the union ${\rm{Aut}}(A)\cup {\rm{Aut}}^-(A)$ . Observe that the union is non-disjoint if and only if A is commutative, in which case ${\rm{Aut}}(A)={\rm{Aut}}^-(A)$ . The set ${\rm{Aut}}^\pm(A)$ is a group under the composition, and moreover, if A is non-commutative and ${\rm{Aut}}^-(A)\ne\emptyset$ , then Aut(A) is a (normal) subgroup of ${\rm{Aut}}^\pm(A)$ of index 2. In particular, $|{\rm{Aut}}(A)|=|{\rm{Aut}}^-(A)|$ , whenever A is non-commutative and ${\rm{Aut}}^-(A)\ne\emptyset$ . We use the analogous notations Aut(X), ${\rm{Aut}}^-(X)$ and ${\rm{Aut}}^\pm(X)$ for the group of automorphisms of a poset X, the set of anti-automorphisms of X and the group ${\rm{Aut}}(X)\cup {\rm{Aut}}^-(X)$ , respectively. As above, Aut(X) either coincides with ${\rm{Aut}}^\pm(X)$ or is a subgroup of index 2 in ${\rm{Aut}}^\pm(X)$ (if X is not an anti-chain and ${\rm{Aut}}^-(X)\ne\emptyset$ ).

1.2. Posets

Let $(X,\le)$ be a partially ordered set (which we usually shorten to “poset”) and $x,y\in X$ . The interval from x to y is the set $\lfloor x,y \rfloor=\{z\in X \,{:}\, x\leq z\leq y\}$ . The poset X is said to be locally finite if all the intervals of X are finite. A chain in X is a linearly ordered (under the induced order) subset of X. The length of a finite chain $C\subseteq X$ is defined to be $|C|-1$ . The length Footnote 1 of a finite poset X, denoted by l(X), is the maximum length of chains $C\subseteq X$ . A walk in X is a sequence $x_0,x_1,\dots,x_m\in X$ , such that $x_i$ and $x_{i+1}$ are comparable and $l(\lfloor x_i,x_{i+1}\rfloor)=1$ (if $x_i\le x_{i+1}$ ) or $l(\lfloor x_{i+1},x_i\rfloor)=1$ (if $x_{i+1}\le x_i$ ) for all $i=0,\dots,m-1$ . A walk $x_0,x_1,\dots,x_m$ is closed if $x_0=x_m$ . A path is a walk satisfying $x_i\ne x_j$ for $i\ne j$ . A cycle is a closed walk $x_0,x_1,\dots,x_m=x_0$ in which $m\ge 4$ and $x_i=x_j\Rightarrow \{i,j\}=\{0,m\}$ for $i\ne j$ . We say that X is connected if for any pair of $x,y\in X$ there is a path $x=x_0,\dots,x_m=y$ . We will denote by Min(X) (resp. Max(X)) the set of minimal (resp. maximal) elements of X. If X is connected and $|X|>1$ , then ${\rm{Min}}(X)\cap {\rm{Max}}(X)=\emptyset$ .

1.3. Incidence algebras

Let X be a locally finite poset and K a field. The incidence algebra [Reference Rota20] I(X, K) of X over K is the K-space of functions $f\,{:}\,X\times X\to K$ such that $f(x,y)=0$ if $x\nleq y$ . This is a unital K-algebra under the convolution product

\begin{equation*} (fg)(x,y)=\sum_{x\leq z\leq y}f(x,z)g(z,y), \end{equation*}

for any $f, g\in I(X,K)$ . Its identity element $\delta$ is given by

\begin{align*}\delta(x,y)=\begin{cases}1, & x=y,\\ \\[-7pt] 0, & x\ne y.\end{cases}\end{align*}

Throughout the rest of the paper, X will stand for a connected finite poset. Then I(X, K) admits the standard basis $\{e_{xy} \,{:}\, x\leq y\}$ , where

\begin{align*}e_{xy}(u,v)=\begin{cases}1, & (u,v)=(x,y),\\ \\[-7pt] 0, & (u,v)\ne(x,y).\end{cases}\end{align*}

We will write $e_{x}=e_{xx}$ . Denote also $B=\{e_{xy} \,{:}\, x<y\}$ . It is a well-known fact (see [Reference Spiegel and O’Donnell22, Theorem 4.2.5]) that the Jacobson radical of I(X, K) is

\begin{equation*} J(I(X,K))=\{f\in I(X,K) \,{:}\, f(x,x)=0 \text{ for all } x\in X\}={\rm{Span}}_K B. \end{equation*}

Diagonal elements of I(X, K) are those $f\in I(X,K)$ satisfying $f(x,y)=0$ for $x\neq y$ . They form a commutative subalgebra D(X, K) of I(X, K) spanned by $\{e_{x} \,{:}\, x \in X\}$ . Clearly, each $f\in I(X,K)$ can be uniquely written as $f=f_D+f_J$ with $f_D\in D(X,K)$ and $f_J\in J(I(X,K))$ .

1.4. Decomposition of $\boldsymbol\phi\boldsymbol\in \boldsymbol{Aut}^{\boldsymbol\pm}(\boldsymbol{{I}}(\boldsymbol{{X}}, \boldsymbol{{K}}))$

Now, we recall the descriptions of automorphisms and anti-automorphisms of I(X, K). Firstly, if X and Y are finite posets and $\lambda\,{:}\, X\to Y$ is an isomorphism (resp. anti-isomorphism), then $\lambda$ induces an isomorphism (resp. anti-isomorphism) $\hat\lambda\,{:}\, I(X,K)\to I(Y,K)$ defined by $\hat\lambda(e_{xy})=e_{\lambda(x)\lambda(y)}$ (resp. $\hat\lambda(e_{xy})=e_{\lambda(y)\lambda(x)}$ ), for all $x\leq y$ in X. An element $\sigma\in I(X,K)$ such that $\sigma(x,y)\neq 0$ , for all $x\leq y$ , and $\sigma(x,y)\sigma(y,z)=\sigma(x,z)$ whenever $x\leq y\leq z$ , determines an automorphism $M_{\sigma}$ of I(X, K) by $M_{\sigma}(e_{xy})=\sigma(x,y)e_{xy}$ , for all $x\leq y$ . Such automorphisms are called multiplicative. Any automorphism (anti-automorphism) of I(X, K) decomposes as

(1) \begin{align}\phi=\hat\lambda \circ \xi \circ M_{\sigma},\end{align}

where $\lambda\in {\rm{Aut}}(X)$ (resp. ${\rm{Aut}}^-(X)$ ), $\xi $ is an inner automorphism, and $M_{\sigma}$ is a multiplicative automorphism of I(X, K). For automorphisms, this was proved in [Reference Baclawski1, Theorem 5] and for anti-automorphisms in [Reference Brusamarello, Fornaroli and Santulo3, Theorem 5] (for more results on automorphisms and anti-automorphisms of incidence algebras see [Reference Brusamarello, Fornaroli and Santulo3, Reference Brusamarello, Fornaroli and Santulo4, Reference Brusamarello, Fornaroli and Santulo5, Reference Brusamarello and Lewis6, Reference Drozd and Kolesnik12, Reference Khripchenko15, Reference Spiegel21, Reference Stanley23]).

1.5. Lie automorphisms of incidence algebras

In this section, we introduce several new notations and recall some definitions and results from [Reference Fornaroli, Khrypchenko and Santulo13].

We denote by $\mathcal{C}(X)$ the set of maximal chains in X. Let $C\,{:}\, u_1<u_2<\dots<u_m$ in $\mathcal{C}(X)$ . A bijection $\theta\,{:}\,B\to B$ is increasing (resp. decreasing) on C if there exists $D\,{:}\, v_1<v_2<\dots<v_m$ in $\mathcal{C}(X)$ such that $\theta(e_{u_iu_j})=e_{v_iv_j}$ for all $1\le i<j\le m$ (resp. $\theta(e_{u_iu_j})=e_{v_{m-j+1}v_{m-i+1}}$ for all $1\le i<j\le m$ ). In this case, we write $\theta(C)=D$ . Moreover, we say that $\theta$ is monotone on maximal chains in X if, for any $C\in \mathcal{C}(X)$ , $\theta$ is increasing or decreasing on C. We denote by $\mathcal{M}(X)$ the set of bijections $B\to B$ which are monotone on maximal chains in X. It is easy to see that $\mathcal{M}(X)$ is a subgroup of the symmetric group S(B). Each $\theta\in\mathcal{M}(X)$ induces a bijection on $\mathcal{C}(X)$ which maps C to $\theta(C)$ .

Let $\theta\,{:}\,B\to B$ be a bijection and $X^2_<=\{(x,y)\in X^2\,{:}\, x<y\}$ . A map $\sigma\,{:}\,X^2_<\to K^*$ is compatible with $\theta$ if $\sigma(x,z)=\sigma(x,y)\sigma(y,z)$ whenever $\theta(e_{xz})=\theta(e_{xy})\theta(e_{yz})$ , and $\sigma(x,z)=-\sigma(x,y)\sigma(y,z)$ whenever $\theta(e_{xz})=\theta(e_{yz})\theta(e_{xy})$ .

Let $\theta\,{:}\,B\to B$ be a bijection and $\Gamma\,{:}\, u_0,u_1,\dots,u_m=u_0$ a closed walk in X. In [Reference Fornaroli, Khrypchenko and Santulo13], we introduced the following 4 functions $X\to\mathbb{N}$ :

\begin{align*}s^+_{\theta,\Gamma}(z)&=|\{i\,{:}\, u_i<u_{i+1}\text{ and }\exists w>z\text{ such that }\theta(e_{zw})=e_{u_iu_{i+1}}\}|,\\[2pt] s^-_{\theta,\Gamma}(z)&=|\{i\,{:}\, u_i>u_{i+1}\text{ and }\exists w>z\text{ such that }\theta(e_{zw})=e_{u_{i+1}u_i}\}|,\\[2pt] t^+_{\theta,\Gamma}(z)&=|\{i\,{:}\, u_i<u_{i+1}\text{ and }\exists w<z\text{ such that }\theta(e_{wz})=e_{u_iu_{i+1}}\}|,\\[2pt] t^-_{\theta,\Gamma}(z)&=|\{i\,{:}\, u_i>u_{i+1}\text{ and }\exists w<z\text{ such that }\theta(e_{wz})=e_{u_{i+1}u_i}\}|.\end{align*}

We call the bijection $\theta\,{:}\,B\to B$ admissible if

(2) \begin{align} s^+_{\theta,\Gamma}(z)-s^-_{\theta,\Gamma}(z)=t^+_{\theta,\Gamma}(z)-t^-_{\theta,\Gamma}(z), \end{align}

for any closed walk $\Gamma\,{:}\,u_0,u_1,\dots,u_m=u_0$ in X and for all $z\in X$ . In particular, if X is a tree, then any bijection $\theta\,{:}\,B\to B$ is admissible. We denote by $\mathcal{AM}(X)$ the set of those $\theta\in\mathcal{M}(X)$ which are admissible.

