Hostname: page-component-745bb68f8f-f46jp Total loading time: 0 Render date: 2025-02-11T09:30:54.358Z Has data issue: false hasContentIssue false

Uniconnected solutions to the Yang–Baxter equation arising from self-maps of groups

Published online by Cambridge University Press:  20 April 2021

Wolfgang Rump*
Affiliation:
Institute for Algebra and Number Theory, University of Stuttgart, Pfaffenwaldring 57, D-70550Stuttgart, Germany
Rights & Permissions [Opens in a new window]

Abstract

Set-theoretic solutions to the Yang–Baxter equation can be classified by their universal coverings and their fundamental groupoids. Extending previous results, universal coverings of irreducible involutive solutions are classified in the degenerate case. These solutions are described in terms of a group with a distinguished self-map. The classification in the nondegenerate case is simplified and compared with the description in the degenerate case.

Type
Article
Copyright
© Canadian Mathematical Society 2021

1 Introduction

Very soon after Drinfeld’s suggestion [Reference Drinfeld9] to study set-theoretic solutions $S\colon X^2\rightarrow X^2$ to the Yang–Baxter equation

$$ \begin{align*}(S\times 1_X)(1_X\times S)(S\times 1_X)=(1_X\times S)(S\times 1_X)(1_X\times S)\end{align*} $$

in $X^3$ , these solutions found increasing attention during the past 30 years (see, e.g., [Reference Angiono, Galindo and Vendramin1, Reference Catino and Rizzo4Reference Dehornoy8, Reference Etingof, Schedler and Soloviev10Reference Gateva-Ivanova12, Reference Gateva-Ivanova14, Reference Gateva-Ivanova and Van den Bergh15, Reference Lebed and Vendramin18, Reference Lu, Yan and Zhu19, Reference Rump21, Reference Rump23, Reference Rump25Reference Weinstein and Xu30]. Due to their close connection with braidings, they arise in various topics, including noncommutative regular rings [Reference Gateva-Ivanova13, Reference Gateva-Ivanova14], regular affine groups [Reference Catino, Colazzo and Stefanelli2Reference Catino and Rizzo4], Hopf–Galois theory [Reference Angiono, Galindo and Vendramin1, Reference Childs6, Reference Featherstonhaugh, Caranti and Childs11, Reference Guarnieri and Vendramin16], and Garside groups [Reference Chouraqui and Godelle7, Reference Dehornoy8, Reference Rump23].

A first systematic study of involutive solutions was undertaken by Etingof et al. [Reference Etingof, Schedler and Soloviev10] who introduced the structure group $(G_X;\circ )$ of a solution $S(x,y)=({{}}^xy,x^y)$ , generated by X, with relations $x\circ y ={{}}^xy\circ x^y$ . A solution S is said to be nondegenerate if the component maps $y\mapsto {{}}^xy$ and $y\mapsto y^x$ are invertible for all $x\in X$ . In [Reference Rump21], it was shown that left nondegenerate involutive solutions S are equivalent to cycle sets, that is, sets X with a single binary operation $\cdot $ such that the self-maps $\sigma (x)\colon X\rightarrow X$ with $\sigma (x)(y):=x\cdot y$ are bijective, and X satisfies the equation

(1.1) $$ \begin{align} (x\cdot y)\cdot (x\cdot z)= (y\cdot x)\cdot (y\cdot z). \end{align} $$

Furthermore, it was shown that finite cycle sets are (two-sided) nondegenerate.

The structure group $G_X$ of a nondegenerate cycle set X is obtained by a process of discrete integration where equation (1.1) plays the role of an integrability condition. More precisely, X embeds into $G_X$ , and the operation on X extends to $G_X$ and makes $G_X$ into a cycle set. The action $G_X\times X\rightarrow X$ given by $(a,x)\mapsto a\cdot x$ leads to a factor group $G(X)$ of $G_X$ , acting faithfully on X, such that $G(X)$ is generated by the permutations $\sigma (x)$ with $x\in X$ . Therefore, $G(X)$ is called the permutation group of X. If X is nondegenerate, $G(X)$ is also a cycle set, induced by the cycle set structure of $G_X$ .

