Hostname: page-component-6bf8c574d5-rwnhh Total loading time: 0 Render date: 2025-02-23T05:35:38.111Z Has data issue: false hasContentIssue false

Constructing skew left braces whose additive group has trivial centre

Published online by Cambridge University Press:  30 January 2025

Adolfo Ballester-Bolinches*
Affiliation:
Departament de Matemàtiques, Universitat de València, Dr. Moliner, 50, 46100 Burjassot, València, Spain e-mail: [email protected] [email protected]
Ramón Esteban-Romero
Affiliation:
Departament de Matemàtiques, Universitat de València, Dr. Moliner, 50, 46100 Burjassot, València, Spain e-mail: [email protected] [email protected]
Paz Jiménez-Seral
Affiliation:
Departamento de Matemáticas, Universidad de Zaragoza, Pedro Cerbuna, 12, 50009 Zaragoza, Spain e-mail: [email protected]
Vicent Pérez-Calabuig
Affiliation:
Departament de Matemàtiques, Universitat de València, Dr. Moliner, 50, 46100 Burjassot, València, Spain e-mail: [email protected] [email protected]
Rights & Permissions [Opens in a new window]

Abstract

A complete description of all possible multiplicative groups of finite skew left braces whose additive group has trivial centre is given. As a consequence, some earlier results of Tsang can be improved and an answer to an open question set by Tsang at Ischia Group Theory 2024 Conference is provided.

Type
Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Canadian Mathematical Society

1 Introduction

Skew brace structure plays a key role in the combinatorial theory of the Yang–Baxter equation. Skew left braces, introduced in [Reference Guarnieri and Vendramin8], can be regarded as extensions of Jacobson radical rings and show connections with several areas of mathematics such as triply factorized groups and Hopf–Galois structures (see [Reference Ballester-Bolinches and Esteban-Romero1, Reference Caranti and Stefanello3, Reference Childs4])

Skew left braces classify solutions of the Yang–Baxter equation (see [Reference Guarnieri and Vendramin8]). This connection to the Yang–Baxter equation motivates the search for constructions of skew braces and classification results.

Recall that a skew left brace is a set endowed with two group structures $(B,+)$ , not necessarily abelian, and $(B,\cdot )$ which are linked by the distributive-like law ${a(b+c)}=ab-a+ac$ for a, b, $c\in B$ .

In the sequel, the word brace refers to a skew left brace.

Given a brace B, there is an action of the multiplicative group on the additive group by means of the so-called lambda map:

$$\begin{align*}\lambda \colon a \in (B,\cdot) \longmapsto \lambda_a \in \operatorname{\mathrm{Aut}}(B,+), \quad \lambda_a(b)=-a+ab, \ \text{for all }a,b\in B.\end{align*}$$

Braces can be described in terms of regular subgroups of the holomorph of the additive group. Recall that the holomorph of a group G is the semidirect product $\operatorname {\mathrm {Hol}}(G) = [G]\operatorname {\mathrm {Aut}}(G)$ . Let B be a brace and set $K = (B,+)$ . Then $H=\{ (a,\lambda _a) \mid a\in B\}$ is a regular subgroup of the holomorph $\operatorname {\mathrm {Hol}}(K)$ isomorphic to $(B, {\cdot })$ (see [Reference Guarnieri and Vendramin8, Theorem 4.2]). If we consider the subgroup $S = KH \leq \operatorname {\mathrm {Hol}}(K)$ , then

$$\begin{align*}S=KH=KE=HE,\end{align*}$$

where $E = \{(0,\lambda _b)\mid b \in B\}$ and $\operatorname {\mathrm {C}}_{E}(K)= K\cap E =H\cap E =1$ . We call $\mathsf {S}(B)=(S, K, H, E)$ the small trifactorized group associated with B.

In [Reference Tsang9], Tsang showed it is possible to construct finite braces by just looking at the automorphism group of the additive group instead of looking at the whole holomorph. This is a significant improvement both from an algebraic and computational approach.

Theorem 1 (see [Reference Tsang9, Corollary 2.2])

If the finite group G is the multiplicative group of a brace with additive group K, then there exist two subgroups X and Y of $\operatorname {\mathrm {Aut}}(K)$ that are quotients of G satisfying

$$\begin{align*}XY=X{\operatorname{\mathrm{Inn}}(K)}=Y{\operatorname{\mathrm{Inn}}(K)}.\end{align*}$$

She looked for a sort of converse of the above theorem in the case of finite braces with an additive group of trivial centre, and proved the following.

