Volume 24, Number 2, Spring/Summer 2024, Pages 155–383 Covered by: Mathematical Reviews zbMATH (formerly Zentralblatt MATH) COBISS SCOPUS Science Citation Index-Expanded (SCIE) Web of Science ISI Alerting Service Current Contents/Physical, Chemical & Earth Sciences (CC/PC & ES) dblp computer science bibliography The University of Primorska The Society of Mathematicians, Physicists and Astronomers of Slovenia The Institute of Mathematics, Physics and Mechanics The Slovenian Discrete and Applied Mathematics Society The publication is partially supported by the Slovenian Research Agency from the Call for co-financing of scientific periodical publications. Contents On local operations that preserve symmetries and on preserving polyhedrality of maps Gunnar Brinkmann, Heidi Van den Camp . . . . . . . . . . . . . . . . . . 155 Algebraic degrees of 2-Cayley digraphs over abelian groups Yongjiang Wu, Jing Yang, Lihua Feng . . . . . . . . . . . . . . . . . . . . 187 A non-associative incidence near-ring with a generalized Möbius function John Johnson, Max Wakefield . . . . . . . . . . . . . . . . . . . . . . . . 207 Valuations and orderings on the real Weyl algebra Lara Vukšić . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231 Regular dessins with moduli fields of the form Q(ζp, p √ q) Nicolas Daire, Fumiharu Kato, Yoshiaki Uchino . . . . . . . . . . . . . . . 273 Generalized X-join of graphs and their automorphisms Javad Bagherian, Hanieh Memarzadeh . . . . . . . . . . . . . . . . . . . . 293 The automorphism group of the zero-divisor digraph of matrices over an antiring David Dolžan, Gabriel Verret . . . . . . . . . . . . . . . . . . . . . . . . . 317 Quotients of skew morphisms of cyclic groups Martin Bachratý . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327 Finite simple groups on triple systems Xiaoqin Zhan, Xuan Pang, Suyun Ding . . . . . . . . . . . . . . . . . . . 347 There is a unique crossing-minimal rectilinear drawing of K18 Bernardo M. Ábrego, Silvia Fernández–Merchant, Oswin Aichholzer, Jesús Leaños, Gelasio Salazar . . . . . . . . . . . . . . . . . . . . . . . . 355 Volume 24, Number 2, Spring/Summer 2024, Pages 155–383 xi ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.01 / 155–186 https://doi.org/10.26493/1855-3974.2749.b64 (Also available at http://amc-journal.eu) On local operations that preserve symmetries and on preserving polyhedrality of maps Gunnar Brinkmann , Heidi Van den Camp * Ghent University, Krijgslaan 281 S9, Ghent, Belgium Received 1 December 2021, accepted 8 July 2023, published online 4 September 2023 Abstract We prove that local operations that preserve all symmetries, as e.g. dual, truncation, me- dial, or join, as well as local operations that are only guaranteed to preserve all orientation- preserving symmetries, as e.g. gyro or snub, preserve the polyhedrality of simple maps. This generalizes a result by Mohar proving this for the operation dual. We give the proof based on an abstract characterization of these operations, prove that the operations are well defined, and also demonstrate the close connection between these operations and Delaney- Dress symbols. Keywords: Embedded graph, map, polyhedral embedding, operation, symmetry, tiling. Math. Subj. Class. (2020): 05C76, 05C10, 05C40, 52B05 1 Introduction Symmetry-preserving operations on polyhedra have been studied for a very long time. They were first applied in ancient Greece. Some of the Archimedean solids can be obtained from Platonic solids by applying the operation which was later called truncation by Kepler. Over the centuries, polyhedra and specific operations on them have been studied extensively [3, 11, 12, 18, 22]. However, a general definition of the concept local symmetry-preserving operation and a systematic way of describing such operations was only presented in 2017 [2]. This description covers a large class of operations on maps, including all well-known symmetry-preserving operations such as truncation, dual, or those operations known as achiral Goldberg-Coxeter operations [4, 5]. Goldberg-Coxeter operations were in fact in- troduced by Caspar and Klug [4] and can be used to construct all fullerenes or certain viruses with icosahedral symmetry. *Corresponding author. E-mail addresses: Gunnar.Brinkmann@ugent.be (Gunnar Brinkmann), heidi.vandencamp@gmail.com (Heidi Van den Camp) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 156 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 v2 v1 v0 Figure 1: On the left, the barycentric subdivision of a hexagonal face is shown. In the middle, the lsp-operation truncation is given and on the right the barycentric subdivision of the result of applying the operation. The blue shaded area shows one chamber of the original hexagon. In addition to these local symmetry-preserving operations (lsp-operations), which pre- serve all the symmetries of a map, there are also operations that are only guaranteed to pre- serve the orientation-preserving symmetries. Well-known examples of such operations are snub and gyro [23], or the chiral Goldberg-Coxeter operations. In [2], a general description of such local orientation-preserving symmetry-preserving operations (lopsp-operations) was also presented. The very general way of describing lsp- and lopsp-operations in [2] al- lows to tackle various problems from a more abstract perspective, and also allows to prove general theorems about the whole class of operations instead of considering each operation separately. In this paper we will use the new description to prove that all those operations (e.g. dual, medial, truncation, snub, . . . ) preserve polyhedrality of maps i.e., if an lsp- or lopsp-operation is applied to a simple 3-connected map of face-width at least 3, then the result is also simple and 3-connected and it has face-width at least 3. As the description in [2] was aimed at a broader audience than just mathematicians, the approach was described in a more intuitive way. In that article an operation is defined as a triangle ‘cut’ out of a simple periodic 3-connected tiling, and it is applied by gluing copies of that triangle into the barycentric subdivision of a map. Another way of looking at it is that the faces of the barycentric subdivision, which are triangles, are further subdivided into smaller triangles. This is done in a way that the subdivisions of the faces of the barycentric subdivision are identical or mirror images of each other, or – in case only orientation pre- serving symmetries must be preserved – in a way that each pair of two triangular faces of the subdivision that share the same edge as well as the same face of the map is subdivided in the same way. In the remainder of this text we will give the conditions for these subdivi- sions that guarantee that the result is the barycentric subdivision of another map – the result of the operation. An example of an lsp-operation and its application is shown in Figure 1. In this article, we will give the more direct definition based on Delaney-Dress symbols that forms the base of this approach and show the connection to the original description. We will also show that for every lsp-operation there is an equivalent lopsp-operation, i.e. a lopsp-operation that has the same result as the lsp-operation when applied to a map. In [2], it is proved that the result of applying an lsp-operation to a polyhedron – that is: a simple 3-connected map embedded in the plane [17] – is also a polyhedron. In [14] this result is also announced for all lopsp-operations. We will modify some concepts that G. Brinkmann et al.: On local operations that preserve symmetries . . . 157 are used in that paper, but due to some serious problems in that paper we will not use the results given there. Originally, lsp- as well as lopsp-operations were only defined for simple plane maps because of their origin in the study of polyhedra. However, there is no mathematical reason why these definitions should not be applied to maps with multiple edges or loops and embeddings of higher genus. The question then arises in how far we can extend the theorem for 3-connected simple plane maps to 3-connected maps of higher genus. In general, lopsp-operations do not necessarily preserve 3-connectivity for maps that are not plane. This is obvious for maps with faces of size 1 or 2, but it is also true for simple maps in general, even if we require the result to be simple. The most striking example of a local symmetry-preserving operation that can turn 3-connected maps into (even simple) maps with lower connectivity is dual. In [1] it is proven that for any k ≥ 1, there exist embeddings of k-connected simple maps M so that the dual M∗ is simple and has a 1-cut. However, even dual always preserves 3-connectivity in simple maps of face-width at least three, as proven in [20]. In Definition 4.1 and Definition 4.8 we will define ck-maps and ck-operations. A map is ck if it is k-connected, it has face-width at least k, and all of its faces have size at least k. In this paper we will prove the general Theorem 4.9 from which the following key result is a corollary. The result in [20] for dual is a special case of this result. The map O(M) is the result of applying the operation O to the map M : Corollary 1.1. Let k ∈ {1, 2, 3}. If M is a ck-map, and O is a ck-lsp- or ck-lopsp- operation, then O(M) is also ck. This theorem is most interesting and relevant for k = 3. This has two reasons. Firstly, the set of c3-operations contains all well-known and intensely studied operations. Lsp- operations that are not c3-lsp-operations were not even included in the original definition of lsp-operations [2]. Secondly, c3-maps, which are in fact simple embedded 3-connected maps of face-width at least three, have some very interesting properties. These maps are also known as polyhedral maps or polyhedral embeddings [20]. They can be defined equiv- alently as simple maps where every facial walk is a simple cycle and any two faces are either disjoint or their intersection consists of only one vertex or one edge. As the name suggests, polyhedral maps are a generalisation of polyhedra to surfaces of higher genus. It turns out that the key property that these operations preserve is not 3-connectivity but polyhedrality. This property is equivalent to being simple and 3-connected in the plane, but only in the plane. The main result of this article follows immediately from Corollary 1.1: If M is a polyhedral map and O is a c3-lsp- or c3-lopsp-operation, then O(M) is also a polyhedral map (Theorem 4.10). In Section 2 we give the definitions of the terminology we will use in this text. It starts with some basic concepts and then the definitions of lsp- and lopsp-operations are given. There is some freedom in the way that lopsp-operations are applied. However, in Section 3 we will prove that the result of applying a lopsp-operation is independent of the choices that are made in its application. Section 4 holds the main results of this paper: We prove a general result that implies that all lopsp-operations preserve polyhedrality of maps. To show that the definition of lopsp-operations we give is equivalent to the original definition in [2], we explore the strong connection between lsp- and lopsp-operations and tilings in Section 5. 