Until now, several types of cycle sets with an underlying group structure have been found. The cycle sets $G_X$ and $G(X)$ are two examples. Both are braces [Reference Rump22], that is, additive abelian groups with a ring-like multiplication where $(G_X;\circ )$ plays the role of an adjoint group, similar to that of a radical ring: $a\circ b=ab+a+b$ . Any brace satisfies

$$ \begin{align*}(a+b)\cdot c=(a\cdot b)\cdot(a\cdot c),\end{align*} $$

so that equation (1.1) follows by the commutativity of addition. Etingof et al. [Reference Etingof, Schedler and Soloviev10] already considered affine solutions to the Yang–Baxter equation. These also can be described as cycle sets with an underlying abelian group. In [Reference Rump24], we studied two other types of cycle set structures on an abelian group. Given the close connection between group structures and symmetries, the cumulative occurrence of groups in relation with cycle sets indicates a high level of inner balance of this structure. Surprisingly, there is another connection with groups arising in the context of coverings.

In this paper, we introduce a class of self-maps of a group G, leading to a cycle set structure on G, such that the universal covering of any cycle set is of that type. In [Reference Rump26], we developed a general covering theory which applies to noninvolutive and even degenerate solutions to the Yang–Baxter equation. Here, we focus upon the involutive case. To highlight the analogy with topological coverings of connected spaces, special attention is payed to irreducible cycle sets, consisting of a single orbit under the permutation group.

By [Reference Rump26, Theorem 6], every cycle set X has a universal covering $\widetilde {p}\colon \widetilde {X}\twoheadrightarrow X$ , a cycle set morphism which does not change the permutation group $G(X)$ , and does not increase the set of $G(X)$ -orbits. Like in topology, it is universal, hence essentially unique, so that it factors through each covering of X. The relationship between X and $\widetilde {X}$ is determined by the fundamental groupoid $\pi _1(X)$ , which can be replaced by a group if X is irreducible. Thus, in order to classify all irreducible cycle sets, the main step consists in the knowledge of cycle sets arising as universal coverings, i.e., those with a trivial fundamental group. In [Reference Rump27], these uniconnected cycle sets, analogous to simply connected spaces, are classified in the nondegenerate case. Here, we remove the restriction by characterizing arbitrary uniconnected cycle sets as groups with a specific self-map (Theorem 3.2).

Using a result of [Reference Rump28], we show that nondegenerate uniconnected cycle sets are equivalent to braces with a distinguished generator (Theorem 3.3). Here, the uniconnected cycle set structure differs from the cycle set structure as a brace! As a corollary, it follows that a uniconnected cycle set is nondegenerate if and only if its underlying group is the adjoint group of a brace. It turns out that a great many of universal coverings are degenerate: For example, each group with a right-invariant lattice structure (e.g., Artin’s braid group) gives rise to a degenerate uniconnected cycle set (Example 1). A class of nondegenerate uniconnected cycle sets is obtained from differential groups (Example 2).

2 Preliminaries

In this section, we briefly recall some facts on cycle sets needed in what follows. A set X with a binary operation $\cdot $ is said to be a cycle set [Reference Rump21] if the left multiplication $\sigma (x)$ : $y\mapsto x\cdot y$ is invertible, and the equation (1.1) holds for all $x,y,z\in X$ . So $\sigma $ defines a map

(2.1) $$ \begin{align} \sigma: X\rightarrow\mathfrak{S}(X) \end{align} $$

into the permutation group $\mathfrak {S}(X)$ . The subgroup $G(X)$ of $\mathfrak {S}(X)$ generated by the image of $\sigma $ is said to be the permutation group of X. The image $\sigma (X)$ admits a well-defined operation

$$ \begin{align*}\sigma(x)\cdot\sigma(y):=\sigma(x\cdot y),\end{align*} $$

so that equation (1.1) is retained for $\sigma (X)$ , the retraction [Reference Etingof, Schedler and Soloviev10, Reference Rump21] of X. In general, $\sigma (X)$ is not a cycle set, unless X is nondegenerate, that is, the square map $x\mapsto x\cdot x$ is bijective. The retraction $\sigma (X)$ is then again nondegenerate, so that the retraction process can be iterated. Finite cycle sets are always nondegenerate [Reference Rump21, Theorem 2].