Let $X=\{x_1,\dots, x_n\}$ . Given $\theta\in \mathcal{AM}(X)$ , a map $\sigma\,{:}\,X_<^2\to K^*$ compatible with $\theta$ and a sequence $c=(c_1,\dots,c_n)\in K^n$ such that $\sum_{i=1}^nc_i\in K^*$ , we define in [Reference Fornaroli, Khrypchenko and Santulo13, Definition 5.17] the following elementary Lie automorphism $\tau=\tau_{\theta,\sigma,c}$ of I(X, K) where, for any $e_{xy}\in B$ ,

\begin{equation*}\tau(e_{xy})=\sigma(x,y)\theta(e_{xy}),\end{equation*}

and $\tau|_{D(X,K)}$ is determined by

\begin{equation*}\tau(e_{x_i})(x_1,x_1)=c_i,\end{equation*}

$i=1,\dots,n$ , as in Lemmas 5.8 and 5.16 from [Reference Fornaroli, Khrypchenko and Santulo13]. As in [Reference Fornaroli, Khrypchenko and Santulo13, Definition 5.15], we say that $\tau$ induces the pair $(\theta,\sigma)$ and in some situations we write $\theta=\theta_{\tau}$ .

As in [Reference Fornaroli, Khrypchenko and Santulo13], we denote by LAut(I(X, K)) the group of Lie automorphisms of I(X, K) and by $\widetilde{{\rm{LAut}}}(I(X,K))$ its subgroup of elementary Lie automorphisms. We will also use the notation $\textrm{Inn}_1(I(X,K))$ for the subgroup of inner automorphisms consisting of conjugations by $\beta\in I(X,K)$ with $\beta_D=\delta$ .

Theorem 1.1. [Reference Fornaroli, Khrypchenko and Santulo13, Theorem 4.15]The group LAut(I(X, K)) is isomorphic to the semidirect product ${\rm{Inn}}_1(I(X, K))\rtimes\widetilde{{\rm{LAut}}}(I(X, K))$ .

2. Proper Lie automorphisms of I(X, K) and proper bijections of B

Let $\varphi\in {\rm{LAut}}(I(X, K))$ . Then $\varphi=\psi\circ \tau_{\theta,\sigma,c}$ , where $\psi\in {\rm{Inn}}_1(I(X, K))$ and $\tau_{\theta,\sigma,c}$ is an elementary Lie automorphism of I(X, K), by Theorem 1.1. Note that $\varphi$ is proper if and only if $\tau_{\theta,\sigma,c}$ is proper. Therefore, all Lie automorphisms of I(X, K) are proper if and only if all elementary Lie automorphisms of I(X, K) are proper.

Let $\varphi=\tau_{\theta,\sigma,c}$ be an elementary Lie automorphism of I(X, K). Suppose that $\varphi$ is proper, $\varphi=\phi+\nu$ , where $\phi\in {\rm{Aut}}(I(X, K))$ or $-\phi\in {\rm{Aut}}^-(I(X, K))$ and $\nu$ is a linear central-valued map on I(X, K) such that $\nu([I(X, K),I(X, K)])=\{0\}$ . If $x<y$ , then $e_{xy}\in J(I(X, K))=[I(X, K),I(X, K)]$ , by [Reference Fornaroli, Khrypchenko and Santulo13, Proposition 2.3]. Thus

(3) \begin{align}\varphi(e_{xy})=\phi(e_{xy}), \forall x<y.\end{align}

By [Reference Spiegel and O’Donnell22, Corollary 1.3.15], for each $x\in X$ there is $\alpha_x\in K$ such that

(4) \begin{align}\varphi(e_{x})=\phi(e_{x})+\alpha_x\delta.\end{align}

Suppose firstly that $\phi\in {\rm{Aut}}(I(X, K))$ . Then, by (1), $\phi=\hat\lambda \circ \xi_f \circ M_{\tau}$ , where $\lambda\in {\rm{Aut}}(X)$ , $\xi_f$ is an inner automorphism and $M_{\tau}$ is a multiplicative automorphism of I(X, K). Thus, by (3),

(5) \begin{align} \theta(e_{xy})= & \sigma(x,y)^{-1}(\hat\lambda \circ \xi_f \circ M_{\tau})(e_{xy})=\sigma(x,y)^{-1}(\hat\lambda \circ \xi_f)(\tau(x,y)e_{xy}) \nonumber\\[3pt] = & \sigma(x,y)^{-1}\hat\lambda(\tau(x,y)fe_{xy}f^{-1})=\sigma(x,y)^{-1}\tau(x,y)\hat\lambda(f)e_{\lambda(x)\lambda(y)}\hat\lambda(f)^{-1}.\end{align}

Analogously, if $-\phi\in {\rm{Aut}}^-(I(X, K))$ , then, by (1), $\phi=\hat\lambda \circ \xi_f \circ M_{\tau}$ , where $\lambda\in {\rm{Aut}}^-(X)$ , $\xi_f$ is an inner automorphism and $M_{\tau}$ is a multiplicative automorphism of I(X, K). Thus, by (3),

(6) \begin{align} \theta(e_{xy})= \sigma(x,y)^{-1}(\hat\lambda \circ \xi_f \circ M_{\tau})(e_{xy})= \sigma(x,y)^{-1}\tau(x,y)\hat\lambda(f)^{-1}e_{\lambda(y)\lambda(x)}\hat\lambda(f).\end{align}

Remark 2.1. Let $x\leq y, u\leq v$ in X, $\alpha\in K^{\ast}$ and $h\in I(X,K)$ an invertible element. Note that if $he_{xy}h^{-1}=\alpha e_{uv}$ , then $(x,y)=(u,v)$ .

Definition 2.2. A bijection $\theta\,{:}\,B\to B$ is said to be properif there exists $\lambda\in {\rm{Aut}}^\pm(X)$ such that $\theta(e_{xy})=\hat\lambda(e_{xy})$ for all $e_{xy}\in B$ . The proper bijections of B form a group, which we denote by $\mathcal{P}(X)$ .

Proposition 2.3. If $|X|>2$ , then the group $\mathcal{P}(X)$ is isomorphic to ${\rm{Aut}}^\pm(X)$ .

Proof. The map sending $\lambda\in {\rm{Aut}}(X)$ (resp. $\lambda\in {\rm{Aut}}^-(X)$ ) to $\theta\in\mathcal{P}(X)$ , such that $\theta(e_{xy})=e_{\lambda(x)\lambda(y)}$ (resp. $\theta(e_{xy})=e_{\lambda(y)\lambda(x)}$ ), is an epimorphism from ${\rm{Aut}}^\pm(X)$ to $\mathcal{P}(X)$ . We only need to prove that it is injective.

It is injective on Aut(X). Indeed, take $\lambda,\mu\in {\rm{Aut}}(X)$ such that $(\lambda(x),\lambda(y))=(\mu(x),\mu(y))$ for all $x<y$ in X. Let x be an arbitrary element of X. Since X is connected and $|X|>1$ , there is $y\in X$ such that either $y<x$ or $y>x$ . In both cases, we get $\lambda(x)=\mu(x)$ . Thus, $\lambda=\mu$ . Similarly, one proves injectivity on ${\rm{Aut}}^-(X)$ .

Assume now that there are $\lambda\in {\rm{Aut}}(X)$ and $\mu\in {\rm{Aut}}^-(X)$ such that $(\lambda(x),\lambda(y))=(\mu(y),\mu(x))$ for all $x<y$ in X. We first show that X must have length at most 1. Indeed, if there were $x<y<z$ in X, then we would have $(\lambda(x),\lambda(y))=(\mu(y),\mu(x))$ and $(\lambda(y),\lambda(z))=(\mu(z),\mu(y))$ , whence $\lambda(x)=\lambda(z)$ , a contradiction. Now, consider a triple $x,y,z\in X$ with $x>z<y$ . It follows from $(\lambda(z),\lambda(x))=(\mu(x),\mu(z))$ and $(\lambda(z),\lambda(y))=(\mu(y),\mu(z))$ that $\mu(x)=\mu(y)$ , a contradiction. Similarly, the existence of a triple $x,y,z\in X$ with $x<z>y$ leads to $\lambda(x)=\lambda(y)$ . If there are two incomparable elements $x,y\in X$ , then there exists a sequence $x=x_1,\dots,x_m=y$ , where $m\ge 3$ and either $x_1<x_2>x_3$ or $x_1>x_2<x_3$ . In both cases, we come to a contradiction. Thus, X is of length at most 1 and any two elements of X are comparable, which means that X is either a singleton or a chain of length 1.

Remark 2.4. If X is a chain of length 1, then $|\mathcal{P}(X)|=1$ , while $|{\rm{Aut}}^\pm(X)|=2$ .

Proposition 2.5. We have $\mathcal{P}(X)\subseteq\mathcal{AM}(X)$ .

Proof. Let $\theta\in\mathcal{P}(X)$ . Then there exists $\lambda\in {\rm{Aut}}^\pm(X)$ as in Definition 2.2. In both cases, $\theta$ is the restriction of $\hat{\lambda}$ to B and, since $\hat{\lambda}$ or $-\hat{\lambda}$ is an elementary Lie automorphism of I(X, K) by [Reference Fornaroli, Khrypchenko and Santulo13, Remark 4.8], then $\theta\in\mathcal{AM}(X)$ by [Reference Fornaroli, Khrypchenko and Santulo13, Remark 5.10].

Lemma 2.6. Let $\varphi\in\widetilde{{\rm{LAut}}}(I(X,K))$ inducing a pair $(\theta,\sigma)$ . Then $\theta$ is proper if, and only if, $\varphi$ is proper.

Proof. Assume first that $\theta(e_{xy})=e_{\lambda(x)\lambda(y)}$ for some $\lambda\in {\rm{Aut}}(X)$ . Then $\theta$ is increasing on any maximal chain in X and thus $\sigma(x,z)=\sigma(x,y)\sigma(y,z)$ for all $x<y<z$ . Extending $\sigma$ to $X^2_\le=\{(x,y) \,{:}\, x\le y\}$ by means of $\sigma(x,x)=1$ for all $x\in X$ , we obtain the multiplicative automorphism $M_\sigma\in {\rm{Aut}}(I(X,K))$ . Consider $\psi=\varphi\circ M^{-1}_\sigma$ . Notice that $\psi(e_{xy})=e_{\lambda(x)\lambda(y)}$ for all $x<y$ and $\psi(e_x)=\varphi(e_x)$ for all $x\in X$ . It suffices to prove that $\psi$ is proper. Indeed, if $\psi=\phi+\nu$ , then $\varphi=\psi\circ M_\sigma=\phi\circ M_\sigma+\nu$ since $M_\sigma$ is identity on D(X, R).