158 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 2 Definitions There are many different, often equivalent, definitions of a map. A short description is that a map is a cellular decomposition of a surface into vertices (0-cells), edges (1-cells), and faces (2-cells). Perhaps more intuitively, a map is an embedding of a topological representation of a graph G onto a surface S. In this text we will only consider 2-cell embeddings, which means that all the connected components of S \G are homeomorphic to 2-dimensional disks. We will only consider oriented surfaces. What we will refer to as map is often called an oriented map in texts where more general maps are also studied. Maps are often studied from a topological point of view. To make some technical details easier to describe rigorously, we will use the combinatorial approach that is given below. This definition is equivalent to the topological ones [16, 21]. We will define a map as a graph together with a rotation system, which for every vertex imposes a cyclic rotational order on the edges incident to that vertex. A graph is a tuple (V,E) where V is the set of vertices and E is the set of edges. Every edge is incident to two vertices that are not necessarily different. If they are the same vertex, then that edge is a loop. Though we are mainly interested in graphs without loops or multiple edges, they will occur in a natural way — e.g. as tools or as the result of an operation — so that we will in general assume that the underlying graph of a map may have multiple edges and loops and explicitly restrict the class where necessary. With every edge of a graph G, we associate two oriented edges, each starting in one vertex of the edge and ending in the other. In the literature these are also called directed edges or darts. If e is one oriented edge, then e−1 is the other oriented edge associated with the same edge of G. For every vertex v of G, a cyclic order is assigned to all oriented edges starting at v. This way, every oriented edge e has a ‘successor’ σ(e). A map – also known as embedded graph or graph embedding – is a connected graph together with such a successor function σ. In a more general context, our maps could be referred to as oriented maps. As we will not consider unoriented maps we just use the term map. The vertices and edges of a map are the vertices and edges of the underlying graph. When drawing maps, the cyclic order around the vertices induced by σ corresponds to the clockwise order of edges around that vertex in the drawing. A map is simple if it has no loops and no multiple edges that are incident with the same 2 vertices. A map is k-connected if it has at least k + 1 vertices and it has no vertex-cut of fewer than k vertices. Consider three oriented edges e1, e2, and e3 incident with a vertex v. We say that e2 is between e1 and e3 if e1, e2, e3 occur in this order in the cyclic order around v, i.e. the cyclic order of edges around v is of the form (. . . e1 . . . e2 . . . e3 . . .) and not (. . . e3 . . . e2 . . . e1 . . .). We say that e and σ(e−1) form an angle in the map. A face of a map M is a cyclic sequence of oriented edges such that every two consecutive edges form an angle. We will use the term facial walk to refer to the closed walk in M corresponding to this cyclic sequence of oriented edges. This definition of face corresponds to the topological notion of a face. For a map M , with VM , EM , and FM denoting the sets of vertices, edges, and faces of M respectively, χ(M) = |VM |−|EM |+ |FM | is the Euler characteristic of M . The genus of M is defined as gen(M) = 2−χ(M)2 . If a map has genus 0, it is called plane. Note that a plane map is not the same as a planar graph. A planar graph is a graph (not embedded) that G. Brinkmann et al.: On local operations that preserve symmetries . . . 159 Figure 2: The left figure shows a map and one of its submaps with bold edges.The right figure shows the internal component of the only face of the submap that has bridges. can be embedded such that it has genus 0. A plane map is one specific genus 0 embedding of a graph. Let M be a map and G′ a subgraph of the underlying graph of M . The map M ′ which is the graph G′ with the embedding induced by that of M is called a submap of M . To formalize when a vertex or edge is ‘in’ a face of one of its submaps we will now define what a bridge for a submap M ′ of M is. There are two kinds of bridges: • If e ∈ EM \EM ′ is an edge with endpoints v, w ∈ VM ′ , then the submap with vertex set {v, w} and edge set {e} is a bridge. • Let C be a component of the submap of M induced by the vertices of M that are not in M ′, and define E′C = {e ∈ EM | e ∩ VC ̸= ∅} and V ′C = {v ∈ VM | ∃e ∈ E′C : v ∈ e}. Then the submap with vertex set V ′C and edge set E′C is a bridge. If a bridge has an edge that is between two edges e and e′ so that e−1 and e′ form an angle in a face of M ′, then the bridge is in that face. All the vertices and edges of the bridge are also said to be in the face. The boundary ∂f of a face f is the submap of M consisting of all the vertices and edges in the facial walk of f . A vertex or edge of M is in the interior of a face of M ′ if it is in that face and it is not in the boundary. If a bridge is in more than one face, we say that those faces are bridged. A face that is not bridged is called simple. Let f be a simple face of M ′. We will define the internal component of f as follows. Start with the submap N of M that consists of the boundary of f together with all bridges in f . Intuitively, we cut along the boundary of f in N in such a way that the facial walk becomes a simple cycle. More formally, we replace every vertex v of N that appears k > 1 times in the facial walk of f by k pairwise different vertices v1, ..., vk. If both oriented edges associated with an edge of M ′ appear in the facial walk, this edge is also split into two different edges between different copies of its vertices. Let (x, v) and (v, y) be the oriented edges that form the angle in M ′ at the i-th occurrence of v. Then we define the rotational order (and also the neighbours) of vi to be the same as the rotational order around v in M , but restricted to the edges between (v, x) and (v, y). Of course some vertices may be replaced by their copies. The result of this is the internal component IC(f) of f . An example of an internal component is illustrated in Figure 2. If IC(f) is plane, we call f internally plane. An important concept in the definition of lsp- and lopsp-operations is the barycentric subdivision of a map. It is obtained by subdividing every face into triangular faces, which we will call chambers. We will also use the barycentric subdivision to define contractible cycles and face-width in a combinatorial way. 160 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 Figure 3: A face in a map M and the corresponding part of BM . Edges of colour 1 are dashed and edges of colour 2 are dotted. The barycentric subdivision BM of a map M is a map that has a unique vertex for every vertex, for every edge and for every face of M . We always assume that BM comes with the natural vertex-colouring that assigns colours 0, 1, and 2 to vertices that correspond to vertices, edges, and faces of M respectively. These colours correspond to their topological dimension. There are edges between vertices of colour 0 and 1 if the corresponding vertex and edge are incident. There are edges between vertices of colour 0 or 1 and colour 2 if the corresponding vertex or edge appears in the boundary of the corresponding face. There are no edges between vertices of the same colour. For i ∈ {0, 1, 2}, an edge is of colour i if it is not incident with a vertex of colour i. We will also refer to vertices and edges of colour i as i-vertices and i-edges. The rotational order of the edges adjacent to a vertex of colour 2 follows the order of the vertices and edges in the corresponding facial walk of M , and similarly for vertices of colour 0 and 1. This is illustrated in Figure 3. Every face of BM is a triangle. Note that in every figure in this text, colours are represented by colours in the order rgb, that is: a red is colour 0, green is colour 1, and black is colour 2. The edges of colour 1 are dashed and the edges of colour 2 are dotted, so that when looking at the figures printed in black and white it should still be clear which edges have which colour. With this rotation system, a short calculation of the Euler characteristic shows that gen(BM ) = gen(M). If x is a face, edge or vertex of M , then to keep notation simple we will also write x for the corresponding vertex of BM . Every face of BM is a triangle, with exactly one vertex and one edge of each colour. We call such a triangle a chamber. Two chambers are adjacent if they share an edge. In the literature, chambers are also called flags. The flag graph of M is the dual of BM , i.e. it is the 3-regular graph that has the chambers as its vertices, and there is an edge between two vertices if their corresponding chambers are adjacent. In some papers flags are defined as triples (v, e, f) where v, e, and f are respectively a vertex, an edge, and a face such that v is a vertex of e and v and e are in face f [7, 19]. We cannot use that approach here because with our general definition of a map there is no 1-to-1 correspondence between chambers and triples (v, e, f). For example, an edge can have the same face on both sides so that there are multiple chambers with the same vertices. Lemma 2.1. A map M , vertex-coloured with colours 0,1, and 2 is the barycentric subdi- vision of another map if and only if: (i) Every face of M is a triangle. G. Brinkmann et al.: On local operations that preserve symmetries . . . 161 (ii) There are no edges between vertices of the same colour. (iii) Every vertex of colour 1 has degree 4. Proof. Let VG and EG be the sets of vertices of M with colours 0 and 1 respectively. Conditions (i) and (ii) imply that every face has exactly one vertex of each colour. It now follows from (iii) that a vertex e ∈ EG of colour 1 has two neighbours in VG and two neighbours f and g of colour 2. This induces an incidence relation on the vertex set VG and the edge set EG that defines a graph G. The rotation system of M induces a rotation system on G. Let N be the map that consists of G with this rotation system. It is not difficult to check that M = BN . The double chamber map DM of a map M is the submap of BM that only contains the edges of colours 1 and 2. A double chamber of a map M is a face in DM . Every double chamber has length four: two (in case of no loops different) vertices of colour 0, one of colour 1, and one of colour 2. Two double chambers are adjacent if they share a 1-edge or two 2-edges. In [2], lsp- and lopsp-operations are – following Goldberg [15] – defined in a geometric way as triangles ‘cut’ out of the barycentric subdivision of a 3-connected tiling of the plane, such that in case of lsp-operations the sides of the triangle are on symmetry axes of the tiling. In this article we give purely combinatorial definitions of lsp- and lopsp- operations, similar to [14] and [13]. The definitions given here are equivalent to those in [2] when restricted to what we will later call c3-operations. The equivalence can be seen by applying operations as defined here to some special periodic tiling, but readers who want to see the equivalence already before starting on the main results of this paper and who want to have a deeper insight into the relation of operations and periodic tilings encoded by Delaney-Dress symbols, can find a direct proof without applications of the operations in Section 5. Definition 2.2. Let O be a 2-connected plane map with vertex set V , together with a colouring c : V → {0, 1, 2}. One of the faces is called the outer face. This face contains three special vertices marked as v0, v1, and v2. We say that a vertex