A brace A is a cycle set with an abelian group structure satisfying

(2.2) $$ \begin{align} a\cdot (b+c) &= (a\cdot b) + (a\cdot c), \end{align} $$
(2.3) $$ \begin{align} (a+b)\cdot c &= (a\cdot b)\:\cdot\: (a\cdot c), \end{align} $$

for all $a,b,c\in A$ . Note that equation (1.1) follows by equation (2.3) and the commutativity of $(A;+)$ . Equation (2.2) says that $\sigma (a)$ is a group automorphism for all $a\in A$ . As a cycle set, any brace A is nondegenerate. Moreover, it is a group with respect to the operation

(2.4) $$ \begin{align} a\circ b:=a^b+b, \end{align} $$

where $a^b:=\sigma (b)^{-1}(a)$ . The group $A^\circ :=(A;\circ )$ is called the adjoint group of A. Typical examples of braces are radical rings R (=Jacobson radicals of rings, viewed as pseudorings) with the operation $a\cdot b:=b(1+a)^{-1}$ . Then, the adjoint group is given by Jacobson’s circle operation [Reference Jacobson17]

$$ \begin{align*}a\circ b=a+ab+b.\end{align*} $$

In accordance with Jacobson’s notation [Reference Jacobson17], we write $a'$ for the inverse in the adjoint group of a brace. Note that the unit element of $A^\circ $ is 0.

For any nondegenerate cycle set X, the permutation group $G(X)$ is the adjoint group of a unique brace $A(X)$ such that $\sigma \colon X\rightarrow A(X)$ is a cycle set morphism. A subset of a brace A is said to be a cycle base [Reference Rump22, Definition 4] if it is invariant under the adjoint group $A^\circ $ and generates the additive group of A. By [Reference Rump28, Proposition 10], a brace A is generated by an element $e\in A$ if and only if $X:=\{e^a\:|\:a\in A\}$ is a cycle base of A. Thus, each transitive cycle base (i.e., such that $A^\circ $ acts transitively on it) is given by a generator $e\in A$ .

A right ideal [Reference Rump22] of a brace A is an additive subgroup which is invariant under the action of $A^\circ $ . By equation (2.4), any right ideal is a subbrace. For radical rings, this concept coincides with the usual one. We say that a brace A is a torsion brace if each $a\in A$ is of finite order in $(A;+)$ . For a torsion brace A, equation (2.2) implies that the primary decomposition of the additive group is a decomposition into right ideals.

The fundamental groupoid [Reference Rump26] $\pi _1(X)$ of a cycle set X has X as set of objects, and pairs $(a,x)\in G(X)\times X$ as morphisms from x to $a(x)$ . Composition is the obvious one:

$$ \begin{align*}x\longrightarrow b(x)\longrightarrow ab(x),\end{align*} $$

given by $(a,b(x))(b,x):=(ab,x)$ . Inverses are $(a,x)^{-1}:=(a^{-1},a(x))$ . The isomorphism classes of objects in $\pi _1(X)$ form a set $C(X)$ , consisting of the $G(X)$ -orbits of X. Any morphism $f\colon X\rightarrow Y$ of cycle sets induces a map $C(f)\colon C(X)\rightarrow C(Y)$ . If f is surjective, it gives rise to a commutative diagram

with a surjective group homomorphism $G(f)$ . We call $f\colon X\twoheadrightarrow Y$ a covering [Reference Rump26] if $C(f)$ and $G(f)$ are bijective. By [Reference Rump26, Theorem 6], any cycle set X admits a covering $\widetilde {X}\twoheadrightarrow X$ which is universal in the sense that it factors uniquely though each covering $Y\twoheadrightarrow X$ , so that $\widetilde {X}\rightarrow Y$ is again a covering. Cycle sets X with no proper coverings $Y\twoheadrightarrow X$ are characterized by their fundamental groupoid which is skeletal.

A cycle set X is said to be indecomposable if $G(X)$ acts transitively on X, that is, $|C(X)|=1$ . Then, $\widetilde {X}$ is indecomposable, too. Thus, $G(X)=G(\widetilde {X})$ acts freely and transitively on $\widetilde {X}$ . Being quite analogous to simply connected topological spaces, such cycle sets are called uniconnected [Reference Rump27].

3 Bracial self-maps of groups

Let $\varepsilon \colon G\rightarrow G$ be a self-map of a group G. We define the kernel of $\varepsilon $ to be the subgroup

$$ \begin{align*}\mbox{Ker}\,\varepsilon:=\{h\in G\:|\;\forall\,g\in G\colon\varepsilon(hg)= \varepsilon(g)\}.\end{align*} $$

For a group homomorphism $\varepsilon $ , this concept coincides with the usual one. Note, however, that $\mbox {Ker}\,\varepsilon $ need not be normal for an arbitrary self-map $\varepsilon $ .