Clearly, $\psi(e_{xy})=\hat\lambda(e_{xy})$ for all $x<y$ , so $\hat\lambda$ is a candidate for $\phi$ . It remains to prove (4) for $\psi$ , that is, to show that $\psi(e_x)=e_{\lambda(x)}+\alpha_x\delta$ for some $\alpha_x\in K$ . The latter is equivalent to

(7) \begin{align} \psi(e_x)(y,y)=\psi(e_x)(\lambda(x),\lambda(x))-1, \end{align}

for all $y\ne \lambda(x)$ . Given $x,y\in X$ with $y\ne \lambda(x)$ , we choose a path $\lambda(x)=u_0,\dots,u_m=y$ from $\lambda(x)$ to y. Let $v_i\in X$ such that $\lambda(v_i)=u_i$ , $0\le i\le m$ . In particular, $v_0=x$ . Then

(8) \begin{align} \nonumber\\[-35pt] \psi(e_x)(y,y)=\psi(e_x)(\lambda(x),\lambda(x))+\sum_{i=0}^{m-1}(\psi(e_x)(u_{i+1},u_{i+1})-\psi(e_x)(u_i,u_i)). \end{align}

If $u_0<u_1$ , then $\theta(e_{xv_1})=e_{u_0u_1}$ , so $\psi(e_x)(u_1,u_1)-\psi(e_x)(u_0,u_0)=-1$ by [Reference Fornaroli, Khrypchenko and Santulo13, Lemma 5.7]. Similarly, if $u_0>u_1$ , then $\theta(e_{v_1x})=e_{u_1u_0}$ , so $\psi(e_x)(u_0,u_0)-\psi(e_x)(u_1,u_1)=1$ . In any case $\psi(e_x)(u_1,u_1)-\psi(e_x)(u_0,u_0)=-1$ . Observe that $v_i\ne x$ for all $i>0$ . Hence $\psi(e_x)(u_{i+1},u_{i+1})-\psi(e_x)(u_i,u_i)=0$ for all such i by [Reference Fornaroli, Khrypchenko and Santulo13, Lemma 5.7]. It follows that the sum on the right-hand side of (8) has only one non-zero term which equals $-1$ , proving (7).

The case $\theta(e_{xy})=e_{\lambda(y)\lambda(x)}$ , where $\lambda\in {\rm{Aut}}^-(X)$ , is similar.

Conversely, suppose that $\varphi=\phi+\nu$ , where $\phi\in {\rm{Aut}}(I(X,K))$ (resp. $-\phi\in {\rm{Aut}}^-(I(X,K))$ ) and $\nu$ is a linear central-valued map on I(X, K) annihilating [I(X, K), I(X, K)]. It follows from (5) (resp. (6)) and Remark 2.1 that $\theta(e_{xy})=e_{\lambda(x)\lambda(y)}$ ( $\theta(e_{xy})=e_{\lambda(y)\lambda(x)}$ ) for all $e_{xy}\in B$ , where $\lambda\in {\rm{Aut}}(X)$ ( $\lambda\in {\rm{Aut}}^-(X)$ ). Therefore, $\theta\in\mathcal{P}(X)$ .

Lemma 2.7. Let $\theta\in\mathcal{M}(X)$ . Let $C_1,C_2\in\mathcal{C}(X)$ such that $\theta$ is increasing on $C_1$ and decreasing on $C_2$ . If there exist $x,y\in C_1\cap C_2$ and $x<y$ , then x is the minimum of $C_1$ and $C_2$ and y is the maximum of $C_1$ and $C_2$ .

Proof. We first notice that there is $z\in C_1$ such that $z<x$ ( $y<z$ ) if, and only if, there is $z'\in C_2$ such that $z'<x$ ( $y<z'$ ), by the maximality of $C_1$ and $C_2$ . Suppose that are $z\in C_1$ and $z'\in C_2$ such that $z,z'<x$ . Then $\theta(e_{zx})\theta(e_{xy})=\theta(e_{zy})$ and $\theta(e_{xy})\theta(e_{z\prime x})=\theta(e_{z^{\prime}y})$ . Thus, there are $s<u<v<t$ such that $\theta(e_{zx})=e_{su}$ , $\theta(e_{xy})=e_{uv}$ , $\theta(e_{z^{\prime}x})=e_{vt}$ and $\theta(e_{zy})=e_{sv}$ , $\theta(e_{z^{\prime}y})=e_{ut}$ . If $\theta^{-1}$ is increasing on a maximal chain containing $s<u<v<t$ , then $\theta^{-1}(e_{st})=\theta^{-1}(e_{su})\theta^{-1}(e_{ut})=e_{zx}e_{z^{\prime}y}$ which implies $z'=x$ , a contradiction. If $\theta^{-1}$ is decreasing on a maximal chain containing $s<u<v<t$ , then $\theta^{-1}(e_{st})=\theta^{-1}(e_{vt})\theta^{-1}(e_{sv})=e_{z^{\prime}x}e_{zy}$ which implies $z=x$ , a contradiction. Therefore, x is the minimum of $C_1$ and $C_2$ . Analogously, y is the maximum of $C_1$ and $C_2$ .

Lemma 2.8. For any $\theta\in\mathcal{M}(X)$ , there is $\sigma\,{:}\,X^2_<\to K^*$ compatible with $\theta$ .

Proof. Let $\theta\in\mathcal{M}(X)$ and $\mathcal C_i$ ( $\mathcal C_d$ ) be the set of all maximal chains in X on which $\theta$ is increasing (decreasing). Let $(x,y)\in X^2_<$ . If $x\in {\rm{Min}}(X)$ , we set $\sigma(x,y)=1$ . Otherwise, by Lemma 2.7, both x and y belong only to maximal chains from $\mathcal{C}_i$ or only to maximal chains from $\mathcal{C}_d$ . In the former case we set $\sigma(x,y)=1$ and, in the latter one, we set $\sigma(x,y)=-1$ .

Let $x<y<z$ in X. Again, by Lemma 2.7, those three elements can be simultaneously only in maximal chains from $\mathcal{C}_i$ or only in maximal chains from $\mathcal{C}_d$ . If they belong to maximal chains from $\mathcal{C}_i$ , then $\sigma(x,z)=1=\sigma(x,y)\sigma(y,z)$ . Otherwise, there are two situations to be considered: $x\in {\rm{Min}}(X)$ or $x\not\in {\rm{Min}}(X)$ . In the former case, $\sigma(x,y)\sigma(y,z)=1\cdot(-1)=-1=-\sigma(x,z)$ . In the latter case, $\sigma(x,y)\sigma(y,z)=(-1)^2=1=-\sigma(x,z)$ . Thus, $\sigma$ is compatible with $\theta$ .

Corollary 2.9. The image of the group homomorphism $\widetilde{{\rm{LAut}}}(I(X,K))\to\mathcal{M}(X)$ , $\varphi\mapsto\theta_\varphi$ , coincides with $\mathcal{AM}(X)$ . In particular, $\mathcal{AM}(X)$ is a group.

Proof. If $\varphi\in\widetilde{{\rm{LAut}}}(I(X,K))$ , then $\theta_\varphi\in\mathcal{AM}(X)$ by Lemma 5.4 and Remark 5.10 from [Reference Fornaroli, Khrypchenko and Santulo13]. Let now $\theta\in\mathcal{AM}(X)$ . By Lemma 2.8, there is $\sigma\,{:}\,X^2_<\to K^*$ compatible with $\theta$ . Choose an arbitrary $c=(c_1,\dots,c_n)\in K^n$ with $\sum_{i=1}^n c_i\in K^*$ , where $n=|X|$ . Then $\tau=\tau_{\theta,\sigma,c}\in \widetilde{{\rm{LAut}}}(I(X,K))$ such that $\theta_\tau=\theta$ .

Theorem 2.10. Every Lie automorphism of I(X,K) is proper if and only if $\mathcal{P}(X)=\mathcal{AM}(X)$ .

Proof. Suppose that every Lie automorphism of I(X, K) is proper. Let $\theta\in\mathcal{AM}(X)$ . By Lemma 2.8, there is $\sigma\,{:}\,X^2_<\to K^*$ compatible with $\theta$ and, by [Reference Fornaroli, Khrypchenko and Santulo13, Lemma 5.16], there is $\varphi\in \widetilde{{\rm{LAut}}}(I(X,K))$ inducing $(\theta,\sigma)$ . By hypothesis, $\varphi$ is proper. Therefore, $\theta\in\mathcal{P}(X)$ by Lemma 2.6.

Conversely, suppose that $\mathcal{P}(X)=\mathcal{AM}(X)$ . Let $\varphi\in \widetilde{{\rm{LAut}}}(I(X,K))$ inducing the pair $(\theta,\sigma)$ . By hypothesis, $\theta$ is proper. Therefore, $\varphi$ is proper, by Lemma 2.6. Thus, every Lie automorphism of I(X, K) is proper.

3. Admissible bijections of B and maximal chains in X

Observe that the definitions of the functions $s^{\pm}_{\theta,\Gamma}$ and $t^{\pm}_{\theta,\Gamma}$ make sense for any sequence $\Gamma\,{:}\,u_0,\dots,u_m$ such that either $u_i<u_{i+1}$ or $u_{i+1}<u_i$ for all $0\le i\le m-1$ . We will call such sequences $\Gamma$ semiwalks. If moreover $u_0=u_m$ , then $\Gamma$ will be called a closed semiwalk.

Lemma 3.1. Let $\theta\in\mathcal{M}(X)$ , $z\in X$ and $\Gamma\,{:}\, u_0,u_1,\dots,u_m=u_0$ a closed semiwalk in X. Let $0{\le}k<k{+}l{\le}m$ such that $u_k<u_{k+1}<\dots<u_{k+l}$ or $u_k>u_{k+1}>\dots>u_{k+l}$ and set $\Gamma'\,{:}\,u_0,\dots,u_k,u_{k+l},\dots,u_m=u_0$ . Then

(9) \begin{align} \nonumber\\[-16pt] s^+_{\theta,\Gamma}(z)-t^+_{\theta,\Gamma}(z)=s^+_{\theta,\Gamma'}(z)-t^+_{\theta,\Gamma'}(z),\ s^-_{\theta,\Gamma}(z)-t^-_{\theta,\Gamma}(z)=s^-_{\theta,\Gamma'}(z)-t^-_{\theta,\Gamma'}(z). \end{align}

Proof. Assume that $u_k<u_{k+1}<\dots<u_{k+l}$ . There are two cases.