Theorem 2 (see [Reference Tsang9, Proposition 2.7])

Suppose that the centre of a finite group $(K, {+})$ is trivial and let P be a subgroup of $\operatorname {\mathrm {Aut}}(K)$ containing $\operatorname {\mathrm {Inn}}(K)$ . If $P=XY$ is a factorization by two subgroups X and Y such that $X\cap Y=1$ , $X{\operatorname {\mathrm {Inn}}(K)}=Y{\operatorname {\mathrm {Inn}}(K)}=P$ and X splits over $X\cap \operatorname {\mathrm {Inn}}(K)$ , then there exists a brace B whose additive group is isomorphic to $(K, {+})$ and whose multiplicative group is isomorphic to a semidirect product $[X\cap \operatorname {\mathrm {Inn}}(K)]Y$ for a suitable choice of the action $\alpha \colon Y\longrightarrow \operatorname {\mathrm {Aut}}(X\cap \operatorname {\mathrm {Inn}}(K))$ .

The above two theorems are the key to prove the main results of [Reference Tsang9, Reference Tsang10].

In [Reference Tsang11], Tsang posed the following question.

Question 3 Is it possible to extend Theorem 2 by dropping the assumption that X splits over $X\cap \operatorname {\mathrm {Inn}}(K)$ ?

The aim of this article is to give a complete characterization of the multiplicative groups of a brace with additive group of trivial centre. As a consequence, we present an improved version of Theorem 2 (on which the main result of [Reference Tsang9] heavily depends), and we give an affirmative answer to Question 3.

Theorem A Let K be a finite group with trivial centre. For every brace B with additive group $K = (B,+)$ and multiplicative group $C = (B,\cdot )$ , there exist subgroups X and Y of $Aut(K)$ satisfying the following properties:

  1. (a) $XY=X{\operatorname {\mathrm {Inn}}(K)}=Y{\operatorname {\mathrm {Inn}}(K)}$ ,

  2. (b) there are two subgroups N and M of $\operatorname {\mathrm {Inn}}(K)$ such that $N \unlhd X$ and $M \unlhd Y$ ,

  3. (c) there exists an isomorphism $\gamma \colon Y/M\longrightarrow X/N$ such that

    $$\begin{align*}\operatorname{\mathrm{Inn}}(K) = \{xy^{-1}\mid x\in X, \, y\in Y,\, \gamma(yM) = xN\},\end{align*}$$
  4. (d) $\lvert K\rvert =\lvert X\rvert \lvert M\rvert =\lvert Y\rvert \lvert N\rvert $ .

In this case,

  1. (e) C has two normal subgroups T and V with $T\cap V=1$ , $X\cong C/T$ and $Y\cong C/V$ , that is, C is a subdirect product of X and Y.

Conversely, for every pair X, Y of subgroups of $\operatorname {\mathrm {Aut}}(K)$ satisfying conditions (a)–(d), there exists a brace B with $K = (B,+)$ and $C= (B,\cdot )$ satisfying (e).

Corollary 4 Let K be a finite group with trivial centre. Suppose that there exist subgroups X, Y of $\operatorname {\mathrm {Aut}}(K)$ such that $X \cap Y = 1$ and $XY=X{\operatorname {\mathrm {Inn}}(K)}=Y{\operatorname {\mathrm {Inn}}(K)}$ . Then there exists a brace with additive group K and a multiplicative group that is isomorphic to a subdirect product of X and Y.