Definition 3.1 We call a cycle set structure $(B;\odot )$ on a group B bracial if it satisfies $a\odot b=b(a\odot 1)$ for all $a,b\in B$ .

Thus, B is uniquely determined by its group structure and the self-map

$$ \begin{align*}\varepsilon(a):=(a\odot 1)^{-1}.\end{align*} $$

We call $\varepsilon $ the structure map and $\varepsilon (B)$ the image of B. The kernel $\mbox {Ker}\,B:=\mbox {Ker}\,\varepsilon $ of $\varepsilon $ will be called the kernel of B. In terms of $\varepsilon $ , we have

Proposition 1 A self-map $\varepsilon \colon B\rightarrow B$ of a group B defines a bracial cycle set if and only if it satisfies the equation

(3.1) $$ \begin{align} \varepsilon\bigl(a\varepsilon(b)^{-1}\bigr)\varepsilon(b)= \varepsilon\bigl(b\varepsilon(a)^{-1}\bigr)\varepsilon(a). \end{align} $$

Proof. For $a,b,c\in B$ , we have

$$\begin{align*} (a\odot b)\odot(a\odot c)&=b\varepsilon(a)^{-1}\odot c\varepsilon(a)^{-1}=c\varepsilon(a)^{-1}\varepsilon(b\varepsilon(a)^{-1})^{-1}\\&= c(\varepsilon\bigl(b\varepsilon(a)^{-1}\bigr)\varepsilon(a))^{-1}.\end{align*}$$

Equation (3.1) says that this expression is symmetric in a and b.▪

Consider the subgroup $\langle \varepsilon (B)\rangle $ of B generated by the image of B. For $a,b\in B$ , we have $a\odot b=b(a\odot 1)=b\varepsilon (a)^{-1}$ . Therefore, the right regular representation of B induces a group isomorphism

(3.2) $$ \begin{align} \pi\colon\langle\varepsilon(B)\rangle\mathop{\longrightarrow}\limits^{\sim}\, G(B) \end{align} $$

onto the permutation group of B, which maps $a\in \langle \varepsilon (B)\rangle $ to the right multiplication $b\mapsto ba^{-1}$ . On the other hand, we have a map

$$ \begin{align*}\sigma\colon B\rightarrow G(B)\end{align*} $$

with $\sigma (a)(b):=a\odot b$ . So the isomorphism (3.2) yields a factorization

(3.3)

Theorem 3.1 Let B be a bracial cycle set. Then, $G(B)$ acts freely on B. The subset $\langle \varepsilon (B)\rangle $ of B is a uniconnected subcycle set. The left cosets of $\langle \varepsilon (B)\rangle $ are the $G(B)$ -orbits of B.

Proof. Assume that $\alpha (b)=b$ for some $\alpha \in G(B)$ and $b\in B$ . Then, $b=b\pi ^{-1}(\alpha )^{-1}$ , which yields $\pi ^{-1}(\alpha )=1$ , that is, $\alpha = 1_B$ . For $a\in \langle \varepsilon (B)\rangle $ and $b\in B$ , we have $b\odot a =a\varepsilon (b)^{-1} \in \langle \varepsilon (B)\rangle $ . Thus, $\langle \varepsilon (B)\rangle $ is a subcycle set. By induction, the equation $b\odot a=a\varepsilon (b)^{-1}$ shows that the cycle set $\langle \varepsilon (B)\rangle $ is indecomposable, hence uniconnected. Moreover, the equation implies that the left cosets of $\langle \varepsilon (B)\rangle $ are the $G(B)$ -orbits of B. ▪

As a consequence, we get a classification of uniconnected cycle sets.

Theorem 3.2 A bracial cycle set B is uniconnected if and only if $\langle \varepsilon (B) \rangle =B$ . Conversely, every uniconnected cycle set B is bracial and satisfies $\langle \varepsilon (B)\rangle =B$ .