Case 1. $\theta^{-1}$ is increasing on a maximal chain containing $u_k<u_{k+1}<\dots<u_{k+l}$ . Then there are $v_k<v_{k+1}<\dots<v_{k+l}$ such that $\theta^{-1}(e_{u_iu_j})=e_{v_iv_j}$ for all $k\le i<j\le k+l$ .

Case 1.1. $z=v_i$ for some $k<i<k+l$ . If $\theta(e_{zw})=e_{u_ju_{j+1}}$ for $w>z$ and $k\le j<k+l$ , then $(z,w)=(v_j,v_{j+1})$ , which implies that $j=i$ and $w=v_{i+1}$ . Similarly $\theta(e_{wz})=e_{u_ju_{j+1}}$ for $w<z$ and $k\le j<k+l$ yields $w=v_{i-1}$ . Since, moreover, $\theta(e_{v_kv_{k+l}})=e_{u_ku_{k+l}}$ and $z\ne v_k$ , there is no $w>z$ such that $\theta(e_{zw})=e_{u_ku_{k+l}}$ . Similarly, there is no $w<z$ such that $\theta(e_{wz})=e_{u_ku_{k+l}}$ . Therefore, $s^+_{\theta,\Gamma'}(z)=s^+_{\theta,\Gamma}(z)-1$ , $t^+_{\theta,\Gamma'}(z)=t^+_{\theta,\Gamma}(z)-1$ , $s^-_{\theta,\Gamma'}(z)=s^-_{\theta,\Gamma}(z)$ and $t^-_{\theta,\Gamma'}(z)=t^-_{\theta,\Gamma}(z)$ .

Case 1.2. $z=v_k$ . Again, if $\theta(e_{zw})=e_{u_ju_{j+1}}$ for $w>z$ and $k\le j<k+l$ , then $w=v_{k+1}$ . However, there is no $w<z$ such that $\theta(e_{wz})=e_{u_ju_{j+1}}$ for some $k\le j<k+l$ , but there is a unique $w>z$ (namely, $w=v_{k+l}$ ) such that $\theta(e_{zw})=e_{u_ku_{k+l}}$ . This means that $s^\pm_{\theta,\Gamma'}(z)=s^\pm_{\theta,\Gamma}(z)$ and $t^\pm_{\theta,\Gamma'}(z)=t^\pm_{\theta,\Gamma}(z)$ .

Case 1.3. $z=v_{k+l}$ . This case is similar to Case 1.2. We have $s^\pm_{\theta,\Gamma'}(z)=s^\pm_{\theta,\Gamma}(z)$ and $t^\pm_{\theta,\Gamma'}(z)=t^\pm_{\theta,\Gamma}(z)$ .

Case 1.4. $z\not\in\{v_k,\dots,v_{k+l}\}$ . Then there is neither $w<z$ such that $\theta(e_{wz})=e_{u_ju_{j+1}}$ nor $w>z$ such that $\theta(e_{zw})=e_{u_ju_{j+1}}$ for some $k\le j<k+l$ . Moreover, there is neither $w<z$ such that $\theta(e_{wz})=e_{u_ku_{k+l}}$ nor $w>z$ such that $\theta(e_{zw})=e_{u_ku_{k+l}}$ . Thus, $s^\pm_{\theta,\Gamma'}(z)=s^\pm_{\theta,\Gamma}(z)$ and $t^\pm_{\theta,\Gamma'}(z)=t^\pm_{\theta,\Gamma}(z)$ .

Case 2. $\theta^{-1}$ is decreasing on a maximal chain containing $u_k<u_{k+1}<\dots<u_{k+l}$ . Then everything from Case 1 remains valid with the replacement of the “ $+$ ”-functions by their “minus;”-analogs and vice versa.

In any case, $s^+_{\theta,\Gamma}(z)-t^+_{\theta,\Gamma}(z)$ and $s^-_{\theta,\Gamma}(z)-t^-_{\theta,\Gamma}(z)$ are invariant under the change of $\Gamma$ for $\Gamma'$ . When $u_k>u_{k+1}>\dots>u_{k+l}$ , the proof is analogous.

Corollary 3.2. Let $\theta\in\mathcal{M}(X)$ . Then $\theta\in\mathcal{AM}(X)$ if and only if (2) holds for any $z\in X$ and any closed semiwalk $\Gamma\,{:}\,u_0,\dots,u_m=u_0$ , $m\ge 2$ .

Proof. The “if” part is trivial. Let us prove the “only if” part. Indeed, the case $m=2$ is explained in the proof of [Reference Fornaroli, Khrypchenko and Santulo13, Lemma 5.13], and if $m\ge 3$ , then $\Gamma$ can be extended to a closed walk $\Delta$ by inserting increasing (if $u_i<u_{i+1}$ ) or decreasing (if $u_i>u_{i+1}$ ) sequences of elements between $u_i$ and $u_{i+1}$ for all $0\le i\le m-1$ . Since (2) holds for $\Delta$ , then by Lemma 3.1 it holds for $\Gamma$ too.

Lemma 3.3. Let $\theta\in\mathcal{AM}(X)$ and $\Gamma\,{:}\,u_0,\dots,u_m=u_0$ , $m\ge 2$ , a closed semiwalk. Let also $x_i<y_i$ , such that $\theta(e_{u_iu_{i+1}})=e_{x_iy_i}$ for $u_i<u_{i+1}$ and $\theta(e_{u_{i+1}u_i})=e_{x_iy_i}$ for $u_i>u_{i+1}$ .

  1. i. If $x_i\in {\rm{Min}}(X)$ for some $0\le i\le m-1$ , then there is $j\ne i$ such that $x_i=x_j$ .

  2. ii. If $y_i\in {\rm{Max}}(X)$ for some $0\le i\le m-1$ , then there is $j\ne i$ such that $y_i=y_j$ .

Proof. We will prove (i), the proof of (ii) is analogous. Assume that $x_i\ne x_j$ for all $j\ne i$ . If $u_i<u_{i+1}$ , then $s^+_{\theta^{-1},\Gamma}(x_i)=1$ and $s^-_{\theta^{-1},\Gamma}(x_i)=0$ , since $\theta^{-1}(e_{x_iy_i})=e_{u_iu_{i+1}}$ and $\theta^{-1}(e_{x_iw})\ne e_{u_ju_{j+1}}$ for any $w\ne y_i$ and $j\ne i$ (otherwise $x_i$ would coincide with some $x_j$ for $j\ne i$ ). Similarly, if $u_i>u_{i+1}$ , then $s^+_{\theta^{-1},\Gamma}(x_i)=0$ and $s^-_{\theta^{-1},\Gamma}(x_i)=1$ . Obviously, $t^{\pm}_{\theta^{-1},\Gamma}(x_i)=0$ , because $x_i\in {\rm{Min}}(X)$ . Thus, (2) fails for the triple $(\theta^{-1},\Gamma,x_i)$ , a contradiction.

Lemma 3.4. Let $\theta\in\mathcal{M}(X)$ . Assume that there exist $C,D\in\mathcal{C}(X)$ such that $\theta$ is increasing on C and decreasing on D. If $x\in C\cap D$ , then either $x\in {\rm{Min}}(X)$ or $x\in {\rm{Max}}(X)$ .

Proof. Let $C\,{:}\,x_1<\dots<x_n$ , $D\,{:}\,y_1<\dots<y_m$ and $x=x_i=y_j$ for some $1\le i\le n$ and $1\le j\le m$ . Suppose that $1<i<n$ . Then $1<j<m$ , since otherwise D would not be maximal. There exist maximal chains $C'\,{:}\,u_1<\dots<u_n$ and $D'\,{:}\,v_1<\dots<v_m$ such that $\theta(e_{x_kx_l})=e_{u_ku_l}$ for all $1\le k<l\le n$ and $\theta(e_{y_py_q})=e_{v_{m-q+1}v_{m-p+1}}$ for all $1\le p<q\le m$ . In particular, $\theta(e_{x_{i-1}x_i})=e_{u_{i-1}u_i}$ , $\theta(e_{x_ix_{i+1}})=e_{u_iu_{i+1}}$ , $\theta(e_{y_{j-1}y_j})=e_{v_{m-j+1}v_{m-j+2}}$ , $\theta(e_{y_jy_{j+1}})=e_{v_{m-j}v_{m-j+1}}$ . Observe that $x_{i-1}<x_i=y_j<y_{j+1}$ . Then either $u_i=v_{m-j}$ , or $u_{i-1}=v_{m-j+1}$ , depending on whether $\theta$ is increasing or decreasing on a maximal chain containing $x_{i-1}<x_i<y_{j+1}$ . Similarly, considering $y_{j-1}<y_j=x_i<x_{i+1}$ we obtain $v_{m-j+1}=u_{i+1}$ or $v_{m-j+2}=u_i$ . If $u_i=v_{m-j}$ , then $v_{m-j+1}=u_{i+1}$ , so that $\{u_i,u_{i+1}\}\subseteq C'\cap D'$ . However, $\theta^{-1}$ is increasing on C $^{\prime}$ and decreasing on D $^{\prime}$ , so $u_i$ is the common minimum of C $^{\prime}$ and D $^{\prime}$ and $u_{i+1}$ is the common maximum of C $^{\prime}$ and D $^{\prime}$ by Lemma 2.7. This contradicts the assumption $1<i<n$ . Similarly, $u_{i-1}=v_{m-j+1}$ implies $v_{m-j+2}=u_i$ , whence $\{u_{i-1},u_i\}\subseteq C'\cap D'$ leading to a contradiction.

Thus, $i\in\{1,n\}$ . If $i=1$ , then necessarily $j=1$ , as otherwise C would not be maximal. Similarly, if $i=n$ , then $j=m$ .

Lemma 3.5. Let $\theta\in\mathcal{M}(X)$ and $C,D\in\mathcal{C}(X)$ , $C\,{:}\,x_1<\dots<x_n$ , $D\,{:}\,y_1<\dots<y_m$ . Assume that $x_i=y_j$ for some $1<i<n$ and $1<j<m$ . If $\theta$ is increasing (resp. decreasing) on C, then it is increasing (resp. decreasing) on D. Moreover, if $\theta(C)\,{:}\,u_1<\dots<u_n$ and $\theta(D)\,{:}\,v_1<\dots<v_m$ , then $u_i=v_j$ (resp. $u_{n-i+1}=v_{m-j+1}$ ).