Proof Assume that $X\cap Y=1$ . Consider $N=X\cap {\operatorname {\mathrm {Inn}}(K)}$ , $M=Y\cap {\operatorname {\mathrm {Inn}}(K)}$ . Then $\lvert X\rvert \lvert M\rvert =\lvert K\rvert $ as $\lvert X\rvert \lvert Y\rvert =\lvert {\operatorname {\mathrm {Inn}}(K)}\rvert \lvert Y\rvert /\lvert Y\cap {\operatorname {\mathrm {Inn}}(K)}\rvert $ . Analogously, $\lvert Y\rvert \lvert N\rvert =\lvert K\rvert $ . Moreover, since

$$\begin{align*}Y/M\cong Y{\operatorname{\mathrm{Inn}}(K)}/{\operatorname{\mathrm{Inn}}(K)}=X{\operatorname{\mathrm{Inn}}(K)}/{\operatorname{\mathrm{Inn}}(K)}\cong X/N,\end{align*}$$

we have an isomorphism $\gamma \colon Y/M\longrightarrow X/N$ given by $\gamma (bM)=aN$ , where $b \in Y$ , $a \in X$ such that $ab^{-1} \in \operatorname {\mathrm {Inn}}(K)$ . Since $X \cap Y = 1$ , for each $k \in K$ , conjugation by k can be expressed as $ab^{-1}$ , for a unique $a\in X$ and $b\in Y$ . Then, the groups X and Y satisfy Statements (a)–(d) of Theorem A, and therefore, there exists a brace whose additive group is K and whose multiplicative group is isomorphic to a subdirect product of X and Y.

Corollary 4 also allows to give a considerably shorter proof of the main results of [Reference Tsang9, Reference Tsang10] about the almost simple groups K that can appear as additive groups of braces with soluble multiplicative group. By Corollary 4, it is enough to find two subgroups X and Y of $\operatorname {\mathrm {Aut}}(K)$ such that $X \cap Y = 1$ and $XY=X{\operatorname {\mathrm {Inn}}(K)}=Y{\operatorname {\mathrm {Inn}}(K)}$ . Therefore, Codes 2, 3, and 4 in the proof of [Reference Tsang9, Theorem 1.3] can be avoided, as well as checking in every case that the subgroup X splits over $X\cap {\operatorname {\mathrm {Inn}}(K)}$ .

In Section 3, we present a worked example of a construction of a brace with additive group $K = \operatorname {\mathrm {PSL}}_2(25)$ by means of subgroups X and Y of $\operatorname {\mathrm {Aut}}(K)$ satisfying all conditions of Theorem A but $X \cap Y \neq 1$ .

2 Proof of Theorem A

Proof of Theorem A

Suppose that B is a brace with additive group K and lambda map $\lambda $ . Let $H=\{(b, \lambda _b)\!\mid b\in B\}$ be the regular subgroup of $\operatorname {\mathrm {Hol}}(K)$ appearing in the small trifactorized group $\mathsf {S}(B)=(S, K, H, E)$ associated with B. Recall that H is isomorphic to the multiplicative group $(C,\cdot )$ of B, $E= \{(0, \lambda _b)\!\mid b \in B\} \leq \operatorname {\mathrm {Hol}}(K)$ , and $S = KH = KE = HE$ with $K\cap E=H\cap E= 1$ .

Observe that S acts on K by means of the homomorphism $\pi \colon (b, \omega )\in S \mapsto \omega \in \operatorname {\mathrm {Aut}}(K)$ . On the other hand, S also acts on K by conjugation. In fact, this action naturally induces a homomorphism $\alpha \colon S \rightarrow \operatorname {\mathrm {Aut}}(K)$ . In particular, for every $b\in B$ and every $k \in K$ , $(0,\lambda _b)(k,1)(0,\lambda _b)^{-1}=(\lambda _b(k), 1)$ , that is, $\alpha (0,\lambda _b) = \lambda _b = \pi (0,\lambda _b)$ . Thus, $\alpha (E) = \pi (E) = \pi (H)$ .

Figure 1 Structure of the multiplicative group in Theorem A.

The restrictions of $\pi $ and $\alpha $ to H induce two actions of H on K, with respective kernels $\operatorname {\mathrm {Ker}}{\pi }|_H=K\cap H\trianglelefteq H$ and $\operatorname {\mathrm {Ker}}{\alpha |_H}=\operatorname {\mathrm {C}}_H(K)\trianglelefteq H$ . Moreover, it holds that

$$ \begin{align*} \operatorname{\mathrm{Ker}}{\pi}|_H\cap \operatorname{\mathrm{Ker}}{\alpha|_H} & = K\cap H\cap \operatorname{\mathrm{C}}_H(K)= K\cap H\cap \operatorname{\mathrm{C}}_S(K) \\ & = H\cap \operatorname{\mathrm{C}}_{K}(K)=H\cap \operatorname{\mathrm{Z}}(K)=1 \quad \text{(see Figure~1).} \end{align*} $$