Proof. The first statement follows by Theorem 3.1. Conversely, let $(B;\odot )$ be any uniconnected cycle set. Choose an element $1\in B$ . Then, there is a bijection $\gamma \colon G(B)\mathop{\longrightarrow}\limits^{\sim}\, B$ with $\gamma (\alpha ):=\alpha (1)$ . So the map $\sigma \colon B\rightarrow G(B)$ with $\sigma (a)(b)=a\odot b$ satisfies $\sigma (a)=\gamma ^{-1}(a\odot 1)$ . Define a multiplication in B as follows:

(3.4) $$ \begin{align} ab:=\gamma^{-1}(b)(a). \end{align} $$

For given $a,b\in B$ , consider the automorphisms $\alpha :=\gamma ^{-1}(a)$ and $\beta :=\gamma ^{-1}(b)$ in $G(B)$ . Because $\gamma (\alpha \beta )=\alpha \gamma (\beta )$ , we have $\gamma \bigl (\gamma ^{-1}(b)\gamma ^{-1}(a)\bigr )= \gamma ^{-1}(b)(a)=ab$ , which yields $\gamma ^{-1}(ab)=\gamma ^{-1}(b) \gamma ^{-1}(a)$ . Thus, equation (3.4) defines a group structure on B with an isomorphism

$$ \begin{align*}\gamma\colon G(B)^{\text{op}}\mathop{\longrightarrow}\limits^{\sim}\, B.\end{align*} $$

In particular, we obtain $\langle \varepsilon (B)\rangle =B$ . For $a,b\in B$ , equation (3.4) gives $a\odot b=\sigma (a)(b)=\gamma ^{-1}(a\odot 1) (b)=b(a\odot 1)$ . Whence B is a bracial cycle set. ▪

In [Reference Rump27, Theorem 2], nondegenerate uniconnected cycle sets are classified in terms of braces with a transitive cycle base. The following example shows that uniconnected cycle sets need not be nondegenerate.

Example 1 Let G be a right $\ell $ -group [Reference Rump23], that is, a group with a lattice order, satisfying

$$ \begin{align*}a\leqslant b\;\Longrightarrow\; ac\leqslant bc\end{align*} $$

for all $a,b,c\in G$ . Consider the self-map $\varepsilon \colon G\rightarrow G$ given by $\varepsilon (a):=a\vee 1$ . Then, $\varepsilon (a\varepsilon (b)^{-1})\varepsilon (b)=\bigl (a(b\vee 1)^{-1}\vee 1\bigr )(b\vee 1)=a\vee b\vee 1$ , which is symmetric in a and b. Hence, $\varepsilon $ makes G into a bracial cycle set. The kernel of $\varepsilon $ is $\{1\}$ , while the image $\varepsilon (G)$ is the positive cone of G. Therefore, the corresponding cycle set $(G;\odot )$ with $a\odot b:=b (a\vee 1)^{-1}$ is uniconnected. However, $(G;\odot )$ is degenerate: For $a\geqslant 1$ , we have $a\odot a=a(a\vee 1)^{-1}=1$ . So the square map $a\mapsto a\odot a$ is not injective if $|G|\ne 1$ .

The diagram (3.3) shows that the structure map $\varepsilon $ of a bracial cycle set B gives the retraction $\varepsilon (B)$ of B. By [Reference Rump21, Section 6], this implies that the cycle set structure of B induces a well-defined binary operation on $\varepsilon (B)$ :

(3.5) $$ \begin{align} \varepsilon(a)\cdot\varepsilon(b):=\varepsilon(b\varepsilon(a)^{-1}). \end{align} $$

Indeed, $\varepsilon (b)=\varepsilon (c)$ implies that $\varepsilon (b\varepsilon (a)^{-1}) \varepsilon (a)=\varepsilon (a\varepsilon (b)^{-1})\varepsilon (b)=\varepsilon (a\varepsilon (c)^{-1}) \varepsilon (c)=\varepsilon (c\varepsilon (a)^{-1})\varepsilon (a)$ , which yields $\varepsilon (b\varepsilon (a)^{-1})=\varepsilon (c\varepsilon (a)^{-1})$ . Thus, $\varepsilon (a) \cdot \varepsilon (b)$ does not depend on the choice of b. So the operation (3.5) satisfies equation (1.1), which makes $\varepsilon (B)$ into a cycle set if B is nondegenerate. In general, however, $(\varepsilon (B);\cdot )$ is not a subcycle set of $(B;\odot )$ .