v has colour i if c(v) = i. This 3-coloured map O is a local symmetry preserving operation, lsp-operation for short, if the following properties hold: (1) Every inner face — i.e. every face that is not the outer face — is a triangle. (2) There are no edges between vertices of the same colour, i.e. the colouring is proper. (3) For each vertex that is not in the outer face: c(v) = 1 ⇒ deg(v) = 4 For each vertex v in the outer face, different from v0, v1, and v2: c(v) = 1 ⇒ deg(v) = 3 and c(v0), c(v2) ̸= 1 162 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 c(v1) = 1 ⇒ deg(v1) = 2 An example of an lsp-operation is shown in the middle of Figure 1. Just like for barycentric subdivisions we say that an edge is of colour i if it is not incident to a vertex of colour i. This is well-defined because of the second property. Every inner face has exactly one vertex and one edge of each colour. We will refer to these triangular faces as chambers. In the original paper [2] only operations that preserve 3-connectivity of polyhedra were discussed, so the result of the operation also had to have only vertices of degree at least 3. In [13] operations were also discussed that produce maps with 1- or 2-cuts, but the restriction that vertices in the result should have degree at least 3 was kept. Our definition of lsp-operations is even more general. With this definition, the result of applying an lsp- operation may have vertices of degree 1 or 2. Application of an lsp-operation: Let O be an lsp-operation and let M be a map. The operation is applied to M by first replacing for i ∈ {0, 1, 2} the i-edges of BM by copies of the part of the boundary of the outer face of O between vj and vk with i ̸= j, k. The copy of vj is identified with the j- vertex and the copy of vk with the k-vertex. Then — depending on the orientation — either a copy of O or a copy of the mirror image of O — which has the same underlying graph as O but the rotation system is the inverse of that of O — is glued into every face of the modified BM . Note that chambers of BM sharing an edge have different orientations. The boundary vertices are identified with their copies. This results in a 3-coloured triangulation. An example of the gluing — restricted to a single face — is given in Figure 1. With Lemma 2.1 and Definition 2.2 it follows that this triangulation is the barycentric subdivision of a map O(M), the result of applying O to M . As any symmetry group acts on the chamber system, lsp-operations preserve all the symmetries of a map. New symmetries can also occur. However, all known examples of 3-connected maps where lsp-operations can increase symmetry are maps of genus at least 1 or they are self-dual. It is an open question whether lsp-operations can increase symmetry in plane 3-connected maps (polyhedra) that are not self-dual. There are also interesting operations such as gyro and snub that are only guaranteed to preserve the orientation-preserving symmetries of maps. These cannot be described by lsp- operations. In the supplementary material of [2] and in [14], local orientation-preserving symmetry-preserving operations (lopsp-operations) are defined similarly to lsp-operations. The most important difference is that here the decoration is glued into double chambers instead of chambers. As with lsp-operations, we will give a more explicit definition of lopsp-operations that is not directly based on tilings. There are some problems that arise in the original definition of lopsp-operations that do not appear for lsp-operations. With the original definition, it is possible to cut different patches out of a tiling that describe the same operation and must be shown to have the same result. That is why we define a lopsp-operation as a plane triangulation, similar to [14], and not as a quadrangle that we can glue directly into double chambers. Although this simplifies the definition of a lopsp-operation, the same problem comes back when it is described how the operation is applied. Definition 2.3. Let O be a 2-connected plane map with vertex set V , together with a colouring c : V → {0, 1, 2} and three special vertices marked as v0, v1, and v2. We say G. Brinkmann et al.: On local operations that preserve symmetries . . . 163 v0 v1 v2 v2 v1 v0,L v0,R Figure 4: On the left, the lopsp-operation gyro is shown. The thick edges are the edges of the path P . On the right the corresponding double chamber patch OP is drawn. that a vertex is of colour i if c(v) = i. The map O is a local orientation-preserving symmetry-preserving operation, lopsp-operation for short, if the following properties hold: (1) Every face is a triangle. (2) There are no edges between vertices of the same colour, i.e. the colouring is proper. (3) For each vertex v different from v0, v1, and v2: c(v) = 1 ⇒ deg(v) = 4 and c(v0), c(v2) ̸= 1 c(v1) = 1 ⇒ deg(v1) = 2 Again we say that an edge has colour i if it is not incident to a vertex of colour i and this is well-defined because of the second property. Note that the edges incident with a vertex have two different colours, and as every face is a triangle, these colours appear alternatingly in the cyclic order around the vertex. The requirement that O is 2-connected is mentioned in the beginning, but would in fact also follow from the other conditions. Again every face has exactly one vertex and one edge of each colour and will be referred to as a chamber. The dual of O will be referred to as the flag structure of O. Application of a lopsp-operation: For vertices v, v′ in a path P we write Pv,v′ for the subpath of P from v to v′. As lopsp-operations are 2-connected, due to Menger’s theorem there are two paths, one from v0 to v1 and one from v0 to v2 that have only v0 in common. These paths together form a longer path P from v1 to v2 through v0. As a submap of O, P has a single face. In this facial walk only v1 and v2 occur once and all other vertices of P occur twice. We say that such a path P is a cut-path of O. Consider the internal component of the only face of 164 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 v2 v1 v0,L v0,R Figure 5: On the left the barycentric subdivision of a hexagonal face is shown. In the middle, a double chamber patch OP of the operation gyro is drawn, and the right image shows the part of BOP (M) corresponding to the hexagonal face. The blue shaded area shows one double chamber. submap P . This is the double chamber patch OP . It can be drawn in the plane, so that the two copies of P form the boundary of the outer face. Figure 4 shows this for the operation gyro. The result of the cutting is a 4-gon with corner vertices v1, v2, and two copies of v0, which we will denote as v0,L and v0,R. The flag structure of OP is the flag structure of O where the edges corresponding to edges of P are removed. The lopsp-operation is now applied by first replacing the edges of a double chamber map DM to form the map DM,P . An edge of colour 2 is replaced by a copy of Pv0,v1 and an edge of colour 1 is replaced by a copy of Pv0,v2 in a way that for i ∈ {0, 1, 2} a copy of vi is identified with a vertex of colour i. Gluing copies of the double chamber patch OP into the faces of DM,P — identify- ing corresponding vertices in DM,P and the copies of double chamber patches — gives a coloured map BOP (M). Note that the orientation inside a double chamber fixes how the different copies of v0 have to be identified. Unlike with lsp-operations, we do not use mirrored copies of O. Figure 5 gives an example — restricted to one face — of this gluing. A side of a double chamber is a path in the boundary of the corresponding face of DM,P that is a copy of the path in OP between v2 and v0,L, between v2 and v0,R, or between v0,L and v0,R. A side is a 1-side if it is between copies of v0 and v2 and it is a 2-side if it is between two copies of v0. Lemma 2.4. Let M be a map and let O be a lopsp-operation with a cut-path P . The 3-coloured map BOP (M) is the barycentric subdivision of a connected map. Proof. This follows immediately from Lemma 2.1. As lsp-operations preserve all symmetries of a map, they also preserve the orientation- preserving symmetries, so one would expect that for every lsp-operation, there is a lopsp- operation that has the same result when applied to any map. This observation allows to prove some properties of the result of applying lsp- or lopsp-operations only for lopsp- operations. The result for lsp-operations can then be deduced from the corresponding lopsp-operation. Such an equivalent lopsp-operation can be obtained in the following way: Let O be an lsp-operation, and let c be the boundary of the outer face of O. Let Olopsp be the map obtained by gluing a mirrored copy of the inner face of c into the outer face, identifying the vertices on c with their copies. The vertices v0, v1, and v2 of O are also the vertices v0, v1, and v2 of Olopsp. G. Brinkmann et al.: On local operations that preserve symmetries . . . 165 Lemma 2.5. If O is an lsp-operation, then Olopsp is a lopsp-operation, and O(M) = Olopsp(M) for any map M . Proof. Olopsp is obviously a triangulation of the disc and there are no edges between ver- tices with the same colour. As Olopsp consists of two copies of O, glued along the boundary c, we can associate a unique vertex o(x) of O with every vertex x of Olopsp. The degree of x in Olopsp is given by deg(x) = { deg(o(x)) if o(x) is not in c 2deg(o(x))− 2 if o(x) is in c . From the degree restrictions for lsp-operations we can now deduce the degree restric- tions in the definition of lopsp-operations for Olopsp. It follows that Olopsp is a lopsp- operation. Choosing the cut-path in Olopsp that corresponds to the path from v1 to v2 through v0 in c for the application of Olopsp shows immediately that the results of applying O and Olopsp are isomorphic: a double chamber is filled in the same way by Olopsp as two adjacent chambers are filled by O. 3 The path invariance of lopsp-operations The cut-path chosen to apply an operation is far from unique, so there are many ways to apply a single lopsp-operation. In this section it is proved that although the ways in which the operation is applied differ, the result of applying a lopsp-operation to a map is independent of the chosen path. An essential tool in proving this are chamber flips, which simulate homotopic deformations. Definition 3.1. Let P be a directed walk in a barycentric subdivision or lopsp-operation. For any two different vertices of a chamber C, there are two different simple paths P0, P1 between these vertices in the boundary of C. If for i ∈ {0, 1} path Pi occurs at a certain position in P , then a chamber flip of C (at this position) is the operation of replacing Pi by P1−i. As a first tool we will discuss transformations of one path into another: Lemma 3.2. Let P, P ′ be two directed paths of the form P = PsR, P ′ = PsR′ from x to y in a lopsp-operation T , so that R′R−1 is the facial walk of an internally plane face f in the submap of T consisting of the vertices and edges of P and P ′. Then there is a sequence of paths P = P0, P1, . . . , Pk = P ′ so that for 1 ≤ i ≤ k path Pi is obtained from Pi−1 by a chamber flip and every vertex of Pi is in Ps or in the boundary or the interior of f . As chamber flips can be reversed, the same is true with the role of P and P ′ interchanged. Proof. We will prove this by induction on the number |C | with C the set of chambers of T inside f . If |C | = 1, then R and R′ are the two paths along the boundary of a chamber, so one can be transformed into the other by one chamber flip and we are done. Now assume that |C | ≥ 2. We