Let $X:=\alpha (H)$ and $Y := \pi (H) = \alpha (E) = \{\lambda _b\mid b\in B\}$ such that $X \cong H/{\operatorname {\mathrm {C}}_H(K)}$ and $Y \cong H/(K\cap H)$ . Since $\alpha (K)=\operatorname {\mathrm {Inn}}(K)$ , we have that

$$ \begin{align*} \alpha(S)&=\alpha(HE)=\alpha(KH)=\alpha(KE)\\ &=\alpha(H)\alpha(E)=\alpha(K)\alpha(H)=\alpha(K)\alpha(E)\\ &=XY=(\operatorname{\mathrm{Inn}}(K))X=(\operatorname{\mathrm{Inn}}(K))Y. \end{align*} $$

Take $R:=(H\cap K){\operatorname {\mathrm {C}}_H(K)} \unlhd H$ . Then, $N:=\alpha (R)\trianglelefteq \alpha (H)=X$ and $ M:= \pi (R)\trianglelefteq \pi (H) = Y$ . It follows that $N =\alpha (H\cap K)\le \alpha (K)=\operatorname {\mathrm {Inn}}(K)$ . On the other hand, $M = \pi (\operatorname {\mathrm {C}}_H(K))$ and if $(b,\lambda _b)\in {\operatorname {\mathrm {C}}_H(K)}$ , then for every $k\in K$ ,

$$\begin{align*}(b, \lambda_b)(k,1)(b,\lambda_b)^{-1}=(b+\lambda_b(k)-b,1)=(k,1), \end{align*}$$

that is, $\lambda _b$ coincides with the inner automorphism of K induced by $-b$ . Thus, $M \le \operatorname {\mathrm {Inn}}(K)$ . Moreover, we see that

$$ \begin{align*} Y/M &\cong (H/\operatorname{\mathrm{Ker}}{\pi}|_H)/(R/\operatorname{\mathrm{Ker}}{\pi}|_H)\cong H/R\\ &\cong(H/\operatorname{\mathrm{Ker}}\alpha|_H)/(R/\operatorname{\mathrm{Ker}}\alpha|_H)\cong X/N; \end{align*} $$

here the isomorphism $\gamma \colon Y/M\longrightarrow X/N$ is given by $\gamma (\lambda _b M)=\alpha _b\lambda _b N$ , where $\alpha _b$ is the inner automorphism of K induced by b. Given $a\in \gamma (\lambda _bM)$ , we have that $a\lambda _b^{-1}\in \alpha _bN\subseteq \operatorname {\mathrm {Inn}}(K)$ . Furthermore, given $x\in \operatorname {\mathrm {Inn}}(K)$ , we have that $x=\alpha _b$ for some $b\in B$ and so $\gamma (\lambda _b M)=\alpha _b\lambda _b N=x\lambda _b N$ with $(\alpha _b\lambda _b)\lambda _b^{-1}=x$ .

Since $\operatorname {\mathrm {Ker}}{\pi |}_H\cap {\operatorname {\mathrm {Ker}}\alpha |_H}=(H\cap K)\cap \operatorname {\mathrm {C}}_H(K)=1$ , we have that $\lvert R\rvert ={\lvert H\cap K\rvert }\lvert {\operatorname {\mathrm {C}}_H(K)}\rvert $ and $\lvert M\rvert =\lvert R/(H\cap K)\rvert =\lvert \operatorname {\mathrm {C}}_H(K)\rvert $ , $\lvert N\rvert =\lvert R/{\operatorname {\mathrm {C}}_H(K)}\rvert =\lvert H\cap K\rvert $ . As $\lvert X\rvert =\lvert K\rvert /\lvert {\operatorname {\mathrm {C}}_H(K)}\rvert $ and $\lvert Y\rvert =\lvert K\rvert /\lvert H\cap K\rvert $ , the claim about the order follows.

Item (e) follows by the fact that H is isomorphic to the multiplicative group $(C,\cdot )$ of B, so that T and V are respectively isomorphic to $\operatorname {\mathrm {Ker}} \alpha |_H$ and $\operatorname {\mathrm {Ker}} \pi |_H$ .