Example 2 Let C be a differential group [Reference Mac Lane20], that is, an abelian group with an endomorphism $d\colon C\rightarrow C$ satisfying $d^2=0$ . For an element $b\in C$ with $db\ne 0$ , we define $\varepsilon (x):=dx+b$ . Then, $\varepsilon (x-\varepsilon (y))+\varepsilon (y)=d(x+y)-db+2b$ . By symmetry, this gives a bracial cycle set C with $x\odot y=y-\varepsilon (x)=y-dx-b$ . Because $1-d$ is invertible, the cycle set is nondegenerate. However, $\varepsilon (C)$ is not a subcycle set. Indeed, $\varepsilon (x)\odot \varepsilon (y)=d(y-b)\notin \varepsilon (C)$ , because $db\ne 0$ . The subcycle set $\langle \varepsilon (C)\rangle = dC+\mathbb {Z} b$ is uniconnected.

Now, we turn our attention to the nondegenerate case. Note that a bracial cycle set B is nondegenerate if and only if the map $a\mapsto a \varepsilon (a)^{-1}$ is bijective.

Definition 3.2 We call a self-map $\varepsilon \colon G\rightarrow G$ of a group G right invariant if the implication

(3.6) $$ \begin{align} \varepsilon(a)=\varepsilon(b)\;\Longrightarrow\;\varepsilon(ac)=\varepsilon(bc) \end{align} $$

holds for $a,b,c\in G$ .

Let $\varepsilon \colon G\rightarrow G$ be a self-map with kernel H. Because $ac=(ab^{-1}) bc$ , condition (3.6) says that $\varepsilon (a)=\varepsilon (b)$ if and only if $ab^{-1}\in H$ . So the fibers of $\varepsilon $ are the right cosets of H. Therefore, $\varepsilon $ induces a bijection $\overline {\varepsilon }\colon G\backslash H\mathop{\longrightarrow}\limits^{\sim}\, \varepsilon (G)$ from the set $G\backslash H$ of right cosets onto the image of $\varepsilon $ . In Example 2, but not in Example 1, the map $\varepsilon $ is right invariant.

Proposition 2 Let G be a group with a right invariant self-map $\varepsilon \colon G\rightarrow G$ . Then,

(3.7) $$ \begin{align} g\cdot\varepsilon(a):=\varepsilon(ag^{-1}) \end{align} $$

(with $g,a\in G$ ) defines a transitive action of G on the image of $\varepsilon $ .

Proof. Because $\varepsilon $ is right invariant, equation (3.7) gives a well-defined map $G\times \varepsilon (G)\rightarrow \varepsilon (G)$ . Furthermore, $gh\cdot \varepsilon (a)=\varepsilon (ah^{-1}g^{-1})=g\cdot \varepsilon (ah^{-1})=g\cdot (h\cdot \varepsilon (a))$ and $1\cdot \varepsilon (a)=\varepsilon (a)$ . Thus, equation (3.7) defines an action on $\varepsilon (G)$ . Because $aG=G$ , this action is transitive. ▪

Proposition 3 The structure map $\varepsilon \colon B\rightarrow B$ of a nondegenerate uniconnected cycle set B is right invariant.

Proof. The diagram (3.3) shows that $\varepsilon (a)=\varepsilon (b)$ is equivalent to $\sigma (a)=\sigma (b)$ . Assuming this, we have to show that $\sigma (ac)= \sigma (bc)$ holds for all $c\in B$ . Because B is uniconnected, it is enough to verify the equivalence

$$ \begin{align*}\sigma(a)=\sigma(b)\;\Longleftrightarrow\;\sigma(c\odot a)=\sigma(c\odot b).\end{align*} $$

Indeed, $\sigma (a)=\sigma (b)$ implies $(c\odot a) \odot (c\odot d)=(a\odot c)\odot (a\odot d)= (b\odot c)\odot (b\odot d)=(c\odot b)\odot (c\odot d)$ for all $d\in B$ . Thus, $\sigma (c\odot a)=\sigma (c\odot b)$ . Conversely, assume that $\sigma (c\odot a)=\sigma (c\odot b)$ . Then, $(a\odot c)\odot (a\odot c)= (c\odot a)\odot (c\odot c)=(c\odot b)\odot (c\odot c)=(b\odot c)\odot (b\odot c)$ . Because B is nondegenerate, we obtain $a\odot c=b\odot c$ . As B is uniconnected, this implies that $\sigma (a)=\sigma (b)$ .▪

For a nondegenerate bracial cycle set B, the group $G(B)$ is the adjoint group of a brace. So the group isomorphism (3.2) also makes $\langle \varepsilon (B)\rangle $ into a brace $A(B)$ . Taking $\pi $ as an identification, we get a factorization

(3.8)

which identifies $\varepsilon (B)$ with the image of $\sigma $ .