Proof. Let $\theta$ be increasing on C. Then it is increasing on D by Lemma 3.4. Using the same idea as in the proof of Lemma 3.4, we have $\theta(e_{x_{i-1}x_i})=e_{u_{i-1}u_i}$ , $\theta(e_{x_ix_{i+1}})=e_{u_iu_{i+1}}$ , $\theta(e_{y_{j-1}y_j})=e_{v_{j-1}v_j}$ , $\theta(e_{y_jy_{j+1}})=e_{v_jv_{j+1}}$ . Considering $x_{i-1}<x_i=y_j<y_{j+1}$ we conclude that $u_i=v_j$ or $u_{i-1}=v_{j+1}$ . Similarly it follows from $y_{j-1}<y_j=x_i<x_{i+1}$ that $u_i=v_j$ or $v_{j-1}=u_{i+1}$ . If $u_i\ne v_j$ , then $u_{i-1}=v_{j+1}$ and $v_{j-1}=u_{i+1}$ . But this is impossible, since $u_{i-1}<u_{i+1}$ and $v_{j+1}>v_{j-1}$ .

The proof for the decreasing case is analogous.

Definition 3.6. Let $C,D\in\mathcal{C}(X)$ . We say that C and D are linkedif there exists $x\in C\cap D$ such that $x\not\in {\rm{Min}}(X)\sqcup {\rm{Max}}(X)$ . Denote by $\sim$ the equivalence relation on $\mathcal{C}(X)$ generated by $\{(C,D)\in\mathcal{C}(X)^2 \,{:}\, C,D\text{ are linked}\}$ .

Lemma 3.7. Each $\theta\in\mathcal{M}(X)$ induces a bijection $\widetilde\theta$ on $\mathcal{C}(X)/{\sim}$ . Moreover, if $\theta$ is increasing (resp. decreasing) on $C\in\mathcal{C}(X)$ , then it is increasing (resp. decreasing) on any $D\sim C$ .

Proof. Let $C,D\in\mathcal{C}(X)$ be linked. Then $\theta(C)$ and $\theta(D)$ are linked by Lemma 3.5. It follows that $C\sim D$ implies $\theta(C)\sim\theta(D)$ , which induces a map $\widetilde\theta\,{:}\,\mathcal{C}(X)/{\sim}\to\mathcal{C}(X)/{\sim}$ . It is a bijection whose inverseis $\widetilde{\theta^{-1}}$ .

Assume that $\theta$ is increasing on $C\in\mathcal{C}(X)$ . Then by Lemma 3.5 it is increasing on any $D\in\mathcal{C}(X)$ which is linked to C. By the obvious induction, this extends to any $D\sim C$ . The decreasing caseis similar.

Definition 3.8. Given $\mathfrak{C}\in\mathcal{C}(X)/{\sim}$ , we define the supportof $\mathfrak{C}$ , denoted ${\rm{supp}}(\mathfrak{C})$ , as the set $\{x{\in}C \,{:}\, C{\in}\mathfrak{C}\}$ .

Remark 3.9. Let $\mathfrak{C},\mathfrak{D}\in\mathcal{C}(X)/{\sim}$ . If $\mathfrak{C}\ne\mathfrak{D}$ , then ${\rm{supp}}(\mathfrak{C})\cap {\rm{supp}}(\mathfrak{D}) \subseteq{\rm{Min}}(X)\sqcup {\rm{Max}}(X)$ .

Indeed, assume that $x\in {\rm{supp}}(\mathfrak{C})\cap {\rm{supp}}(\mathfrak{D})$ , where $x\not\in {\rm{Min}}(X)$ and $x\not\in {\rm{Max}}(X)$ . There are $C\in\mathfrak{C}$ and $D\in\mathfrak{D}$ such that $x\in C\cap D$ . But then C and D are linked, so $C\sim D$ , whence $\mathfrak{C}=\mathfrak{D}$ .

Theorem 3.10. Let $\theta\in\mathcal{AM}(X)$ and $\mathfrak{C}\in\mathcal{C}(X)/{\sim}$ . Then there exists an isomorphism or an anti-isomorphism of posets $\lambda\,{:}\,{\rm{supp}}(\mathfrak{C})\to {\rm{supp}}(\widetilde\theta(\mathfrak{C}))$ such that for all $x<y$ from ${\rm{supp}}(\mathfrak{C})$ one has

(10) \begin{align} \theta(e_{xy})=\hat\lambda(e_{xy}). \end{align}

Proof. In view of Lemma 3.7, we may assume that $\theta$ is increasing on all $C\in\mathfrak{C}$ or decreasing on all $C\in\mathfrak{C}$ . Consider the case of an increasing $\theta$ . We are going to construct the corresponding $\lambda\,{:}\,{\rm{supp}}(\mathfrak{C})\to {\rm{supp}}(\widetilde\theta(\mathfrak{C}))$ . Let $x\in {\rm{supp}}(\mathfrak{C})$ and $C\,{:}\,x_1<\dots<x_n$ a maximal chain from $\mathfrak{C}$ containing x. Denote by $C'\,{:}\,u_1<\dots<u_n$ the image of C under $\theta$ . If $x=x_i$ for some $1\le i\le n$ , then we put $\lambda(x_i)=u_i$ . We still need to show that the definition does not depend on the choice of C.

If $1<i<n$ , then this is true by Lemma 3.5.

If $i=1$ , then $x\in {\rm{Min}}(X)$ . If there exists another maximal chain $D\,{:}\,y_1<\dots<y_m$ from $\mathfrak{C}$ containing x, then $x=y_1$ . We thus need to show that $u_1=v_1$ , where $D'\,{:}\,v_1<\dots<v_m$ is the image of D under $\theta$ . Since $C\sim D$ , there are $C=C_1,\dots,C_k=D$ such that $C_j$ and $C_{j+1}$ are linked for all $1\le j\le k-1$ . Denote by $z_j$ an element of $C_j\cap C_{j+1}$ , $1\le j\le k-1$ , which is neither minimal nor maximal in X. Set also $z_0=z_k=x$ . Observe that $z_j,z_{j+1}\in C_{j+1}$ for all $0\le j\le k-1$ , so that either $z_j\le z_{j+1}$ or $z_j\ge z_{j+1}$ . Let $\Gamma\,{:}\,z_0,z_1,\dots,z_k=z_0$ . Clearly, $z_0\ne z_j$ and $z_j\ne z_k$ for all $1\le j\le k-1$ , as $z_0=z_k\in {\rm{Min}}(X)$ , while $z_j\not\in {\rm{Min}}(X)$ . Moreover, we will assume that $z_j\ne z_{j+1}$ for all $1\le j\le k-2$ , since otherwise we may just remove the repetitions (and at least 2 elements will remain). Let also $a_j<b_j$ such that $\theta(e_{z_jz_{j+1}})=e_{a_jb_j}$ if $z_j<z_{j+1}$ , and $\theta(e_{z_{j+1}z_j})=e_{a_jb_j}$ if $z_j>z_{j+1}$ , $0\le j\le k-1$ . Observe that $a_0=u_1$ , since $x=z_0<z_1\in C$ , and $a_{k-1}=v_1$ , since $x=z_k<z_{k-1}\in D$ . In particular, $a_0,a_{k-1}\in {\rm{Min}}(X)$ . Since $z_j$ is not minimal for all $1\le j\le k-2$ , then neither is $a_j$ , so that $a_j\not\in\{a_0,a_{k-1}\}$ for such j. But then we must have $a_0=a_{k-1}$ , that is, $u_1=v_1$ , by Lemma 3.3 (i).

The case $i=n$ is similar. The map $\lambda\,{:}\,{\rm{supp}}(\mathfrak{C})\to {\rm{supp}}(\widetilde\theta(\mathfrak{C}))$ is thus constructed.

We now prove that $\lambda(x)<\lambda(y)$ and (10) holds for all $x<y$ from ${\rm{supp}}(\mathfrak{C})$ . By construction, this is true for x and y belonging to the same $C\in\mathfrak{C}$ . Let now $x<y$ be arbitrary elements of ${\rm{supp}}(\mathfrak{C})$ . Choose $C\in\mathfrak{C}$ containing x and $D\in\mathfrak{C}$ containing y. If $x\not\in {\rm{Min}}(X)$ , then any $C'\in\mathcal{C}(X)$ containing x and y is linked to C, so that $C'\in\mathfrak{C}$ . The case when $y\not\in {\rm{Max}}(X)$ is similar. Let now $x\in {\rm{Min}}(X)$ and $y\in {\rm{Max}}(X)$ . As above, we choose $C=C_1,\dots,C_k=D$ , such that $C_j,C_{j+1}\in\mathfrak{C}$ are linked for all $1\le j\le k-1$ , and $z_j\in C_j\cap C_{j+1}$ , $1\le j\le k-1$ , which is neither minimal nor maximal in X. We set $z_0=x$ , $z_k=y$ and $\Gamma\,{:}\,z_0,z_1,\dots,z_k,z_{k+1}=z_0$ . We also assume that $z_j\ne z_{j+1}$ for all $0\le j\le k$ and denote by $a_j<b_j$ the elements satisfying $\theta(e_{z_jz_{j+1}})=e_{a_jb_j}$ if $z_j<z_{j+1}$ and $\theta(e_{z_{j+1}z_j})=e_{a_jb_j}$ if $z_j>z_{j+1}$ , $0\le j\le k$ . As above, observe that $a_0,a_k\in {\rm{Min}}(X)$ , while $a_1,\dots,a_{k-1}\not\in {\rm{Min}}(X)$ . Then $a_0=a_k$ by Lemma 3.3 (i). Similarly it follows from Lemma 3.3 (ii) that $b_{k-1}=b_k$ . But $a_0=\lambda(x)$ and $b_{k-1}=\lambda(y)$ , since $e_{a_0b_0}=\theta(e_{z_0z_1})=e_{\lambda(z_0)\lambda(z_1)}=e_{\lambda(x)\lambda(z_1)}$ and $e_{a_{k-1}b_{k-1}}=\theta(e_{z_{k-1}z_k})=e_{\lambda(z_{k-1})\lambda(z_k)}=e_{\lambda(z_{k-1})\lambda(y)}$ . Hence, $\lambda(x)=a_k<b_k=\lambda(y)$ and $\theta(e_{xy})=\theta(e_{z_{k+1}z_k})=e_{a_kb_k}=e_{\lambda(x)\lambda(y)}=\hat\lambda(e_{xy})$ .