Now, suppose that $\operatorname {\mathrm {Aut}}(K)$ possesses subgroups X and Y satisfying conditions (a)–(d). Let

$$\begin{align*}W=\{(x,y)\mid x\in X,\, y\in Y,\, \gamma(yM)=xN\}\end{align*}$$

be a subdirect product of X and Y with amalgamated factor group $Y/M\cong X/N$ (see [Reference Doerk and Hawkes6, Chapter A, Definition 19.2]). By [Reference Doerk and Hawkes6, Chapter A, Proposition 19.1], and the hypothesis, we have that $\lvert W\rvert =\lvert K\rvert $ . Since $\operatorname {\mathrm {Z}}(K)$ is trivial, the map $\zeta \colon K\longrightarrow \operatorname {\mathrm {Inn}}(K)$ , where $\zeta (k)$ is the inner automorphism of K induced by k, is an isomorphism. By hypothesis, the map $W\longrightarrow \operatorname {\mathrm {Inn}}(K)$ given by $(x,y)\longmapsto xy^{-1}$ is surjective. Since $\lvert W\rvert =\lvert \operatorname {\mathrm {Inn}}(K)\rvert =\lvert K\rvert $ , it is a bijection. We can consider $H=\{(b,y)\mid (x,y)\in W,\, \zeta (b)=xy^{-1}\}\subseteq \operatorname {\mathrm {Hol}}(K)$ . Given $(b, y)$ , $(b_1, y_1)\in H$ , we have that $(b, y)(b_1,y_1)=(b+y(b_1),yy_1)$ , $\zeta (b)=xy^{-1}$ , and $\zeta (b_1)=x_1y_1^{-1}$ with $(x,y)$ , $(x_1, y_1)\in B$ . Then

$$\begin{align*}\zeta(b+y(b_1))=\zeta(b)\zeta(y(b_1))=\zeta(b)y\zeta(b_1)y^{-1}=xy^{-1}yx_1y_1^{-1}y^{-1}=(xx_1)(yy_1)^{-1}\end{align*}$$

with $(xx_1, yy_1)=(x,y)(x_1,y_1)\in W$ . Furthermore, if $(b, y)\in H$ , with $\zeta (b)=xy^{-1}$ , we have that $(b, y)^{-1}=(y^{-1}(-b), y^{-1})$ and

$$\begin{align*}\zeta(y^{-1}(-b))=y^{-1}\zeta(-b)y=y^{-1}\zeta(b)^{-1}y=y^{-1}yx^{-1}y=x^{-1}{(y^{-1})}^{-1}\end{align*}$$

with $(x^{-1}, y^{-1})=(x,y)^{-1}\in W$ . We conclude that H is a subgroup of $\operatorname {\mathrm {Hol}}(K)$ . As the projection onto its first component is surjective, it turns out that it H is a regular subgroup of $\operatorname {\mathrm {Hol}}(K)$ by [Reference Ballester-Bolinches, Esteban-Romero and Pérez-Calabuig2, Proposition 2.5] and so it is isomorphic to the multiplicative group of a brace with additive group K (see [Reference Guarnieri and Vendramin8, Theorem 4.2]).

We finish the proof by showing that the map $\phi \colon H\to W$ given by $(b, y)\longmapsto (\zeta (b)y, y)$ , where $\zeta (b)=xy^{-1}$ and $(x, y)\in W$ , is an isomorphism. Indeed, if $\zeta (b)=xy^{-1}$ , $\zeta (b_1)=x_1y_1^{-1}$ , where $(x,y)$ , $(x_1,y_1)\in W$ , we have that

$$ \begin{align*} \phi(b,y)\phi(b_1,y_1)&=(\zeta(b)y,y)(\zeta(b_1)y_1,y_1)=(x,y)(x_1,y_1)=(xx_1,yy_1),\\ \phi((b,y)(b_1,y_1))&=\phi(b+y(b_1),yy_1)=(\zeta(b+y(b_1))yy_1,yy_1)\\ &=(\zeta(b)y\zeta(b_1)y^{-1}yy_1,yy_1)=(xy^{-1}yx_1y_1^{-1}y_1,yy_1)\\ &=(xx_1,yy_1). \end{align*} $$

We conclude that $\phi $ is a group homomorphism. Assume that $\phi (b, y)=(\zeta (b)y,y)=(1,1)$ , with $\zeta (b)=xy^{-1}$ and $(x,y)\in W$ , then $y=1$ and so $\zeta (b)=x=1$ , which implies that $b=0$ . Consequently, $\phi $ is injective. As W and H are finite and have the same order, we obtain that $\phi $ is an isomorphism. Since C is isomorphic to H we have just proved that (e) holds for C.