Theorem 3.3 Let B be a brace, generated by $e\in B$ . Then, there is a unique self-map $\varepsilon \colon B\rightarrow B$ with $\varepsilon (0)=e$ such that $\varepsilon $ is the structure map of a nondegenerate uniconnected cycle set with group $B^\circ $ . Conversely, every nondegenerate uniconnected cycle set arises in this way.

Proof. By [Reference Rump28, Proposition 10], $X:=\{e^a\:|\:a\in B\}$ is a transitive cycle base of B. So [Reference Rump27, Theorem 2] shows that the operation

$$ \begin{align*}a\odot b:=b\circ(e^a)'\end{align*} $$

makes B into a nondegenerate uniconnected cycle set. In particular, $(B;\odot )$ is a bracial cycle set with $\varepsilon (a)=(a\odot 0)'=e^a$ . Thus, $\varepsilon (B)=X$ and $\varepsilon (0)=e$ . By Propositions 2 and 3, the structure map $\varepsilon $ is unique: $\varepsilon (a)=\varepsilon (0\circ a)= \varepsilon (0)^a=e^a$ .

Conversely, let B be a nondegenerate uniconnected cycle set. Because $\varepsilon (B)$ is the retraction of B, the map $\varepsilon \colon B\twoheadrightarrow \varepsilon (B)$ is a cycle set morphism. So the cycle set morphism (3.8) shows that $\iota $ is a morphism of cycle sets. Thus, $\varepsilon (B)$ is a cycle base of $A(X)$ . Equation (3.5) shows that it is transitive.

By equation (3.5), the action (3.7) coincides with the adjoint action in $A(B)$ . Hence, $\varepsilon (a)=a^{-1}\cdot \varepsilon (1)=\varepsilon (1)^a$ for all $a \in A(B)$ . Thus, $a\odot b=b\varepsilon (a)^{-1}=b(\varepsilon (1)^a)^{-1}$ . By [Reference Rump27, Theorem 2], the proof is complete. ▪

Corollary A uniconnected cycle set B is nondegenerate if and only if B is the adjoint group of a brace.

Proof. By [Reference Rump27, Proposition 2], a cycle set is nondegenerate if and only if it is a cycle base of a brace. ▪