prove that there are at least two chambers in C that have a connected intersection with ∂f that contains at least one edge: Let Ff be the dual of T restricted to 166 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 C and without edges that correspond to edges in ∂f . If T is the barycentric subdivision of a map then Ff is part of the flag graph of that map. There are at least two chambers in C that contain an edge of ∂f . Assume that there is a chamber C such that C ∩ ∂f is disconnected. This chamber C splits the set C into two parts, i.e. the vertex corresponding to C is a cut-vertex of Ff . In each component of Ff \ C there is at least one chamber that shares an edge with ∂f . Let C0 be a chamber that contains an edge of ∂f that has the largest distance dmax to C along a path in Ff . If this chamber has a disconnected intersection with ∂f , then its corresponding vertex is a cut-vertex of Ff . This implies that there is a chamber that shares an edge with ∂f and has a larger distance to C than dmax, which is in contradiction with the maximality of dmax. Repeating this argument for the other component of Ff \ C, it follows that in each of the two components there is a chamber that has a connected intersection with the facial walk ∂f that contains at least one edge. Assume that one of these two chambers intersects ∂f in a single edge or in two edges of P or of P ′. Then we can do a chamber flip to obtain either a path P1 or Pk−1, so that we can apply induction to P1, P ′ or P, Pk−1 and use that each chamber flip can be undone by a reverse chamber flip. If the intersection of neither of the two chambers with ∂f is one or two edges of P or P ′, then both intersections consist of one edge of P and one edge of P ′. For one of the chambers, the shared vertex of those edges is the first vertex of R and R′. Applying a chamber flip replacing the edge of P , we get a path P1 to which we can apply induction. Lemma 3.3. Let Q,Q′ be two directed paths from x to y in a lopsp-operation, and z a vertex not contained in either of the paths. Then there is a sequence of paths Q = Q0, Q1, . . . , Qk = Q′ from x to y so that for 1 ≤ i ≤ k the path Qi is obtained from Qi−1 by a chamber flip and none of the paths contain z. Proof. We will prove this by backwards induction on the number n of edges in the begin- ning of Q that Q and Q′ have in common. Remember that for vertices v, v′ in a path Q we write Qv,v′ for the subpath of Q from v to v′. If n = |Q′|, then Q = Q′, so assume that n < |Q′| and that the assumption is true for n′ > n. Then there is a first vertex a in Q that is incident with an edge that is in Q′ but not in Q. Let b be the next vertex after a in Q′ that Q′ shares with Q. We will show that Q can be transformed to Q′x,aQ ′ a,bQb,y in the described way, so that we can apply induction to transform Q′x,aQ ′ a,bQb,y into Q ′. Let c be the cycle Qa,b ∪Q′a,b. We call the face of c containing z the exterior. Note that neither Q′x,a = Qx,a nor Qb,y intersects c in a vertex other than a or b. There are four possibilities for the position of Qx,a and Qb,y . These are depicted in Figure 6. If Qx,a or Qb,y are in the interior of c, we use them as part of the face boundary when applying Lemma 3.2, otherwise we do not. As Lemma 3.2 already allows to consider paths that start with a common part outside the face, we can choose P, P ′ from Lemma 3.2 in the following way: Qb,y outside: Choose P = Qx,aQa,b, P ′ = Q′x,aQ ′ a,b. Qb,y inside: Choose P = Qx,aQa,bQb,y , P ′ = Q′x,aQ ′ a,bQb,y . G. Brinkmann et al.: On local operations that preserve symmetries . . . 167 Q′a,b Qx,a x a b y z Qb,y Qa,b Q′a,b Qa,x x a b y z Qb,y Qa,b Q′a,b Qx,a Qb,y x a b y z Qa,b Q′a,b Qx,a Qb,y x a b y z Qa,b Figure 6: The four different cases in the proof of Lemma 3.3 are shown here. The shaded area represents the interior. Note that in case Qx,a is outside c it forms the Ps from Lemma 3.2, otherwise Ps consists of a single vertex. In each case Lemma 3.2 can be applied to prove that Q can be transformed to Q′x,aQ ′ a,bQb,y in the described way, and as the beginning of Q ′ x,aQ ′ a,bQb,y has more than n edges in common with Q′, we can apply reverse induction. Let M be a map, O a lopsp-operation with cut-path P and OP the corresponding dou- ble chamber patch. Recall that BOP (M) is obtained by gluing copies of OP into DM . Therefore every vertex v in BOP (M) is in at least one copy of OP . If v is in more than one copy, v corresponds to the same vertex of O in each of these copies. Similarly, every edge or face of BOP (M) also corresponds to exactly one edge or face of O respectively. This allows us to define a surjective mapping πP , that maps every vertex, edge, and face of BOP (M) to its corresponding vertex, edge, or face of O. The mapping πP is not a bijection, but we can define a kind of inverse function π−1P . It maps a set X of vertices, edges or faces in O to the set of all the vertices, edges or faces in BOP (M) whose image under πP is in X . If we apply π−1P to a single vertex, edge or face x of O, we will often write π−1P (x) instead of π −1 P ({x}). For submaps M ′ of O the image π−1P (M ′) is a subset of vertices and edges of BOP (M). If these form a connected graph, we interpret it as a map with the embedding induced by BOP (M). The definition of BOP (M) depends on P . We will now prove that the result of an operation is independent of P , so that we can define O(M) for a lopsp-operation O. Theorem 3.4. Let O be a lopsp-operation and let P and Q be two cut-paths in O. Let M be a map. Then BOP (M) ∼= BOQ(M). Proof. The idea of this proof is as follows. We define a submap BOP (M)|Q of BOP (M) and prove that the underlying graph of this map is isomorphic as a graph to DM,Q. Then we prove that they are also isomorphic as maps, and that the internal component of each face of BOP (M)|Q is isomorphic to OQ. It follows that BOP (M) is isomorphic to BOQ(M). 168 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 Let e be an edge of DM , and let j be 1 if e has colour 2 and 2 if e has colour 1. Let P e be the copy of Pvj ,v0 in BOP (M) that replaced e. By Lemma 3.3 there is a series of paths Pvj ,v0 = P0, . . . , Pk = Qvj ,v0 from vj to v0 in O, so that the path Pi+1 is obtained from Pi by a chamber flip of a chamber Ci and none of v0, v1, v2 occur as interior points of any of the paths. We define a sequence of paths P e = P e0 , . . . , P e k in BOP (M) with πP (P e i ) = Pi for 0 ≤ i ≤ k. The path P ei+1 will be obtained from P ei by applying a chamber flip to a chamber C ∈ π−1P (Ci). The chamber flips in O on the paths Pi replace subpaths of one or two edges. In case of one edge it is clear that a corresponding chamber flip can be performed on P ei in BOP (M). In case of two edges, we have to prove that the two corresponding edges of P ei are also contained in the same chamber. As P e i is a path, the two edges share one of their vertices, say v. By definition of the paths Pi we get that πP (v) /∈ {v0, v1, v2}. For such a vertex v it is true that if e1, e2, . . . , ek is the cyclic order of edges around v, then πP (e1), πP (e2), . . . , πP (ek) is the cyclic order of the edges around the vertex πP (v). If a chamber flip is applied to the edges πP (ej) and πP (ej+1) to go from Pi to Pi+1, then we can apply a chamber flip to the edges ej and ej+1 to go from P ei to a new path P e i+1. Thus our sequence of paths P e = P e0 , . . . , P e k in BOP (M) with πP (P e i ) = Pi is defined for 0 ≤ i ≤ k and πP (P ek ) = Qvj ,v0 . We denote P ek as Qe. Note that Qe is isomorphic to Qvj ,v0 , not to Q. Let BOP (M)|Q be the map consisting of all the vertices and edges of BOP (M) contained in Qe for some edge e. With the rotational orders induced by BOP (M) we have that BOP (M)|Q is a submap of BOP (M). First we prove that as (non-embedded) graphs, BOP (M)|Q and DM,Q are isomorphic. Two paths Qe and Qe ′ can only intersect in their endpoints: Every other vertex v of Qe and Qe ′ satisfies πP (v) ̸∈ {v0, v1, v2}, which implies that v has only two incident edges that are mapped to edges in Q by πP . It follows that two paths of the form Qe are either disjoint — except possibly for their endpoints — or identical. We prove by induction that P ei and P e′ i are disjoint (except for their endpoints) for all 0 ≤ i ≤ k and edges e and e′ in DM . If e and e′ are edges of a different colour this is trivial as at least one of their endpoints is different. Assume that e and e′ have the same colour. By our previous argument it suffices to show that their first edge is different. For i = 0 this is clear. Assume that it is true for i − 1. Let εi and ε′i be the first edges of P ei and P e ′ i respectively. We can assume that they are both incident with the same vertex x ∈ π−1P ({v0, v1, v2}). The paths P ei and P e′ i are obtained from P e i−1 and P e′ i−1 by one chamber flip for each path. Either εi = εi−1 and ε′i = ε ′ i−1, or the chamber flips replace εi−1 and ε ′ i−1 by both their previous edges or both their next edges in the rotational order around x. As εi−1 and ε′i−1 are different edges, εi and ε′i are also different edges, which proves our statement. It follows that BOP (M)|Q and DM,Q are isomorphic as graphs. Next we prove that they are also isomorphic as maps. Let m denote the total number of chamber flips necessary to transform first Pv1,v0 to Qv1,v0 and then Pv2,v0 to Qv2,v0 . With every face (that is: double chamber) D of DM and 0 ≤ i ≤ m we can now associate a closed walk Wi that consists of the four paths P e1i , P e2 i , P e3 i , P e4 i in BOP (M) where e1, . . . , e4 are the four edges of D, in the same order as they appear in D. Claim: BOP (M)|Q is a submap of BOP (M) that is isomorphic as a map to DM,Q and the internal component of each face is isomorphic to OQ. Let C be the set of all chambers in BOP (M), and let n be the number of chambers in O. We will define functions αi : C → Z (0 ≤ i ≤ m) with the following properties: G. Brinkmann et al.: On local operations that preserve symmetries . . . 