3 A worked example

In general, we do not have that $X\cap Y=1$ . Let us consider $K=\operatorname {\mathrm {PSL}}_2(25)$ . Its automorphism group $A=\operatorname {\mathrm {Aut}}(K)$ is generated by $\operatorname {\mathrm {Inn}}(K)$ , the diagonal automorphism d induced by the conjugation by the matrix

$$\begin{align*}\mathsf{D}= \begin{bmatrix} \zeta&0\\ 0&1 \end{bmatrix}\in\operatorname{\mathrm{GL}}_2(25),\end{align*}$$

where $\zeta $ is a primitive $24$ th-root of unity of $\operatorname {\mathrm {GF}}(25)$ , and the field automorphism f. The group A possesses a subgroup X generated by the inner automorphisms $c_1$ , $c_2$ , and $c_3$ induced by the matrices

$$\begin{align*}\mathsf{C}_1= \begin{bmatrix} \zeta^{4}&0\\ 0&\zeta^{20} \end{bmatrix},\qquad\mathsf{C}_2= \begin{bmatrix} 1&0\\ \zeta&1 \end{bmatrix},\qquad \mathsf{C}_3= \begin{bmatrix} 1&0\\ 1&1 \end{bmatrix}, \end{align*}$$

respectively, and $df$ . We have that $c_1$ has order $3$ , $\langle c_2, c_3\rangle $ is an elementary abelian group of order $25$ , $c_1$ normalises $\langle c_2, c_3\rangle $ , $(df)c_1(df)^{-1}=c_1^{-1}$ , $df$ has order $8$ , and $df$ normalizes $\langle c_2, c_3\rangle $ . Then the group $\langle df, c_1, c_2, c_3\rangle $ has order $600$ .

Let $u_1$ and $u_2$ be the inner automorphisms induced by the conjugation by

$$\begin{align*}\mathsf{U_1}=\begin{bmatrix} \zeta^3&\zeta^{16}\\ \zeta^{13}&\zeta^{11} \end{bmatrix},\qquad \mathsf{U_2}= \begin{bmatrix} \zeta^{5}&\zeta^5\\ \zeta^9&\zeta^{22} \end{bmatrix}. \end{align*}$$

Let $Y=\langle u_1, dfu_2\rangle $ . We have that $u_1$ has order $13$ . Let

$$\begin{align*}\mathsf{R}= \begin{bmatrix} \zeta&0\\ 0&\zeta \end{bmatrix}\in\operatorname{\mathrm{Z}}(\operatorname{\mathrm{GL}}_2(25)),\qquad \mathsf{T}= \begin{bmatrix} 3&0\\ 4&2 \end{bmatrix} \end{align*}$$

and let t be the automorphism induced by conjugation by $\mathsf {T}$ . Then $(dfu_2)^2=dfu_2dfu_2=d^{f}{{u_2}}{}d^5u_2$ is the automorphism induced by conjugation by

(1) $$ \begin{align} \mathsf{D}\mathsf{U}_2^{(5)}\mathsf{D} ^5\mathsf{U}_2=\mathsf{R}^{15}\mathsf{T},\end{align} $$

where $\mathsf {U}_2^{(5)}$ denotes the matrix whose entries are obtained from the entries of $\mathsf {U}_2$ by applying the Frobenius field automorphism, that is, $(dfu_2)^2=t$ . As $(\mathsf {R}^{15}\mathsf {T})^2=\mathsf {R}^3$ , we conclude that $dfu_2$ has order $4$ . We can also check that $(dfu_2)u_1(dfu_2)^{-1}=u_1^8$ . It follows that Y has order $52$ .