References

Angiono, I., Galindo, C., and Vendramin, L., Hopf braces and Yang–Baxter operators. Proc. Amer. Math. Soc. 145(2017), no. 5, 19811995.CrossRefGoogle Scholar
Catino, F., Colazzo, I., and Stefanelli, P., On regular subgroups of the affine group. Bull. Aust. Math. Soc. 91(2015), no. 1, 7685 10.1017/S000497271400077XCrossRefGoogle Scholar
Catino, F., Colazzo, I., and Stefanelli, P., Regular subgroups of the affine group and asymmetric product of radical braces. J. Algebra 455(2016), 164182.10.1016/j.jalgebra.2016.01.038CrossRefGoogle Scholar
Catino, F. and Rizzo, R., Regular subgroups of the affine group and radical circle algebras. Bull. Aust. Math. Soc. 79(2009), no. 1, 103107.10.1017/S0004972708001068CrossRefGoogle Scholar
Cedó, F., Jespers, E., and Okniński, J., Retractability of set theoretic solutions of the Yang–Baxter equation. Adv. Math. 224(2010), no. 6, 24722484.10.1016/j.aim.2010.02.001CrossRefGoogle Scholar
Childs, L. N., Fixed-point free endomorphisms and Hopf–Galois structures. Proc. Amer. Math. Soc. 141(2013), no. 4, 12551265.10.1090/S0002-9939-2012-11418-2CrossRefGoogle Scholar
Chouraqui, F. and Godelle, E., Finite quotients of groups of I-type. Adv. Math. 258(2014), 4668.10.1016/j.aim.2014.02.009CrossRefGoogle Scholar
Dehornoy, P., Set-theoretic solutions of the Yang–Baxter equation, RC-calculus, and Garside germs. Adv. Math. 282(2015), 93127.10.1016/j.aim.2015.05.008CrossRefGoogle Scholar
Drinfeld, V. G., On some unsolved problems in quantum group theory. In: Quantum groups (Leningrad, 1990), Lecture Notes in Mathematics, 1510, Springer, Berlin, Germany, 1992, pp. 18.Google Scholar
Etingof, P., Schedler, T., and Soloviev, A., Set-theoretical solutions to the quantum Yang–Baxter equation. Duke Math. J. 100(1999), 169209.CrossRefGoogle Scholar
Featherstonhaugh, S. C., Caranti, A., and Childs, L. N., Abelian Hopf–Galois structures on prime-power Galois field extensions. Trans. Amer. Math. Soc. 364(2012), no. 7, 36753684.CrossRefGoogle Scholar
Gateva-Ivanova, T., Noetherian properties of skew-polynomial rings with binomial relations. Trans. Amer. Math. Soc. 343(1994), 203219.CrossRefGoogle Scholar
Gateva-Ivanova, T., Skew polynomial rings with binomial relations. J. Algebra 185(1996), 710753.10.1006/jabr.1996.0348CrossRefGoogle Scholar
Gateva-Ivanova, T., Quadratic algebras, Yang–Baxter equation, and Artin–Schelter regularity. Adv. Math. 230(2012), nos. 4–6, 21522175.CrossRefGoogle Scholar
Gateva-Ivanova, T. and Van den Bergh, M., Semigroups of I-type. J. Algebra 206(1998), 97112.10.1006/jabr.1997.7399CrossRefGoogle Scholar
Guarnieri, L. and Vendramin, L., Skew braces and the Yang–Baxter equation. Math. Comp. 86(2017), 25192534.CrossRefGoogle Scholar
Jacobson, N., Structure of rings. Vol. 37. American Mathematical Society Colloquium Publications, Providence, RI, 1964.Google Scholar
Lebed, V. and Vendramin, L., Homology of left non-degenerate set-theoretic solutions to the Yang–Baxter equation. Adv. Math. 304(2017), 12191261.10.1016/j.aim.2016.09.024CrossRefGoogle Scholar
Lu, J. H., Yan, M., and Zhu, Y. C., On the set-theoretical Yang–Baxter equation. Duke Math. J. 104(2000), 118.CrossRefGoogle Scholar
Mac Lane, S., Homology. Springer, Berlin, Göttingen, and Heidelberg, Germany, 1963.10.1007/978-3-642-62029-4CrossRefGoogle Scholar
Rump, W., A decomposition theorem for square-free unitary solutions of the quantum Yang–Baxter equation. Adv. Math. 193(2005), 4055.10.1016/j.aim.2004.03.019CrossRefGoogle Scholar
Rump, W., Braces, radical rings, and the quantum Yang–Baxter equation. J. Algebra 307(2007), 153170.CrossRefGoogle Scholar
Rump, W., Right $l$ groups, geometric Garside groups, and solutions of the quantum Yang–Baxter equation. J. Algebra 439(2015), 470510.10.1016/j.jalgebra.2015.04.045CrossRefGoogle Scholar
Rump, W., Quasi-linear cycle sets and the retraction problem for set-theoretic solutions of the quantum Yang–Baxter equation. Algebra Colloq. 23(2016), no. 1, 149166.CrossRefGoogle Scholar
Rump, W., Classification of cyclic braces, II. Trans. Amer. Math. Soc. 372(2019), no. 1, 305328.10.1090/tran/7569CrossRefGoogle Scholar
Rump, W., A covering theory for non-involutive set-theoretic solutions to the Yang–Baxter equation. J. Algebra 520(2019), 136170.10.1016/j.jalgebra.2018.11.007CrossRefGoogle Scholar
Rump, W., Classification of indecomposable involutive set-theoretic solutions to the Yang–Baxter equation. Forum Math. 32(2020), no. 4, 891903.CrossRefGoogle Scholar
Rump, W., One-generator braces and indecomposable set-theoretic solutions to the Yang–Baxter equation. Proc. Edinb. Math. Soc. 63(2020), 676696.10.1017/S0013091520000073CrossRefGoogle Scholar
Tate, J. and Van den Bergh, M., Homological properties of Sklyanin algebras. Invent. Math. 124(1996), 619647.10.1007/s002220050065CrossRefGoogle Scholar
Weinstein, A. and Xu, P., Classical solutions of the quantum Yang–Baxter equation. Comm. Math. Phys. 148(1992), no. 2, 309343.10.1007/BF02100863CrossRefGoogle Scholar