It is clear that $\lambda$ is a bijection whose inverse is the map $\mu\,{:}\,{\rm{supp}}(\widetilde\theta(\mathfrak{C}))\to {\rm{supp}}(\mathfrak{C})$ corresponding to $\theta^{-1}$ . Thus, $\lambda$ is an isomorphism between ${\rm{supp}}(\mathfrak{C})$ and ${\rm{supp}}(\widetilde\theta(\mathfrak{C}))$ .

The case of a decreasing $\theta$ is analogous.

The following example shows that the admissibility of $\theta$ in Theorem 3.10 cannot be dropped.

Example 3.11. Let $X=\{1,\dots,10,1',\dots 9',7''\}$ with the following Hasse diagram.

Then $\mathcal{C}(X)/{\sim}$ consists of 2 classes whose supports are $Y=\{1,\dots,10\}$ and $Z=\{1',\dots,9',10,7''\}$ . Observe that there exists $\theta\in\mathcal{M}(X)$ mapping one ${\sim}$ -class to another. It is defined as follows: $\theta(e_{ij})=e_{i'j'}$ for all $i\le j$ in X with $(i,j)\ne (5,7)$ , $\theta(e_{i'j'})=e_{ij}$ for all $i'\le j'$ in X with $(i',j')\ne(5',7'')$ , $\theta(e_{57})=e_{5'7''}$ and $\theta(e_{5'7''})=e_{57}$ (to make the definition shorter, we set $10':=10$ ). However, Y and Z are not isomorphic or anti-isomorphic because $|Y|\ne|Z|$ . The reason is that $\theta\not\in\mathcal{AM}(X)$ . Indeed, for $\Gamma\,{:}\,5<7>6<8>5$ we have $s^\pm_{\theta,\Gamma}(7')=0$ , $t^+_{\theta,\Gamma}(7')=0$ and $t^-_{\theta,\Gamma}(7')=1$ .

As a consequence of Theorems 2.10 and 3.10, we have the following result which generalizes [Reference Fornaroli, Khrypchenko and Santulo13, Corollary 5.19], where X was a chain.

Corollary 3.12. If $|\mathcal{C}(X)/{\sim}|=1$ , then each $\varphi\in {\rm{LAut}}(I(X,K))$ is proper.

Observe, however, that the condition $|\mathcal{C}(X)/{\sim}|=1$ is not necessary for all $\varphi\in {\rm{LAut}}(I(X,K))$ to be proper, as the following example shows.

Example 3.13. Let $X=\{1,2,3,4,5,6\}$ with the following Hasse diagram.

Note that $\mathcal{C}(X)/{\sim}$ consists of 2 classes whose supports are $Y=\{1,2,4,5\}$ and $Z=\{1,3,5,6\}$ . For any $\theta\in\mathcal{AM}(X)$ , there are 2 possibilities for the corresponding isomorphisms $\lambda_1$ and $\lambda_2$ between the supports: either $\lambda_1\,{:}\,Y\to Y$ and $\lambda_2\,{:}\,Z\to Z$ , or $\lambda_1\,{:}\,Y\to Z$ and $\lambda_2\,{:}\,Z\to Y$ . In the former case $\lambda_1={\rm{id}}_Y$ and $\lambda_2={\rm{id}}_Z$ , and in the letter case $\lambda_2=\lambda^{-1}_1$ , where $\lambda_1$ maps an element $y\in Y$ to the element $z\in Z$ which is symmetric to y with respect to the vertical line passing through the vertices 1 and 5. In both cases, $\lambda_1$ and $\lambda_2$ are the restrictions of an automorphism of X to Y and Z, respectively.

4. Sets of length one

4.1. Admissible bijections of B and crowns in X

Before proceeding to the case $l(X)=1$ , we will prove a useful fact which holds for X of an arbitrary length.

Definition 4.1. Let n be an integer greater than 1. By a weak n-crownwe mean a poset $P=\{x_1,\dots,x_{n},y_1,\dots,y_{n}\}$ where

(11) \begin{align} x_i<y_i\ {for\ all}\ 1\le i\le n,\ x_{i+1}<y_i\ {for\ all}\ 1\le i\le n-1\ {and}\ x_1<y_{n}.\end{align}

An n-crownis a weak n-crown which has no other pairs of distinct comparable elements except (11). It is thus fully determined by n up to an isomorphism and will be denoted by ${\rm{Cr}}_n$ . A poset P is called a weak crown(resp. crown), if it is a weak n-crown (resp. n-crown) for some $n\ge 2$ . We say that a poset X has a weak crown (resp. crown) if there is a subset $Y\subseteq X$ which is a weak crown (resp. crown) under the induced partial order.

Posets without crowns are known to satisfy some “good” properties [Reference Dokuchaev and Novikov10, Reference Dräxler11, Reference Rival19].

Lemma 4.2. Let $\theta\in\mathcal{M}(X)$ . Then $\theta\in\mathcal{AM}(X)$ if and only if (2) holds for any $z\in X$ and any weak crown $\Gamma\,{:}\,u_0,\dots,u_m=u_0$ in X.

Proof. The “only if” part is obvious. For the “if” part take any closed semiwalk $\Gamma\,{:}\,u_0,u_1,\dots,u_m=u_0$ and $z\in X$ . Let $0\le k<k+l\le m$ such that $u_k<u_{k+1}<\dots<u_{k+l}$ . We define $\Gamma'\,{:}\,u_0,\dots,u_k,u_{k+l},\dots,u_m=u_0$ . By Corollary 3.2 equality (2) holds for $\Gamma$ if and only if it holds for $\Gamma'$ . The same is true for any $\Gamma'$ obtained from $\Gamma$ by removing intermediate terms in a decreasing sequence of consecutive vertices. Thus, doing so for all maximal sequences in $\Gamma$ we finally get $\Gamma'$ whose vertices form either a sequence $x<y>z$ or a weak crown. However, the case $\Gamma'\,{:}\,x<y>z$ can be ignored, because (2) always holds for such $\Gamma'$ as shown in the proof of [Reference Fornaroli, Khrypchenko and Santulo13, Lemma 5.13].

4.2. The crownless case

Let now $l(X)=1$ . Observe that $\mathcal{M}(X)=S(B)$ . Moreover, any $C\in\mathcal{C}(X)$ is linked only to itself, so $|\mathcal{C}(X)/{\sim}|=|\mathcal{C}(X)|$ and Theorem 3.10 becomes useless.

Definition 4.3. We say that $\theta\in\mathcal{M}(X)$ is separatingif there exists a pair of non-disjoint $C,D\in\mathcal{C}(X)$ such that $\theta(C)$ and $\theta(D)$ are disjoint.

Remark 4.4. Any separating $\theta$ is not proper.

Lemma 4.5. Let $l(X)=1$ . If $|{\rm{Min}}(X)|>1$ and $|{\rm{Max}}(X)|>1$ , then there are disjoint $C,D\in\mathcal{C}(X)$ .

Proof. Choose arbitrary $x<y$ in X. Obviously, $x\in {\rm{Min}}(X)$ and $y\in {\rm{Max}}(X)$ . Let $U=\{u\in {\rm{Min}}(X)\mid u\not\le y\}$ and $V=\{v\in {\rm{Max}}(X)\mid x\not\le v\}$ . If $U\ne\emptyset$ , then take $u\in U$ . Clearly, $u\ne x$ . Since X is connected, there exists $v>u$ , and $v\ne y$ by the definition of U. Then $C\,{:}\,x<y$ and $D\,{:}\,u<v$ are disjoint. The case $V\ne\emptyset$ is similar. Suppose now that $U=V=\emptyset$ . This means that $x\le v$ for any $v\in {\rm{Max}}(X)$ and $y\ge u$ for any $u\in {\rm{Min}}(X)$ . Choose $u\in {\rm{Min}}(X)\setminus\{x\}$ and $v\in {\rm{Max}}(X)\setminus\{y\}$ . Then $C_1\,{:}\,x<v$ and $D_1\,{:}\,u<y$ aredisjoint.

Proposition 4.6. Let $l(X)=1$ . Then $\mathcal{M}(X)=\mathcal{P}(X)$ if and only if $|{\rm{Min}}(X)|=1$ or $|{\rm{Max}}(X)|=1$ .

Proof. If $|{\rm{Min}}(X)|=1$ , say ${\rm{Min}}(X)=\{x\}$ , then any $\theta\in\mathcal{M}(X)$ can be identified with a bijection $\lambda$ of Max(X) such that $\theta(e_{xy})=e_{x\lambda(y)}$ for all $y>x$ . But $\lambda$ extends to an automorphism of X by means of $\lambda(x)=x$ , so that $\theta(e_{xy})=e_{\lambda(x)\lambda(y)}$ . A symmetric argument works in the case $|{\rm{Max}}(X)|=1$ .

Suppose now that $|{\rm{Min}}(X)|>1$ and $|{\rm{Max}}(X)|>1$ . By Lemma 4.5 there are disjoint $C\,{:}\,x<y$ and $D\,{:}\,u<v$ . Choose a path $x=x_0,x_1,\dots,x_m=u$ . Since $x,u\in {\rm{Min}}(X)$ , then $m\ge 2$ and $x_0<x_1>x_2$ . We define $\theta(e_{x_0x_1})=e_{xy}$ , $\theta(e_{xy})=e_{x_0x_1}$ , $\theta(e_{x_2x_1})=e_{uv}$ , $\theta(e_{uv})=e_{x_2x_1}$ and $\theta(e_{ab})=e_{ab}$ for any other $e_{ab}\in B$ . Clearly, $\theta\in S(B)=\mathcal{M}(X)$ and it is separating, in particular, not proper.

If X does not contain a weak crown, then $\mathcal{AM}(X)=\mathcal{M}(X)$ by Lemma 4.2. Hence, we obtain the following.

Corollary 4.7. Let $l(X)=1$ and assume that X does not contain a weak crown. Then $\mathcal{AM}(X)=\mathcal{P}(X)$ if and only if $|{\rm{Min}}(X)|=1$ or $|{\rm{Max}}(X)|=1$ .

4.3. The crown case

We will now consider two classes of posets of length one which have crowns. We begin with the case of X being a crown and are going to calculate the groups $\mathcal{P}(X)$ and $\mathcal{AM}(X)$ explicitly. Thus, in this subsection $X={\rm{Cr}}_n=\{x_1,\dots,x_n,y_1,\dots,y_n\}$ .

Definition 4.8. The chains $x_i<y_i$ , $1\le i\le n$ , will be called odd, and $x_{i+1}<y_i$ , $1\le i\le n-1$ , and $x_1<y_{n}$ will be called even. Thus, each element of ${\rm{Cr}}_n$ belongs to exactly one odd chain and to exactly one even chain.