169 C C C C αi+1(C) = αi(C)− 1 αi+1(C) = αi(C) + 1 αi+1(C) = αi(C) + 1 αi+1(C) = αi(C)− 1 Figure 7: The evolution of α after chamber flips. The bold paths with arrows are Wi and Wi+1. (i) Let C,C ′ in BOP (M) be two adjacent chambers sharing the directed edges e and e−1, so that C is on the left of e. For e′ ∈ {e, e−1} we define ni(e′) as the number of times e′ occurs in the cyclic walk Wi. Then αi(C)− αi(C ′) = ni(e)− ni(e−1). (ii) For every chamber C in O: ∑ C′∈π−1P (C) αi(C ′) = 1 As a consequence of (ii) we have ∑ C∈C αi(C) = n. The walk W0 is an internally plane facial walk of DM,P with an internal component that is isomorphic to OP . We define α0(C) = 1 if C is a chamber on the inside of W0 and α0(C) = 0 if C is on the outside. As W0 has exactly one copy of each chamber in O inside we get (ii) for α0. As α0 only differs for neighbouring chambers if they share an edge of W0, and then in the way described by (i), we also get (i). For i > 0 we define αi inductively. Let C be the chamber of O to which a chamber flip is applied when changing Wi−1 to Wi. These chamber flips occur in two places of Wi−1, and in fact in different directions. Two chambers C−, C+ with πP (C−) = πP (C+) = C are involved, C− on the left of the cyclic walk and C+ on the right. We now define αi(C −) = αi−1(C −) − 1 and αi(C+) = αi−1(C+) + 1. This is illustrated in Figure 7. As we once add one and once subtract one for two chambers with the same image under πP , (ii) is immediate. Property (i) can be checked easily by looking at αi for C−, C+, and the neighbouring chambers sharing an edge with them. For i = 0, The function αi describes whether a chamber is inside or outside Wi. For other i this is not always the case. If Wi self-intersects the intuitive meaning of αi is less obvious. For j = 1 or j = 2, the two edges of Wi incident to the j-vertex x of D are always moved in the same direction by the chamber flips. This implies that {αi(C) | x ∈ C} is the same set for every 0 ≤ i ≤ m. As {α0(C) | x ∈ C} ⊆ {0, 1} — it can be {1} if 170 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 M has a loop — every chamber that contains x is mapped to 0 or 1 by αm. The degree in Wm of every vertex that is not in π−1P ({v0, v1, v2}) is two, so we can follow Wm from v1 and from v2 to the copies of v0 to conclude that for each edge of Wm, the two chambers C and C ′ containing it have αm(C) ∈ {0, 1} and αm(C ′) ∈ {0, 1}. The αm values of two adjacent chambers can only differ if their shared edge is in Wm, so 0 and 1 are the only values of αm. Note that this argument only works because every vertex of Wm that is not in π−1P ({v0, v1, v2}) has degree 2 in Wm. For 0 < i < m this is not necessarily the case, and for those values of i the mapping αi may have values different from 0 and 1. Consider the submap Fm of the dual of BOP (M) — i.e. the flag graph of the map N such that BN = BOP (M) — induced by the chambers C with αm(C) = 1, where edges in that map corresponding to edges in Wm are removed. By (ii) it follows that for every chamber CO in O there is exactly one vertex in Fm. For every edge in the flag structure of OQ there is exactly one edge in Fm, as by (i) adjacent chambers have the same value under αm if their shared edge is not in Wm. In fact, these are all the edges of Fm: The maximum degree of a vertex in Fm is 3, as a chamber is adjacent to three others. Let k be the number of edges in Q. As there are 2k edges in Wm, we get 2·|EFm | = ∑ v∈Fm deg(v) ≤ 3n−2k. We also have 2|EFm | ≥ 2|EOQ | = 3n− 2k and thus |EFm | = |EOQ |. It follows that the flag structure of OQ is isomorphic to Fm. As there are no edges in the flag structure of OQ that correspond to edges of Q, the walk Wm is the facial walk of a face of BOP (M)|Q. It follows that BOP (M)|Q and DM,Q are isomorphic maps and the internal component of each face of BOP (M)|Q is isomorphic to OQ. Definition 3.5. Let O be a lopsp-operation and let M be a map. Choose any cut-path P in O. The result O(M) of applying O to M is the map with barycentric subdivision BOP (M). By Lemma 2.4 and Theorem 3.4, O(M) is well-defined and independent of the chosen path. We can also define the map π := πP as it is independent of the chosen path. 4 The effect of lsp- and lopsp-operations on polyhedrality Polyhedral maps are simple maps that are 3-connected and have ‘face-width’ at least three. The face-width (or representativity) of a map is a measure of ‘local planarity’. Embeddings of high face-width share certain properties with plane maps. We will define face-width in a combinatorial way, using barycentric subdivisions. It is not difficult to prove that the definition given here is equivalent to the definition in e.g. [20]. A cycle in a map M is contractible if – as a submap of M – it has a simple internally plane face. The face-width of a map M , denoted fw(M), is the minimal length of a non- contractible cycle in BM , divided by two. If M has no non-contractible cycles, i.e. M is plane, then we define fw(M) = ∞. Definition 4.1. For k ≥ 1 we define a map M to be ck if: • M has no cut-sets with fewer than k vertices • fw(M) ≥ k • The size of every face of M is at least k • The degree of every vertex of M is at least k G. Brinkmann et al.: On local operations that preserve symmetries . . . 171 The condition that neither cuts with fewer than k vertices nor vertices with degree smaller than k may be present instead of just requiring the map to be k-connected is chosen in order to deal with small boundary cases. For example, a cycle is 2-connected, but its dual is a map with only two vertices so it is not 2-connected. Both a cycle and its dual are c2. A polyhedral map is a simple, 3-connected map that has face-width at least three. Lemma 4.2. A map is c3 if and only if it is polyhedral. Proof. It suffices to prove that every c3-map is simple and has at least four vertices. The rest of the statement is trivial when the definitions are written out. Let M be a c3-map. Facial loops and facial 2-cycles are excluded by the restrictions on face sizes and non-facial loops and non-facial 2-cycles imply either smaller cuts or a smaller face-width. Therefore M is simple. It has at least 4 vertices as it has minimum degree at least 3 and it is simple. The reason why the term c3 is used instead of polyhedral in this article is that many results are proven for ck maps for general k ∈ {1, 2, 3}. The following lemma characterises c2- and c3-maps by a condition based on the cham- ber system. A 4-cycle in a barycentric subdivision is called trivial if it has a face that has no vertex or only a single colour-1 vertex in its interior. Lemma 4.3. Let M be a map. (i) M is c2 if and only if BM has no cycles of length 2. (ii) M is c3 if and only if M is c2, and BM has no nontrivial cycles of length 4. Proof. (i): Let M be a map and assume that M is not c2. There are four possible reasons for not being c2: the existence of a cutvertex, the existence of a facial loop (a face of size 1), the existence of a vertex of degree 1, or the existence of a non-contractible 2-cycle in BM . The last three immediately imply the existence of a 2-cycle in BM , so assume that M has a cutvertex v. If there is a loop in M , then there is a 2-cycle in BM , so assume that M has no loops. Then vertex v has neighbours x and y in different components such that y follows x in the rotational order around v. The facial walk (x, v), (v, y), (y, w1), . . . , (wk, x) of the face f containing this angle must also contain v as one of the wi as otherwise part of the facial walk would be a path from x to y in M \ {v}. This implies that in the barycentric subdivision there are 2 edges between v and the vertex corresponding to f — a 2-cycle. Conversely, assume that there is a 2-cycle c in BM . If there is a 0- or 2-vertex of degree two in BM then there is a vertex of degree 1 or a face of size 1 in M , so we can assume that every vertex of BM has degree at least four. We can also assume that every cycle of length 2 in BM is contractible, as otherwise fw(M) = 1 and we are done. This implies that every 2-cycle has two well-defined sides. Assume w.l.o.g. that c is innermost, that is: it contains no 2-cycle in its simple, plane face fc. Let v and w be the vertices of c. Assume that v has colour 1. Then the two neigh- bours of v that are not w are on different sides of c. Every face has only three edges and there are no vertices of degree 2, so there are two edges between each of these neighbours of v and w. This is not possible as c is innermost. It follows that the two vertices of c have colours 0 and 2 respectively. There is at least one vertex ef of colour 1 in the interior of fc. This vertex ef has degree 4. If there are no 0-vertices in the interior of fc then there must 172 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 u v f1 f2 x y Figure 8: This figure clarifies the proof of (ii) of Lemma 4.3. The blue vertices are all in the same component of M \ {x, y}, and the black vertices are not in that component. be two edges between ef and the 0-vertex of c. This is a contradiction with the assumption that c is innermost. It follows that fc has a 0-vertex in its interior. As every 2-cycle has two well-defined sides, there is an innermost 2-cycle in the other face gc of c. Using the same arguments as for fc on that cycle we get that gc also has a 0-vertex in its interior. Every path in BM between the 0-vertices in the two faces using only 2-edges must pass through the 0-vertex of c. It follows that this vertex is a cutvertex of M , so that M is not c2. (ii): Let M be a map and assume that M is not c3. If it is also not c2 we are done, so assume that M is c2. There are four possible reasons for not being c3: the existence of a cutset {x, y} of size two, the existence of a face of size two, the existence of a vertex with degree 2, or the existence of a non-contractible 4-cycle in BM . Again the last three, as well as double edges forming a non-facial 2-cycle in M , immediately give a nontrivial 4-cycle in BM , so assume that there are no double edges or loops in M , but there is a 2-cut {x, y}. Both x and y have neighbours in different components. Let u ̸= y be a neighbour of x, so that the previous vertex in the rotational order around x is not in the same component of M \ {x, y} as u. Let v be the last vertex, as seen from u, in the rotational order around x such that v and all vertices in the rotational order between u and v are in the same component as u, as shown in Figure 8. Note that u = v is possible. If the edges (x, u) and (v, x) would belong to the same face then there would be a colour-1 cycle of length 2 in BM , so they are in different faces f1 and f2 of M . Both f1 and f2 must also contain y as otherwise the next, resp. previous neighbour of x would belong to the same component as u and v. The cycle x, f1, y, f2 is a nontrivial cycle of length 4 in BM . Conversely, assume that M is c3 and that M is not c2 or BM has a nontrivial cycle of length 4. As M is c3 it is also c2, so BM has a nontrivial cycle c of length 4. Note that there are no double edges in BM as M is c2, and M is simple and 3-connected by Lemma 4.2. The cycle c is contractible because fw(M) ≥ 3. It therefore has two well-defined sides. Assume first that c has no vertices of colour 0 on one side. Then there must be a 2-vertex on that side, as there cannot only be vertices of colour 1. This 2-vertex can have degree at most 4 as it can be adjacent to at most two 0-vertices in c and there are no double edges in BM . This implies a facial 2-cycle in M , a contradiction. It follows that there is at least one 0-vertex on each side of c. Every colour-2 path between 0-vertices on different sides passes through c. This implies that the vertices and edges of M corresponding to vertices of c form a cut of M . Ignoring edges if one of their incident vertices is also in c, G. Brinkmann et al.: On local operations that preserve symmetries . . . 