By [Reference Conway, Curtis, Norton, Parker and Wilson5], X and Y are maximal subgroups of the almost simple group $\operatorname {\mathrm {Inn}}(K)\langle df\rangle $ . Observe that $(df)^4c_3^2=(dfdf)^2c_3^2=(dd^5)^2c_3^2=d^{12}c_3^2$ is induced by $\mathsf {D}^{12}\mathsf {C}_3^2=\mathsf {R}^{18}\mathsf {T}$ , consequently, $(df)^4c_3^2=t$ . This, together with Equation (1), shows that $t\in X\cap Y$ . We note that $\operatorname {\mathrm {Inn}}(K)X=\operatorname {\mathrm {Inn}}(K)Y=\operatorname {\mathrm {Inn}}(K)\langle df\rangle $ . Moreover, $\lvert {X\cap Y}\rvert $ divides $\gcd (\lvert X\rvert , \lvert Y\rvert )=4$ . If $\lvert X\cap Y\rvert = 4$ , then $X\cap Y$ is contained in $X\cap \operatorname {\mathrm {Inn}}(K)$ , but it is not contained in $Y\cap \operatorname {\mathrm {Inn}}(K)$ . This shows that $\lvert X\cap Y\rvert \le 2$ . Hence $\lvert X\cap Y\rvert = 2$ . As $XY\subseteq \operatorname {\mathrm {Inn}}(K)\langle df\rangle $ ,

$$\begin{align*}15\,600=\lvert \operatorname{\mathrm{Inn}}(K)\langle df\rangle\rvert\ge \lvert XY\rvert=\frac{\lvert X\rvert \lvert Y\rvert}{\lvert X\cap Y\rvert}=15\,600\cdot \frac{2}{\lvert X\cap Y\rvert}=15\,600,\end{align*}$$

and so $XY=\operatorname {\mathrm {Inn}}(K)\langle df\rangle $ .

Let $N=\langle c_1, c_2, c_3, (df)^2\rangle \trianglelefteq X$ , $M=\langle u_1\rangle \trianglelefteq Y$ . Then $\lvert N\rvert =150$ , $\lvert M\rvert =13$ , $N\le X\cap \operatorname {\mathrm {Inn}}(K)$ , $M\le Y\cap \operatorname {\mathrm {Inn}}(K)$ , $Y/M\cong X/N\cong C_4$ , and $\lvert K\rvert =\lvert X\rvert \lvert M\rvert = \lvert Y\rvert \lvert N\rvert $ . The isomorphism between $Y/M$ and $X/N$ is given by $\gamma ((dfu_2)^rM)=(df)^rN$ for $0\le r < 4$ , and, since $d^6\in \operatorname {\mathrm {Inn}}(K)$ , it is clear that

$$ \begin{align*} (df)(dfu_2)^{-1}&=c_0u_0^{-1}\in \operatorname{\mathrm{Inn}}(K),\\ (df)^2(dfu_2)^{-2}&=d^6t^{-1}\in \operatorname{\mathrm{Inn}}(K),\\ (df)^3(dfu_2)^{-3}&=(df)(d^6t^{-1}u_2^{-1})(df)^{-1}\in\operatorname{\mathrm{Inn}}(K). \end{align*} $$