Lemma 4.9. Let $\theta\in\mathcal{M}({\rm{Cr}}_n)$ . Then $\theta\in\mathcal{AM}({\rm{Cr}}_n)$ if and only if for any pair of distinct non-disjoint chains $C,D\in\mathcal{C}({\rm{Cr}}_n)$ , the images $\theta(C)$ and $\theta(D)$ have opposite parities.

Proof. Observe that (2) is invariant under cyclic shifts of $\Gamma$ (any such shift does not change the functions $s^{\pm}_{\theta,\Gamma}$ and $t^{\pm}_{\theta,\Gamma}$ ). Thus, for admissibility it is enough to consider $\Gamma\,{:}\,x_1<y_1>x_2<\dots <y_{n}>x_1$ , since any other cycle in ${\rm{Cr}}_n$ is a cyclic shift of $\Gamma$ .

The “only if” case. Let $\theta\in\mathcal{AM}({\rm{Cr}}_n)$ and $C,D\in\mathcal{C}({\rm{Cr}}_n)$ , such that $C\cap D=\{z\}$ , where $z\in {\rm{Min}}({\rm{Cr}}_n)\sqcup {\rm{Max}}({\rm{Cr}}_n)$ . Suppose that $z\in {\rm{Min}}({\rm{Cr}}_n)$ . Then there are only two elements $w,w'\in {\rm{Max}}({\rm{Cr}}_n)$ such that $z<w,w'$ . Thus, $t_{\theta,\Gamma}^{\pm}(z)=0$ and $\theta$ is admissible if and only if $s_{\theta,\Gamma}^{+}(z)=s_{\theta,\Gamma}^{-}(z)=1$ , which only occur if $\theta(C)$ and $\theta(D)$ have opposite parities. The case when $z\in {\rm{Max}}({\rm{Cr}}_n)$ is similar.

The “if” case. Let $\theta\in\mathcal{M}({\rm{Cr}}_n)$ and $z\in {\rm{Cr}}_n$ be arbitrary. Again, we consider the case $z\in {\rm{Min}}({\rm{Cr}}_n)$ , so that $t_{\theta,\Gamma}^{\pm}(z)=0$ . Choose $w,w'\in {\rm{Max}}({\rm{Cr}}_n)$ with $z<w,w'$ and put $C\,{:}\,z<w$ and $D\,{:}\,z<w'$ . Since $\theta(C)$ and $\theta(D)$ have opposite parities, then $s_{\theta,\Gamma}^{\pm}(z)=1$ and (2) is satisfied. Similarly, one handles the case $z\in {\rm{Max}}({\rm{Cr}}_n)$ .

Proposition 4.10. The group $\mathcal{AM}({\rm{Cr}}_n)$ is isomorphic to $(S_{n}\times S_{n})\rtimes \mathbb{Z}_2$ .

Proof. Denote by $\mathcal{O}$ and $\mathcal{E}$ the subsets of $\mathcal{C}({\rm{Cr}}_n)$ formed by the odd and even chains, respectively, and let $\mathcal{G}=\{\theta\in \mathcal{M}({\rm{Cr}}_n) \,{:}\, \theta(\mathcal{O})=\mathcal{O} \text{ or } \theta(\mathcal{O})=\mathcal{E}\}$ . We will first prove that $\mathcal{AM}({\rm{Cr}}_n)=\mathcal{G}$ . For any $\theta\in \mathcal{AM}({\rm{Cr}}_n)$ , if $\theta(e_{x_1y_1})\in \mathcal{O}$ , then $\theta(e_{x_2y_1})\in \mathcal{E}$ by Lemma 4.9. It follows that $\theta(e_{x_2y_2})\in \mathcal{O}$ by the same reason. Applying this argument consecutively to $e_{x_3y_2},e_{x_3y_3},\dots,e_{x_ky_k},e_{x_1y_k}$ , we obtain $\theta(\mathcal{O})=\mathcal{O}$ . Similarly, if $\theta(e_{x_1y_1})\in \mathcal{E}$ , then $\theta(\mathcal{O})=\mathcal{E}$ . Thus, $\theta\in \mathcal{G}$ . On the other hand, let $\theta\in \mathcal{G}$ and $C_1,C_2\in\mathcal{C}({\rm{Cr}}_n)$ , $C_1\neq C_2$ , such that $C_1\cap C_2\neq\emptyset$ . Then $C_1$ and $C_2$ have opposite parities, say, $C_1\in \mathcal{O}$ and $C_2\in \mathcal{E}$ . If $\theta(\mathcal{O})=\mathcal{O}$ , then $\theta(\mathcal{E})=\mathcal{E}$ due to the bijectivity of $\theta$ . Analogously, if $\theta(\mathcal{O})=\mathcal{E}$ , then $\theta(\mathcal{E})=\mathcal{O}$ . So, in either case, $\theta(C_1)$ and $\theta(C_2)$ have opposite parities. Therefore, $\theta\in \mathcal{AM}({\rm{Cr}}_n)$ , by Lemma 4.9.

We now prove that $\mathcal{G}\cong (S_{n}\times S_{n})\rtimes \mathbb{Z}_2$ . Consider $\mathcal{H}=\{\theta\in \mathcal{M}({\rm{Cr}}_n) \,{:}\, \theta(\mathcal{O})=\mathcal{O}\}$ . Clearly, $\mathcal{H}$ is a (normal) subgroup of $\mathcal{G}$ of index 2. Since $|\mathcal{O}|=|\mathcal{E}|=n$ , we have $\mathcal{H}\cong S_{n}\times S_{n}$ . Define $\theta\in\mathcal{M}({\rm{Cr}}_n)$ as follows: $\theta(e_{x_iy_i})=e_{x_{i+1}y_i}$ and $\theta(e_{x_{i+1}y_i})=e_{x_iy_i}$ for $1\le i\le n-1$ , $\theta(e_{x_{n}y_{n}})=e_{x_1y_{n}}$ and $\theta(e_{x_1y_{n}})=e_{x_{n}y_{n}}$ . By definition $\theta\in\mathcal{G}\setminus\mathcal{H}$ and $\theta$ has order 2. Therefore, $\mathcal{G}=\mathcal{H}\cdot\langle\theta\rangle\cong(S_{n}\times S_{n})\rtimes \mathbb{Z}_2$ .

Proposition 4.11. The group $\mathcal{P}({\rm{Cr}}_n)$ is isomorphic to $\mathbb{Z}_{2n}\rtimes \mathbb{Z}_2$ .

Proof. In view of Proposition 2.3, it suffices to prove that ${\rm{Aut}}^\pm({\rm{Cr}}_n)\cong \mathbb{Z}_{2n}\rtimes \mathbb{Z}_2$ . To this end, we will show that ${\rm{Aut}}^\pm({\rm{Cr}}_n)\cong D_{2n}$ , where $D_{2n}$ is the group of symmetries of a regular 2n-gon which is known to be isomorphic to $\mathbb{Z}_{2n}\rtimes \mathbb{Z}_2$ . Denote $x_i$ by $u_{2i-1}$ and $y_i$ by $u_{2i}$ , for all $i=1,\ldots,n$ , identifying $u_j$ with the j-th vertex of a regular 2n-gon, whose vertices are indexed consecutively according to the counterclockwise orientation. For the sake of simplicity, we shall consider the indices modulo 2n in the rest of the proof.

Given $\varphi\in {\rm{Aut}}^\pm({\rm{Cr}}_n)$ , set $i_{\varphi}$ to be the integer modulo 2n such that $\varphi(u_{2n})=u_{i_{\varphi}}$ . Notice that if $i_{\varphi}$ is even then $\varphi\in {\rm{Aut}}({\rm{Cr}}_n)$ , otherwise $\varphi\in {\rm{Aut}}^-({\rm{Cr}}_n)$ . In any case, since the only elements of ${\rm{Cr}}_n$ comparable with $u_{2n}$ , besides itself, are $u_{2n-1}$ and $u_1$ , then $\varphi(u_1)=u_{i_{\varphi}\pm 1}$ . If $\varphi(u_1)=u_{i_{\varphi}+1}$ , then it can be easily shown inductively that $\varphi(u_j)=u_{i_{\varphi}+j}$ for any $j=1,\ldots,2n$ . This corresponds to the counterclockwise rotation by an angle of $i_{\varphi}\pi/n$ in $D_{2n}$ . If $\varphi(u_1)=u_{i_{\varphi}-1}$ , again by an easy inductive argument, $\varphi(u_j)=u_{i_{\varphi}-j}$ for all $j=1,\ldots,2n$ . If $i_{\varphi}$ is even, $\varphi$ corresponds to the reflection across the diagonal containing $u_j$ and $u_{j+n}$ , where $2j=i_{\varphi}$ . Otherwise $\varphi$ corresponds to the reflection across the line which contains the midpoints of the sides $u_ju_{j+1}$ and $u_{j+n}u_{j+n+1}$ , where $i_{\varphi}=2j+1$ . Since $i_{\varphi}$ can be any of the 2n indices of the vertices considered, all the 4n elements of $D_{2n}$ (2n rotations and 2n reflections) can occur as elements of ${\rm{Aut}}^\pm({\rm{Cr}}_n)$ and we obtain the claimed isomorphism.

Corollary 4.12. We have $\mathcal{P}({\rm{Cr}}_2)=\mathcal{AM}({\rm{Cr}}_2)$ and $\mathcal{P}({\rm{Cr}}_n)\ne \mathcal{AM}({\rm{Cr}}_n)$ for all $n>2$ .

Proof. Indeed, $|\mathcal{AM}({\rm{Cr}}_2)|=|\mathcal{P}({\rm{Cr}}_2)|$ and $|\mathcal{AM}({\rm{Cr}}_n)|=2(n!)^2>4n!>4n=|\mathcal{P}({\rm{Cr}}_2)|$ for $n>2$ .

4.4. The case of the ordinal sum of two anti-chains

We will now proceed to the case of sets of length one which have as many crowns as possible.

Definition 4.13. Given positive integers mand n, denote by ${\rm{K}}_{m,n}$ the poset $\{x_1,\dots,x_m,y_1,\dots,y_n\}$ , where $x_i<y_j$ for all $1\le i\le m$ and $1\le j\le n$ , and there is no other pair of distinct comparable elements.