173 this cut consists of 2 vertices, a vertex and an edge or two edges. For each of the edges we can choose one of its incident vertices such that we find a cut-set consisting of 2 vertices, which is a contradiction with the 3-connectivity of M . Lemma 4.3 is very useful to determine whether a map is c2 or c3. It will often be used in the following lemmas and theorems. The main theorem of this last section is Theorem 4.9, which shows the equivalence of different definitions of ck-lopsp-operations and states that when applying ck-lopsp-operations with k ∈ {1, 2, 3} to certain maps, the result is ck. The most difficult part of its proof is captured in Theorem 4.5 for c2-maps and Theorem 4.7 for c3-maps. Lemma 4.4. Let O be a lopsp-operation with a cut-path P of minimal length. (i) If the vertices of an edge e in O are both in Pv0,vi for an i ∈ {1, 2}, then e or an edge with the same vertices as e is also in Pv0,vi . (ii) If the vertices of an edge in OP are in different copies of Pv0,v1 , then there is a nontrivial 4-cycle in two copies of OP sharing their copies of Pv0,v1 . Proof. (i): This follows immediately from the minimality of the length of P . (ii): If Pv0,v1 is v0 = t1, . . . , tk = v1 and we denote one copy with t1, . . . , tk and the other with t′1, . . . , t ′ k, then — again due to minimality and as O has no loops — such an edge connects w.l.o.g. ti with t′i+1 for some 1 ≤ i < (k − 1). Considering two copies of OP sharing the copies of Pv0,v1 , this gives a 4-cycle c = ti, t ′ i+1, t ′ i, ti+1 with v1 in the interior. If c was trivial, then v1 would be a 1-vertex adjacent to all 4 vertices on c — also ti and t′i — which contradicts the minimality of P . Theorem 4.5. Let M be a c2-map and let O be a lopsp-operation. Then O(M) is c2 if and only if for each cut path P in O we have that there is no 2-cycle in OP . Note that we must consider 2-cycles in OP and not in O. It is possible that there are 2- cycles in O that do not induce 2-cycles in OP for any cut-path. For example, the operation gyro, shown in Figure 4, has several 2-cycles but none in OP for any cut-path P . Proof. If there is a 2-cycle in OP for a cut-path P then each copy of OP inserted into DM,P contains a copy of this 2-cycle. Lemma 4.3 now implies that O(M) is not c2. Conversely, assume that O(M) is not c2. Then there is a 2-cycle c in BO(M). Let x and y be the vertices of c and let e1 and e2 be its edges. Let P be a cut-path in O of minimal length. Note that by Lemma 4.3 applied to M , every double chamber has two different 0-vertices and therefore the boundary of each double chamber is simple. It follows that if there exists a double chamber that contains both edges of c in its interior or on its boundary, then c induces a cycle in a copy of OP and we are done. Assume that e1 and e2 are in different double chambers D1 and D2 respectively. Both D1 and D2 contain x and y, so those vertices are on the boundary of both double chambers. Assume first that x and y are not both on copies of Pv0,v1 or both on copies of Pv0,v2 . Then D1 and D2 share their 1-vertex and their 2-vertex which implies a 2-cycle in BM , a contradiction. It follows that x and y are both on copies of Pv0,v1 or both on copies of Pv0,v2 . If they are in the same copy of Pv0,v1 or Pv0,v2 , then by Lemma 4.4(i) an edge e0 with the same vertices is also in that copy of Pv0,v1 or Pv0,v2 , so that OP contains a 2-cycle. 174 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 x y e1 D1 D2 e2 e′1 Figure 9: This figure clarifies a step in the proof of Theorem 4.5. It shows that a 2-cycle in BO(M) with its vertices on different copies of the same Pv0,vi cannot exist. The middle vertex can be the 1- or the 2-vertex of the double chambers. In either case, the left- and rightmost vertices are the 0-vertices. t′0=t0 t1t ′ 1 tit ′ i tjt ′ j tkt ′ k tst ′ s Figure 10: The double chamber patch in the proof of Lemma 4.6. There can be no edge {t′l, tm} with l < j and m > i. The last possibility is that x and y are in different copies of Pv0,v1 or in different copies of Pv0,v2 . As O does not contain loops we have π(x) ̸= π(y). Applying Jordan’s curve the- orem to c we get that the edge e′1 in D2 with π(e1) = π(e ′ 1) cannot exist — a contradiction (see Figure 9). Lemma 4.6. Let M be a c3-map and let O be a lopsp-operation with a cut-path P of min- imal length. Let f be a 2-vertex of DM , and consider the submap Sf of BO(M) consisting of all the edges and vertices in the double chambers with 2-vertex f . If there is a nontrivial 4-cycle c in Sf , then there is a 2-cycle in OP or c is contained in either only one of these double chambers, or in two adjacent double chambers. Proof. As M is c3, the submap of all vertices and edges of O(M) belonging to one of the double chambers containing f is plane. The map formed by the vertices and edges on the 2-sides of the double chambers in Sf is a simple cycle, that we consider to be the boundary of the outer face of the map Sf . There are at least three double chambers in Sf . If c contains only edges on edges of DM , then c is the boundary of one double chamber, so assume that c contains at least one edge in the interior of a double chamber. The boundary ∂D of every double chamber D in Sf is a cycle. As Sf is plane, it follows with the Jordan curve theorem that c must cross ∂D an even number of times. By cross we mean that there is a subpath of c whose first and last vertex are on different sides of c, and whose other vertices are all in ∂D. Let D be a double chamber in Sf that contains G. Brinkmann et al.: On local operations that preserve symmetries . . . 175 an edge of c in its interior. If c crosses ∂D 0 times, then c is contained in one copy of OP and we are done. If c crosses ∂D 4 times then there must be an edge outside of D that has both its vertices on ∂D. In this case Lemma 4.4 implies that there is a 2-cycle in OP and we are done. We can therefore assume that c crosses ∂D exactly twice. Note that every crossing is on a 1-side. Assume that the vertices of these crossings are on the same 1-side of D. If there would be only one edge of c in D this again leads to a 2-cycle in OP with Lemma 4.4. If there is no crossing in f it is clear that a chamber D′ adjacent to D also has two crossings with c on the same 1-side. If f is in c that follows from the fact that f is on every 1-side and there must be at least two edges of c in a double chamber that has 2 crossings with c on the same 1-side. It follows that c is completely contained in D and D′, i.e. in two adjacent double chambers. We can now assume that c has two crossings with ∂D that are on different 1-sides and not in f . As c crosses into the double chambers adjacent to D those must also have two crossings on different 1-sides. Repeating this argument we get that every double chamber in Sf has two crossings with c on different 1-sides. It follows that every double chamber in Sf contains a subpath of c connecting vertices different from f on their two 1-sides. As there are at least three double chambers in Sf and there must be at least one edge of c in each one, there are exactly three or four double chambers in Sf . In each case there are at least two adjacent double chambers that contain only one edge of c. Let e1 and e2 be the only edges of c in two adjacent double chambers. As the vertices of e1 and e2 are on different 1-sides of OP the edges e1 and e2 are not in P . Therefore the edges π(e1) and π(e2) induce unique edges, that we will also denote with π(e1) and π(e2), in OP . If the vertices of Pv2,v0 in O are — in this order — t0, t1, . . . , ts, then we will denote the vertices on the different 1-sides of OP with t0, t1, . . . , ts, resp. t′0, t ′ 1, . . . , t ′ s. We have π(e1) = {ti, t′j} and π(e2) = {tj , t′k} with w.l.o.g. 0 < i < j ≤ s. Note that i ̸= j as there are no loops in O. Due to the Jordan curve theorem applied to the cycle t0 = t′0, t ′ 1, . . . , t ′ j , ti, ti−1, . . . , t0 there is no edge {tm, t′l} in OP with m > i and l < j. As j > i and π(e2) = {tj , t′k} is in OP , it follows that k > j. This situation is shown in Figure 10. If there are four double chambers in Sf we can repeat this argument on every pair of adjacent double chambers. With x1, x2, x3, x4 the vertices of c in cyclic order and π(xa) = tia we get that ia < ia+1 for all 1 ≤ a ≤ 3 and i4 < i1, so that by transitivity i1 < i1, a contradiction. If there are only three double chambers in Sf , then there must be two edges e3 and e4 of c in the same double chamber. The edges π(e3) and π(e4) form a path from t′i to tk. Such a path would have to cross both the edges π(e1) and π(e2), which is only possible if the path has at least three edges — it must contain ti or t′j and tj or t ′ k— which is a contradiction. Theorem 4.7. Let M be a c3-map and let O be a lopsp-operation. Then O(M) is c3 if and only if it is c2 and for each cut-path P in O we have that there is no nontrivial 4-cycle in a patch of two adjacent copies of OP sharing one of their sides. Proof. The implication that O(M) is not c3 if there is a 2- or nontrivial 4-cycle for some cut-path P is obvious, as corresponding pairs of two adjacent copies of OP in BO(M) would contain such cycles. For the other implication we assume that O(M) is not c3 but it is c2, and that there is no nontrivial 4-cycle in a patch of two adjacent copies of OP for a cut-path P . We will come to a contradiction by constructing such a 4-cycle. Let P be a cut-path in O of minimal 176 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 length. We will refer to the copies of v2 or v0 in the double chamber patches as the corners of the double chamber patches. By Lemma 4.3 there is a nontrivial 4-cycle c in BO(M). Let X be a set of double chambers in DM of minimal size, so that the union of all double chamber patches for double chambers in X contains c. For simplicity we will also refer to the set of those double chamber patches as X . The fact that c has four edges implies that 1 ≤ |X| ≤ 4. If |X| = 1 then c can be thought of as a 4-cycle in OP , which we assumed does not exist, so |X| > 1. We make the following observations: (i) Every double chamber in X shares at least two of its corners with other elements of X: As |X| > 1 the cycle c ‘enters’ and ‘leaves’ any double chamber D ∈ X in two different vertices. Both of these vertices must be in the intersection of D with the other double chambers of X . (ii) If two double chambers share two corners they also share the side containing those two corners. This follows immediately from the fact that there are no double edges in M and BM . (iii) If the intersection of a double chamber D ∈ X with the other double chambers of X is exactly one side, then there are at least two edges of c in D: Assume that there is only one edge e of c in D, and let D′ be the double chamber of X sharing the side with D. Both vertices of e are on the side shared by D and D′, but e itself is not, as otherwise D could be removed from X and X would not be minimal. If the vertices of e are on the same edge of DM , then Lemma 4.4 implies that there is a 2-cycle in OP , a contradiction with Theorem 4.5. If the vertices of e are on different copies of Pv0,v1 then Lemma 4.4 implies a nontrivial 4-cycle in two adjacent copies of OP , a contradiction. It follows from (i) and (ii) that if |X| = 2 then c is a 4-cycle in two adjacent copies of OP , so |X| > 2. We will now prove that there is a cycle of length 4 in DM that contains at least one edge of each double chamber in X . We call such a cycle a saturating 4-cycle. Assume first that every double chamber in X shares the same 2-vertex. With (i), (ii), and (iii), it follows easily that X consists of all the three or four double chambers corre- sponding to one face of M . Lemma 4.6 now implies that O(M) is not c2 or that there is a 4-cycle in two adjacent copies of OP . Both are contradictions. Now assume that there is a 2-vertex f of DM such that there is only one edge e of c in the union of all the double chamber patches of X with 2-vertex f . The edge e has both its vertices on the same 2-side. As at least one of its vertices is not a corner, both double chambers sharing that 2-side are in X . This is a contradiction with (iii). It follows that there are two 2-vertices f and g in DM such that the unions Xf and Xg of double chambers patches in X with 2-vertex f and g respectively each contain exactly two edges of c. With (i), (ii), and (iii) it follows that there are 0-vertices v and w of double chambers in Xf and Xg such that v, f, w, g is a saturating 4-cycle. As M is c3, Lemma 4.3 implies that the saturating 4-cycle is the boundary of a double chamber, or two double chambers sharing the 1-vertex. If e is the 1-vertex in these one or two double chambers, then c is contained in the set Ne consisting of all six double chambers that share a side with a double chamber containing e. We say that the two double chambers with 1-vertex e are the central double chambers, and the four other double chambers in Ne are the extremal double chambers. The three possible configurations of Ne with respect to shared sides are shown in Figure 11. Using the fact G. Brinkmann et al.: On local operations that preserve symmetries . . . 