Let $z\in XY\cap \operatorname {\mathrm {Inn}}(K)$ . Recall that $X\cap Y=\langle t\rangle $ . Then there exist $x\in X$ , $y\in Y$ with $z=xy^{-1}=(xt)(yt)^{-1}$ . We observe that $t=(df)^4c_3^2\in N$ , but $t\notin M$ by order considerations. Given $x\in X$ , $y\in Y$ , there exist r, $s\in \{0,1,2,3\}$ such that $xN=(df)^rN$ and $yM=(dfc_3)^sM$ . We also observe that $x\in \operatorname {\mathrm {Inn}}(K)$ if, and only if, $y\in \operatorname {\mathrm {Inn}}(K)$ . To prove that we can choose $x\in X$ , $y\in Y$ such that $z=xy^{-1}$ and $\gamma (yM)=xN$ , it is enough to prove that for such a choice we have that $z=xy^{-1}$ and $r=s$ . Note that if $x\in N$ , then $r=0$ ; if $x\in \operatorname {\mathrm {Inn}}(K)\setminus N$ , then $r=2$ ; and if $x\notin \operatorname {\mathrm {Inn}}(K)$ , then $r\in \{1,3\}$ . Analogously, if $y\in M$ , then $s=0$ ; if $y\in \operatorname {\mathrm {Inn}}(K)\setminus M$ , then $s=2$ ; and if $y\notin \operatorname {\mathrm {Inn}}(K)$ , then $s\in \{1,3\}$ . We also have that $tM=(dfu_2)^2M$ and that $tN=N$ , as $t\in \operatorname {\mathrm {Inn}}(K)$ , $t\in N$ , but $t\notin M$ . If $x\in N$ and $y\in M$ , we can choose $r=s=0$ and $\gamma (yM)=xN$ . Suppose that $x\in N$ and $y\notin M$ . Then $y\in \operatorname {\mathrm {Inn}}(K)$ and so, $xN=N$ and $yM=(dfu_2)^2M$ . Consequently, $xtN=N$ , $ytN=N$ , and $\gamma (ytN)=xtN$ . Suppose that $x\notin N$ and $y\in M$ . We have that $x\in \operatorname {\mathrm {Inn}}(K)$ and so, $xN=(df)^2N$ and $yM=M$ . It follows that $xtN=(df)^2N$ and $ytM=(dfu_2)^2M$ , that is, $\gamma (ytM)=xtN$ . Suppose that x, $y\in \operatorname {\mathrm {Inn}}(K)$ , $x\notin N$ , and $y\notin M$ . Then $xN=(df)^2N$ , $yM=(dfu_2)^2M$ , and $\gamma (yM)=xN$ . Finally, suppose that x and $y\notin M$ . Then $xN=(df)^rN$ and $yM=(dfu_2)^sM$ , with r, $s\in \{1,3\}$ . If $r=s$ , then $\gamma (yM)=xN$ . If $r\ne s$ , then $xtN=(df)^rN$ and $ytM=(dfu_2)^{s+2}M$ , with $r\equiv s+2\pmod {4}$ . Thus $\gamma (ytM)=xtN$ .

It follows that X, Y satisfy all conditions of Theorem A. We can also check with GAP [7] all this information about these subgroups.

Acknowledgements

We would like to thank the referee for his thoughtful comments toward improving our article.

References

Ballester-Bolinches, A. and Esteban-Romero, R., Triply factorised groups and the structure of skew left braces . Commun. Math. Stat. 10(2022), 353370.Google Scholar
Ballester-Bolinches, A., Esteban-Romero, R., and Pérez-Calabuig, V., Enumeration of left braces with additive group ${C}_4\times {C}_4\times {C}_4$ . Math. Comput. 93(2024), no. 346, 911919.Google Scholar
Caranti, A., and Stefanello, L., From endomorphisms to bi-skew braces, regular subgroups, the Yang-Baxter equation, and Hopf-Galois structures . J. Algebra 587(2021), 462487.Google Scholar
Childs, L. N., Skew braces and the Galois correspondence for Hopf Galois structures . J. Algebra 511(2018), 270291.Google Scholar
Conway, J. H., Curtis, R. T., Norton, S. P., Parker, R. A., and Wilson, R. A., Atlas of finite groups, Oxford University Press, London, 1985.Google Scholar
Doerk, K. and Hawkes, T., Finite soluble groups, volume 4 of De Gruyter Expositions in Mathematics, Walter de Gruyter & Co., Berlin, 1992.Google Scholar
The GAP Group, GAP—Groups, algorithms, and programming, Version 4.14.0, 2024. http://www.gap-system.org.Google Scholar
Guarnieri, L. and Vendramin, L., Skew-braces and the Yang-Baxter equation . Math. Comput. 86(2017), no. 307, 25192534.Google Scholar
Tsang, C. (S. Y.), Non-abelian simple groups which occur as the type of a Hopf–Galois structure on a solvable extension . Bull. London Math. Soc. 55(2023), no. 5, 23242340.Google Scholar
Tsang, C. (S. Y.), Finite almost simple groups whose holomorph contains a solvable regular subgroup . Adv. Group Theory Appl., in press, arXiv preprint: 2312.15745, 2023.Google Scholar
Tsang, C. (S. Y.), Factorizations of groups and skew braces . In: Ischia Group Theory 2024 Conference, Ischia, Italy, 2024.Google Scholar
Figure 0

Figure 1 Structure of the multiplicative group in Theorem A.