Observe that ${\rm{K}}_{m,n}$ is the ordinal sum [Reference Stanley24] of two anti-chains of cardinalities m and n. The Hasse diagram of ${\rm{K}}_{m,n}$ is a complete bipartite graph [Reference Bollobás2], so that ${\rm{Aut}}({\rm{K}}_{m,n})\cong S_m\times S_n$ . It is also clear that ${\rm{K}}_{m,n}$ is anti-isomorphic to ${\rm{K}}_{n,m}$ , so we may assume that $m\le n$ . The cases $m=1$ and $m=n=2$ (a 2-crown) were considered in Proposition 4.6 and Corollary 4.12.

Proposition 4.14. Let $2\le m\le n$ . Then $\mathcal{P}({\rm{K}}_{m,n})=\mathcal{AM}({\rm{K}}_{m,n})$ .

Proof. Let $\theta\in\mathcal{AM}({\rm{K}}_{m,n})$ . Fix $j\in \{1,\ldots, n\}$ and write $\theta(e_{x_iy_j})=e_{u_iv_i}$ , $1\le i\le m$ . Denote by $U_j$ and $V_j$ the sets of all $u_i$ and $v_k$ , respectively. We first prove that for any pair of $u_i\in U_j$ and $v_k\in V_j$ there is l such that $\theta(e_{x_ly_j})=e_{u_iv_k}$ . This is trivial if $u_i=u_k$ or $v_i=v_k$ , so let $u_i\ne u_k$ and $v_i\ne v_k$ . Consider the cycle $\Gamma\,{:}\,u_i<v:i>u_k<v:k>u_i$ . We have $s^{\pm}_{\theta,\Gamma}(y_j)=0$ and $t^+_{\theta,\Gamma}(y_j)=2$ . Since $\theta$ is admissible, we must have $t^-_{\theta,\Gamma}(y_j)=2$ . But this means that $\theta(e_{x_ly_j})=e_{u_iv_k}$ for some $1\le l\le m$ (and $\theta(e_{x_py_j})=e_{u_kv_i}$ for some $1\le p\le m$ ), as desired. As a consequence, we obtain $|U_j|\cdot|V_j|=m$ .

We now prove that $|U_j|=1$ or $|V_j|=1$ . Assume that $|U_j|\ge 2$ and $|V_j|\ge 2$ . Since $|U_j|\cdot|V_j|=m$ , we conclude that $|U_j|\le \frac m2$ and $|V_j|\le \frac m2\le\frac n2$ . It follows that there exist $z<w$ such that $z\not\in U_j$ and $w\not\in V_j$ . Consider the cycle $\Gamma\,{:}\,u_1<v_1>zu_1$ . Clearly, $s^{\pm}_{\theta,\Gamma}(y_j)=0$ , $t^+_{\theta,\Gamma}(y_j)=1$ and $t^-_{\theta,\Gamma}(y_j)=0$ , a contradiction.

Case 1. $|V_j|=1$ for all $1\le j\le n$ . Then there exists a bijection $\lambda$ of $\{y_1,\dots,y_n\}$ such that $\{\theta(e_{x_1y_j}),\dots,\theta(e_{x_my_j})\}=\{e_{x_1\lambda(y_j)},\dots,e_{x_m\lambda(y_j)}\}$ for all $1\le j\le n$ . We will prove that $\theta(e_{x_iy_j})=e_{u_i\lambda(y_j)}$ and $\theta(e_{x_iy_k})=e_{z_i\lambda(y_k)}$ imply $u_i=z_i$ for $j\ne k$ . Suppose that $u_i\ne z_i$ and consider the cycle $\Gamma\,{:}\,u_i<\lambda(y_j)>z_i<\lambda(y_k)>u_i$ . We have $t^{\pm}_{\theta,\Gamma}(x_i)=0$ , $s^+_{\theta,\Gamma}(x_i)=2$ and $s^-_{\theta,\Gamma}(x_i)=0$ , a contradiction. Thus, there exists a bijection $\mu$ of $\{x_1,\dots,x_m\}$ such that $\theta(e_{x_iy_j})=e_{\mu(x_i)\lambda(y_j)}$ for all $1\le i\le m$ and $1\le j\le n$ . But this means that $\theta$ corresponds to the automorphism of ${\rm{K}}_{m,n}$ acting as $\lambda$ on ${\rm{Max}}({\rm{K}}_{m,n})$ and as $\mu$ on ${\rm{Min}}({\rm{K}}_{m,n})$ . So, $\theta\in\mathcal{P}({\rm{K}}_{m,n})$ .

Case 2. $|U_j|=1$ for some $1\le j\le n$ . Then $|V_j|=m$ . We will prove that this is possible only if $m=n$ . Assume that $m<n$ and let $U_j=\{p_j\}$ . Since $\theta(e_{x_iy_j})=e_{p_jv_i}$ for all $1\le i\le m$ and $|V_j|<n$ , there exist $z<w$ such that $z\ne p_j$ and $w\not\in V_j$ . Taking the cycle $\Gamma\,{:}\,p_j<v_1>zp_j$ , we obtain $s^{\pm}_{\theta,\Gamma}(y_j)=0$ , $t^+_{\theta,\Gamma}(y_j)=1$ and $t^-_{\theta,\Gamma}(y_j)=0$ , a contradiction. Thus, $m=n$ . We now prove that $|U_k|=1$ for all $1\le k\le n$ . If $|U_k|\ne 1$ , then $k\ne j$ and $|V_k|=1$ , say $V_k=\{q_k\}$ . We have $\{\theta(e_{x_1y_j}),\dots,\theta(e_{x_ny_j})\}=\{e_{p_jy_1},\dots,e_{p_jy_n}\}$ and $\{\theta(e_{x_1y_k}),\dots,\theta(e_{x_ny_k})\}=\{e_{x_1q_k},\dots,e_{x_nq_k}\}$ . Since $j\ne k$ , these sets must be disjoint. But $e_{p_jq_k}$ belongs to their intersection, a contradiction. Thus, $|U_k|=1$ for all $1\le k\le n$ . Replacing $\theta$ by $\theta'\circ\theta$ , where $\theta'(e_{xy})=e_{\mu(y)\mu(x)}$ and $\mu$ is the anti-automorphism of X which interchanges $x_i$ and $y_i$ for all $1\le i\le n$ , we get the situation of Case 1, so that $\theta'\circ\theta\in\mathcal{P}(X)$ . Since $\theta'\in\mathcal{P}({\rm{K}}_{m,n})$ , we concludethat $\theta\in\mathcal{P}({\rm{K}}_{m,n})$ .

Acknowledgements

The second author was partially supported by CNPq 404649/2018-1.

Footnotes

1 Often also called the height.

References

Baclawski, K., Automorphisms and derivations of incidence algebras, Proc. Amer. Math. Soc. 36(2) (1972), 351356.CrossRefGoogle Scholar
Bollobás, B., Modern graph theory, vol. 184. Grad. Texts Math. (Springer, New York, NY, 1998).CrossRefGoogle Scholar
Brusamarello, R., Fornaroli, É. Z. and Santulo, E. A. Jr., Anti-automorphisms and involutions on (finitary) incidence algebras, Linear Multilinear Algebra 60(2) (2012), 181188.CrossRefGoogle Scholar
Brusamarello, R., Fornaroli, É. Z. and Santulo, E. A. Jr., Classification of involutions on finitary incidence algebras, Int. J. Algebra Comput., 24(8) (2014), 10851098.CrossRefGoogle Scholar
Brusamarello, R., Fornaroli, É. Z. and Santulo, E. A. Jr., Multiplicative automorphisms of incidence algebras, Commun. Algebra 43(2) (2015), 726736.CrossRefGoogle Scholar
Brusamarello, R. and Lewis, D. W., Automorphisms and involutions on incidence algebras, Linear Multilinear Algebra 59(11) (2011), 12471267.CrossRefGoogle Scholar
Cao, Y., Automorphisms of certain Lie algebras of upper triangular matrices over a commutative ring, J. Algebra 189(2) (1997), 506513.Google Scholar
Cecil, A. J., Lie isomorphisms of triangular and block-triangular matrix algebras over commutative rings, Master’s thesis, University of Victoria, 2016.Google Scholar
Doković, D. Ž., Automorphisms of the Lie algebra of upper triangular matrices over a connected commutative ring, J. Algebra 170(1) (1994), 101110.CrossRefGoogle Scholar
Dokuchaev, M. and Novikov, B., On colimits over arbitrary posets, Glasg. Math. J. 58(1) (2016), 219228.CrossRefGoogle Scholar
Dräxler, P., Completely separating algebras, J. Algebra 165(3) (1994), 550565.CrossRefGoogle Scholar
Drozd, Y. and Kolesnik, P., Automorphisms of incidence algebras, Comm. Algebra 35(12) (2007), 38513854.CrossRefGoogle Scholar
Fornaroli, É. Z., Khrypchenko, M. and Santulo, E. A. Jr, Lie automorphisms of incidence algebras, To appear in Proc. Amer. Math. Soc. (2020). (arXiv:2012.06661v3). DOI: 10.1090/proc/15786.CrossRefGoogle Scholar
Hua, L. K., A theorem on matrices over a sfield and its applications, J. Chinese Math. Soc. (N.S.) 1 (1951), 110163.Google Scholar
Khripchenko, N. S., Automorphisms of finitary incidence rings, Algebra and Discrete Math. 9(2) (2010), 7897.Google Scholar
Martindale, W. S. 3rd, Lie isomorphisms of primitive rings, Proc. Amer. Math. Soc. 14 (1963), 909916.CrossRefGoogle Scholar
Martindale, W. S. 3rd, Lie isomorphisms of prime rings, Trans. Amer. Math. Soc. 142 (1969), 437455.CrossRefGoogle Scholar
Martindale, W. S. 3rd, Lie isomorphisms of simple rings, J. Lond. Math. Soc. 44 (1969), 213221.CrossRefGoogle Scholar
Rival, I., A fixed point theorem for finite partially orderes sets, J. Comb. Theory, Ser. A 21 (1976), 309318.CrossRefGoogle Scholar
Rota, G.-C., On the foundations of combinatorial theory. I. Theory of Möbius functions, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 2(4) (1964), 340368.Google Scholar
Spiegel, E., On the automorphisms of incidence algebras, J. Algebra 239(2) (2001), 615623.CrossRefGoogle Scholar
Spiegel, E. and O’Donnell, C. J., Incidence Algebras (Marcel Dekker, New York, NY, 1997).Google Scholar
Stanley, R., Structure of incidence algebras and their automorphism groups, Bull. Am. Math. Soc. 76 (1970), 12361239.CrossRefGoogle Scholar
Stanley, R. P., Enumerative combinatorics. Vol. 1, vol. 49. Camb. Stud. Adv. Math. (Cambridge University Press, Cambridge, 1997).CrossRefGoogle Scholar