177 e g f e g f e g f Figure 11: The three possible configurations of the six double chambers in the set Ne are shown here. Different vertices in the drawing represent different vertices of DM . that there are no nontrivial 4-cycles in BM it can easily be verified that no two vertices in Figure 11 represent the same vertex of DM . We already proved that |X| ≥ 3 and that there are two edges of c in Xf and two in Xg . Therefore we can assume w.l.o.g. that Xf consists of two double chambers. It follows from (iii) that these two double chambers are extremal double chambers. The two vertices of the edge e are in c. If the third vertex of c in Xf is f , then with Lemma 4.4 and the fact that there are no 2-cycles in OP , it follows that the edges of c in Xf are 1-edges of DM . In that case we can replace the two extremal double chambers in X by the central double chamber so X was not of minimal size, a contradiction. If the third vertex x of c in Xf is not f , then the extremal double chambers share a side and x is on this side. Let e1 and e2 be the edges of c in these double chambers. The edge e1 corresponds to an edge in OP from w.l.o.g. v0,R to an internal vertex of P(v0,L),v2 and e2 corresponds to an edge from v0,L to P(v0,R),v2 . This would imply crossing edges in the plane map OP , a contradiction. It follows that Xf and Xg each consist of only one double chamber, which by (i) must be the central double chamber. Then c is a nontrivial 4-cycle in a patch of two adjacent copies of OP , a contradiction. For a lopsp-operation O, let TO be the tiling of the Euclidean plane obtained by apply- ing O to the regular hexagonal tiling of the plane. We say TO is the associated tiling of O. With the definition for lopsp-operations from [2] this is the tiling from which O is defined. In Section 5 we will further explore the fundamental connection between lopsp-operations and tilings. We will use the connectivity of the associated tiling of a lopsp-operation to define when an operation is ck. With Theorem 4.5 and Theorem 4.7 we will then prove an equivalent characterisation in Theorem 4.9 which does not depend on the associated tiling. Definition 4.8. For k ∈ {1, 2, 3} a lopsp-operation O is ck if the associated tiling TO is k-connected and all faces have size at least k. An lsp-operation is ck if the equivalent lopsp-operation Olopsp is ck. For a ck-lopsp-operation O and a map M with minimum face size at least k and min- imum degree at least k that is not necessarily ck, the map O(M) also has minimum face size at least k and minimum degree at least k. For the vertices of BO(M) of colour 0 or 2 178 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 that are not 0-vertices or 2-vertices of DM , the fact that they have degree at least 2k follows from O being a ck-lopsp-operation, as these degrees also occur in the tiling TO. For the others it follows from the degrees of 0-vertices and 2-vertices in DM . In case M is also ck, we have a stronger result: Theorem 4.9. The following statements are equivalent for k ∈ {1, 2, 3} and a lopsp- operation O: (a) O is a ck-lopsp-operation (b) For all ck-maps M , O(M) is ck. (c) There exists a ck-map M such that O(M) is ck. Proof. (a) ⇒ (b) For k = 1 this is trivial. Assume that there is a c2-map M such that O(M) is not c2. By Theorem 4.5 there is a 2-cycle in OP for some cut-path P in O. This cycle induces a cycle in BTO , which implies a 1-cut or a face of size 1 in the tiling TO. It follows that O is not a c2-operation — a contradiction. Similarly, if there is a c3-map M such that O(M) is not c3, we find a 2-cut in TO using the cycle from Theorem 4.7. (b) ⇒ (a) For k = 1 this is trivial. The tiling TO is obtained by inserting copies of OP (for some P ) into the double cham- ber map of the hexagonal tiling. Let us call the map formed by the subdivided 2-edges of the double chamber map of the hexagonal tiling the hexagonal skeleton. Assume now that O is not a c2-lopsp operation. Then there is a face f of size 1 or a 1-cut {x} in TO. In case of a 1-cut, at least one of the components of TO \ {x}, say C0, is finite. Let C denote a finite submap of the hexagonal skeleton that contains f , resp. C0 together with the cut vertex x. Using Goldberg-Coxeter operations (see [2] or [4]) with sufficiently large parameters to construct large icosahedral fullerenes, we get a fullerene F , that is c3 and contains an isomorphic copy of C with the paths between the 0-vertices replaced by edges. Applying O to that fullerene, we get a submap S of O(F ) that has a face f ′ of size 1 or that is isomorphic to C0 and where all vertices corresponding to vertices of C0 have — except for the vertex x′ corresponding to x — only neighbours in S. So f ′ or the vertex x′, which is a cut-vertex of O(F ), are contradictions to the assumption. The case k = 3 is completely analogous, with also a 2-face and a 2-cut in the argument. (b) ⇒ (c) Trivial. (c) ⇒ (b) Note that the conditions in Theorems 4.5 and 4.7 are — except for M being ck — independent of M , as OP and the union of two copies of OP sharing a side are the same for all these M . This implies that if O(M) is ck for some ck-map M , then O(N) is ck for any ck-map N . One of the main results of this paper, Theorem 4.10, now follows from Theorem 4.9 and Lemma 2.5. Theorem 4.10. If M is a polyhedral map and O is a c3-lsp- or c3-lopsp-operation, then O(M) is also a polyhedral map. G. Brinkmann et al.: On local operations that preserve symmetries . . . 179 σ1 v2 v1 v0 C1 C2 C3 v2 σ0 σ0 σ0 σ2 σ1 m01 m12 C1 12 3 C2 12 3 C3 3 3 Figure 12: The operation truncation and the Delaney-Dress symbol encoding a tiling from which the operation can be obtained when the original definition is applied. 5 Connection to tilings In a series of papers [8, 9, 10], Andreas Dress (in later papers together with coauthors) developed a finite symbol encoding the topology as well as the symmetry of periodic tilings. He attributed the idea to Matthew Delaney and called these symbols Delaney symbols. In later papers by other authors, these symbols are called Delaney-Dress symbols. In [6] and [10] Delaney-Dress symbols of periodic tilings of the Euclidean plane and the hyperbolic plane are characterized. In this section we show that there is a very fundamental connection between l(op)sp- operations and Delaney-Dress symbols and therefore to tilings. Recall that we defined the associated tiling TO of a lopsp-operation O as the tiling that is the result of applying O to the hexagonal tiling of the plane, i.e. the tiling with Schläfli symbol {6, 3}. We will find the same tiling in a different way using Delaney-Dress symbols, and we will see that from a mathematical point of view the choice of the tiling {6, 3} is quite arbitrary. The hexagonal tiling was chosen because it was also used in the original definition of lsp-operations in [2], where in turn it was chosen as a tribute to a paper by Goldberg [15]. By proving this connection it follows that our abstract combinatorial definitions are equivalent – in the 3-connected case – to the definitions of lsp- and lopsp-operations in [2]. As a topological definition of tilings falls outside the scope of this article, we will directly start with the combinatorial characterization described in [6, 9, 10]. We will sketch the connection to tilings, but for a detailed description we refer the reader to [6] or [10]. Theorem 5.1 (A.W.M. Dress [9]). Let D be a set together with an action (from the right) of the Coxeter group Σ = ⟨σ0, σ1, σ2 | σ2i = 1⟩ on D , and for (i, j) ∈ {(0, 1), (0, 2), (1, 2)} let mij : D → N be maps with m02(C) = 2 for all C ∈ D . The tuple (D ,m01,m02,m12) is the Delaney-Dress symbol of a tiling of the Euclidean plane if and only if the following properties hold: (1) D has finitely many elements (2) Σ acts transitively on D 180 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 (3) For i, j ∈ {0, 1, 2}, i < j, mij is constant on ⟨σi, σj⟩-orbits and C(σiσj)mij(C) = C for all C ∈ D (4) We have C (D ,m01,m02,m12) = ∑ C∈D ( 1 m01(C) + 1 m12(C) − 1 m02(C) ) = 0 Such Delaney-Dress symbols encode the combinatorial structure of periodic tilings of the Euclidean plane, together with a symmetry group acting on the tiling. If C (D ,m01, m02,m12) ̸= 0, the tuple can also be a Delaney-Dress symbol, but then it encodes a periodic tiling of the hyperbolic plane (C < 0) or — in case additional divisibility rules are fulfilled — the sphere (C > 0) [6]. The elements of D are the orbits of chambers of the tiling under the symmetry group. An element C ∈ D with Cσi = C represents an orbit of chambers with mirror symmetries of the tiling stabilizing the edges of colour i. If there are no C ∈ D with Cσi = C, the symmetry group contains no pure reflections, but maybe sliding reflections. If there are no odd cycles, that is Cσi1 . . . σik ̸= C for odd k, all symmetries are orientation preserving. The maps m01 and m12 give information about the symmetry group of the tiling. Let {i, j, k} = {0, 1, 2}, i < j and for C ∈ D let rij(C) = min{r | C(σiσj)r = C}. Note that rij is constant on ⟨σi, σj⟩-orbits. If a ⟨σi, σj⟩-orbit C⟨σi,σj⟩ contains no C ′ with C ′σi = C ′ or C ′σj = C ′, then the vertices of colour k of the corresponding chambers in the tiling are centers of an fr-fold rotation with fr = mij(C)/rij(C). If an orbit C⟨σi,σj⟩ contains a C ′ with C ′σi = C ′ or C ′σj = C ′, then with fm = 2mij(C)/rij(C) for fm > 1 the vertices of colour k of the chambers in orbit C are intersections of mirror axes with an angle of 360/fm degrees. We will now associate a tuple (DO,m01,m02,m12) with an lsp- or lopsp-operation O and prove that it is a Delaney-Dress symbol. In fact, it will be a Delaney-Dress symbol of the tiling O(T ) where T is the tiling with Schläfli symbol {6, 3}, i.e. the hexagonal tiling of the plane where every vertex has degree 3 and every face has 6 edges. Due to the relation between Delaney-Dress symbols and tilings as described in [6] and [10], this also shows the equivalence of the combinatorial definitions of lsp- and lopsp-operations defined here and the geometric ones given in [2]. There a l(op)sp-operation is described as a ‘triangle’ cut out of a tiling in such a way that certain conditions on the symmetry are satisfied. One could replace the values 3 and 6 we will use for defining the mappings mij by, for example, 4 and 4, and Theorem 5.3 would still be true. It would however be the Delaney- Dress symbol of the tiling that can be obtained by applying O to the square tiling of the plane, which is 4-regular and every face has 4 edges. By using other numbers, other tilings — even spherical or hyperbolic ones — could be used as source tilings. All of those tilings can be used to define l(op)sp-operations in the geometric way that was described in [2] for the hexagonal tiling. Let O be an lsp-operation and let DO be the set of chambers of O. We define the action of Σ on DO by letting Cσi = C ′ if C and C ′ share their i-edge, and Cσi = C if the i-edge of C is in the outer face of O. For (i, j) ∈ {(0, 1), (0, 2), (1, 2)}, let vij(C) be the vertex of chamber C that is not of colour i or j. We get: G. Brinkmann et al.: On local operations that preserve symmetries . . . 181 rij(C) = min {r | C(σiσj)r = C} =  ∣∣C⟨σi,σj⟩∣∣ 2 = deg(vij(C)) 2 if vij(C) is not in the outer face∣∣C⟨σi,σj⟩∣∣ = deg(vij(C))− 1 if vij(C) is in the outer face To find the Delaney-Dress symbol of the tiling obtained by applying O to {6, 3} we define mij : DO → N as follows: mij(C) =  rij(C) · 2 if vij(C) = v1 rij(C) · 3 if vij(C) = v0 rij(C) · 6 if vij(C) = v2 rij(C) if vij(C) /∈ {v0, v1, v2} Note that the requirements for the vertex degrees in an lsp-operation imply that for all C ∈ DO, the value m02(C) is 2. We define D(O) = (DO,m01,m02,m12) and call it the Delaney-Dress symbol cor- responding to the lsp-operation O. This correspondence is illustrated for the operation truncation in Figure 12. Theorem 5.2 states that it is in fact a Delaney-Dress symbol of a tiling of the Euclidean plane. By our previous remarks there is a 2-fold rotation around each copy of v1 in that tiling, a 3-fold rotation around each copy of v0, and a 6-fold rotation around each copy of v2. There are also intersections of mirror axes with 90◦, 60◦, and 30◦ angles at v1, v0, and v2 respectively. This is the symmetry we expect when applying an lsp-operation to tiling {6, 3}. This is also the symmetry that is required to define an lsp-operation from a tiling with the geometric definition. Theorem 5.2. If O is an lsp-operation, then D(O) = (DO,m01,m02,m12) is the Delaney- Dress symbol of a tiling of the Euclidean plane. Proof. We have to prove the properties in Theorem 5.1. The first two properties are obvi- ous, so we will focus on the other two. (3): Let (i, j) ∈ {(0, 1), (0, 2), (1, 2)}. A ⟨σi, σj⟩-orbit consists of all the chambers sharing the same vertex vij(C), so that by definition mij is constant on ⟨σi, σj⟩- orbits. It is clear that C(σiσj)mij(C) = C(σiσj)rij(C)·k = C (4): Let {i, j, k} = {0, 1, 2}. For a vertex v of colour k and i < j we define α(v) =∑ C∈DO v∈C ( 1 mij(C) ) . Counting the number of chambers with a certain vertex and using the definition of mij , we get that 182 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 v2 v1 v0,L v0,R C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 m01 m12 C1 5 6 C2 5 6 C3 5 3 C4 5 3 C5 5 3 C6 5 3 C7 5 3 C8 5 3 C9 5 3 C10 5 3 Figure 13: On the left, the double chamber patch of the lopsp-operation gyro is shown and on the right the corresponding Delaney-Dress symbol. α(v) =  2 if v is an inner vertex 1 if v is an outer vertex different from vi for i = 0, 1, 2 1/2 if v = v1 1/3 if v = v0 1/6 if v = v2 . Let n be the number of vertices (and equivalently edges) in the outer face. As every vertex of O has exactly one colour we get that: C (DO,m01,m02,m12) = ∑ C∈DO ( 1 m01(C) + 1 m12(C) − 1 m02(C) ) = ∑ C∈DO ( 1 m01(C) + 1 m12(C) + 1 m02(C) ) − ∑ C∈DO (1) = ∑ v∈VO α(v)− (|FO| − 1) = (|VO| − n) · 2 + (n− 3) · 1 + 1 2 + 1 3 + 1 6 − (|FO| − 1) = 2|VO| − |FO| − n− 1 By counting the number of directed edges associated with edges in the triangulated disk O in two ways, we get that 2|EO| = 3(|FO| − 1) + n or equivalently |FO| = 2|EO| − 2|FO|+ 3− n. We also know that O is plane, so |VO| − |EO|+ |FO| = 2. It follows that: C (DO,m01,m02,m12) = 2|VO| − 2|EO|+ 2|FO| − 3 + n− n− 1 = 0 G. Brinkmann et al.: On local operations that preserve symmetries . . . 183 We will now prove the corresponding result for lopsp-operations. Let O be a lopsp- operation and let DO be the set of chambers of O. We define the action of Σ on DO by letting Cσi = C ′ if C and C ′ share their i-edge. For lopsp-operations there is no outer face, so rij(C) is always deg(vij) 2 . We define m01,m02,m12 : DO → N exactly as before: mij(C) =  rij(C) · 2 if vij(C) = v1 rij(C) · 3 if vij(C) = v0 rij(C) · 6 if vij(C) = v2 rij(C) if vij(C) /∈ {v0, v1, v2} Again m02(C) = 2 for all C ∈ DO. We define D(O) = (DO,m01,m02,m12) and in Theorem 5.3 we prove that it is a Delaney-Dress symbol. The operation gyro and its corresponding Delaney-Dress symbol are shown as an example in Figure 13. Once again, the tiling described by the Delaney-Dress symbol is the result of applying the operation to the hexagonal tiling of the plane. In Section 4 we named this tiling the associated tiling TO of O. There are 2-, 3-, and 6-fold rotations at the copies of v1, v0, and v2 respectively. In lopsp-operations there is no chamber C such that Cσi = C so there are no pure reflections encoded in the Delaney-Dress symbol. This is the symmetry required in the geometric definition of lopsp-operations. Theorem 5.3. If O is a lopsp-operation, then D(O) = (DO,m01,m02,m12) is the Delaney- Dress symbol of a tiling of the Euclidean plane. Proof. We prove the properties in Theorem 5.1. Again, the first two are obvious. (3): As in the proof of Theorem 5.2. (4): Let {i, j, k} = {0, 1, 2}. For a vertex v ∈ VO of colour k and i < j we again define α(v) = ∑ C∈DO v∈C ( 1 mij(C) ) . Counting the number of chambers with a given vertex v and using the definition of mij , we get that α(v) =  2 if v /∈ {v0, v1, v2} 1 if v = v1 2/3 if v = v0 1/3 if v = v2 . We can now compute C (DO,m01,m02,m12): 184 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 C (DO,m01,m02,m12) = ∑ C∈DO ( 1 m01(C) + 1 m12(C) − 1 m02 ) = ∑ C∈DO ( 1 m01(C) + 1 m12(C) + 1 m02 − 1 ) = ∑ v∈VO α(v)− |DO| = (|VO| − 3) · 2 + 2 3 + 1 + 1 3 − |FO| = 2|VO| − |FO| − 4 As O is a triangulation, we get that 2|EO| = 3|FO| and as O is plane, we have |VO| − |EO|+ |FO| = 2. It follows that: C (DO,m01,m02,m12) = 2|VO| − 2|EO|+ 2|FO| − 4 = 0 In Lemma 2.5 we proved that for every lsp-operation there is an equivalent lopsp- operation Olopsp. Lemma 5.4 proves formally that the Delaney-Dress symbols of the lsp- operation and its corresponding lopsp-operation in fact encode isomorphic tilings. Lemma 5.4. The Delaney-Dress symbols D(O) and D(Olopsp) are Delaney-Dress sym- bols of combinatorially isomorphic tilings. Proof. Mapping each chamber Clopsp of D(Olopsp) onto the corresponding chamber C of D(O), we have (in the notation of [10]) a morphism between the symbols and in the nota- tion of [6] a Delaney map f , that is: For all k ∈ {0, 1, 2}, (i, j) ∈ {(0, 1), (0, 2), (1, 2)}, and chambers C of D(Olopsp) we have f(Cσk) = (f(C))σk and mij(C) = mij(f(C)). The existence of such a morphism guarantees (see [6, 10]) that D(O) and D(Olopsp) code combinatorially isomorphic tilings and that the tiling coded by D(Olopsp) can be obtained from the tiling coded by D(O) by symmetry breaking — That is: modifying the tiling, so that the combinatorial structure is preserved, but some metric symmetries of the tiling are destroyed. 6 Future work In the last section of [2] many open problems are described. They are sometimes just formulated for lsp-operations, but are often as relevant and interesting for lopsp-operations, so we refer the reader to [2]. A very interesting question is whether ambo is ‘essentially’ the only lsp-operation that can increase the symmetry of polyhedra, i.e. plane 3-connected maps. More specifically: Assume that for an lsp-operation O and a polyhedron M , the polyhedron O(M) has more symmetries than M . Is M self-dual and can O be written as the product of ambo and other lsp-operations? For lopsp-operations this is certainly not true. For example, applying gyro to the tetrahedron gives the dodecahedron, which has a much larger symmetry group. Classifying lopsp-operations that can introduce new symmetries would be an interesting problem, but maybe even more difficult than solving the problem for lsp-operations. G. Brinkmann et al.: On local operations that preserve symmetries . . . 185 We know that there is at least one lsp-operation (dual) that does not always preserve 3-connectivity for maps, if the face-width is at most two [1], so an obvious question is which other operations do not always preserve 3-connectivity. This was answered for lsp- operations in [24], where the class of such operations, called edge-breaking operations, was characterized. Recently, these results have been extended to lopsp-operations. An article with the new results has been submitted [25]. Another problem mentioned in [2] — the generation of lsp-operations for a given in- flation factor — has been solved [13]. Such an algorithm not only allows the generation of lsp-operations, but also the generation of polyhedra and other maps with some specific symmetry groups of the embedding. For generating lopsp-operations a program has been written very recently, but it has not been published yet. ORCID iDs Gunnar Brinkmann https://orcid.org/0000-0003-4168-0877 Heidi Van den Camp https://orcid.org/0000-0001-7634-3681 References [1] D. Bokal, G. Brinkmann and C. T. Zamfirescu, The connectivity of the dual, J. Graph The- ory 101 (2022), 182–209, doi:10.1002/jgt.22819, https://doi.org/10.1002/jgt. 22819. [2] G. Brinkmann, P. Goetschalckx and S. Schein, Comparing the constructions of Goldberg, Fuller, Caspar, Klug and Coxeter, and a general approach to local symmetry-preserving op- erations, Proc. R. Soc. Lond., A, Math. Phys. Eng. Sci. 473 (2017), 14 pp., doi:10.1098/rspa. 2017.0267, id/No 20170267, https://doi.org/10.1098/rspa.2017.0267. [3] R. Bruce King and M. V. Diudea, The chirality of icosahedral fullerenes: a comparison of the tripling (leapfrog), quadrupling (chamfering), and septupling (capra) transformations, J. Math. Chem. 39 (2006), 597–604, doi:10.1007/s10910-005-9048-7, https://doi.org/ 10.1007/s10910-005-9048-7. [4] D. L. Caspar and A. Klug, Physical principles in the construction of regular viruses, in: Cold Spring Harbor Symposia on Quantitative Biology, Cold Spring Harbor Laboratory Press, volume 27, 1962 pp. 1–24, doi:10.1101/sqb.1962.027.001.005, https://doi.org/10. 1101/sqb.1962.027.001.005. [5] H. Coxeter, Virus macromolecules and geodesic domes, A spectrum of mathematics (1971), 98–107. [6] O. Delgado-Friedrichs, Data structures and algorithms for tilings I, Theor. Comput. Sci. 303 (2003), 431–445, doi:10.1016/S0304-3975(02)00500-5, https://doi.org/10.1016/ S0304-3975(02)00500-5. [7] A. Dress and G. Brinkmann, Phantasmagorical fulleroids, MATCH Commun. Math. Com- put. Chem. 33 (1996), 87–100, https://match.pmf.kg.ac.rs/electronic_ versions/Match33/match33_87-100.pdf. [8] A. W. Dress, Regular polytopes and equivariant tessellations from a combinatorial point of view, in: Algebraic Topology Göttingen 1984, Springer, pp. 56–72, 1985. [9] A. W. M. Dress, Presentations of discrete groups, acting on simply connected manifolds, in terms of parametrized systems of Coxeter matrices - a systematic approach, Adv. Math. 63 (1987), 196–212, doi:10.1016/0001-8708(87)90053-3, https://doi.org/10.1016/ 0001-8708(87)90053-3. 186 Ars Math. Contemp. 24 (2024) #P2.01 / 155–186 [10] A. W. M. Dress and D. Huson, On tilings of the plane, Geom. Dedicata 24 (1987), 295–310, doi:10.1007/bf00181602, https://doi.org/10.1007/bf00181602. [11] F. V. Fomin and D. M. Thilikos, On self duality of pathwidth in polyhedral graph embeddings, J. Graph Theory 55 (2007), 42–54, doi:10.1002/jgt.20219, https://doi.org/10.1002/ jgt.20219. [12] M. D. R. Francos, Chamfering operation on k-orbit maps, Ars Math. Contemp. 7 (2014), 519– 536, doi:10.26493/1855-3974.541.133, https://doi.org/10.26493/1855-3974. 541.133. [13] P. Goetschalckx, K. Coolsaet and N. Van Cleemput, Generation of local symmetry-preserving operations on polyhedra, Ars Math. Contemp. 18 (2020), 223–239, doi:10.26493/1855-3974. 1931.9cf, https://doi.org/10.26493/1855-3974.1931.9cf. [14] P. Goetschalckx, K. Coolsaet and N. Van Cleemput, Local orientation-preserving symmetry preserving operations on polyhedra, Discrete Math. 344 (2021), 10 pp., doi:10.1016/j.disc. 2020.112156, id/No 112156, https://doi.org/10.1016/j.disc.2020.112156. [15] M. Goldberg, A class of multi-symmetric polyhedra, Tôhoku Math. J. 43 (1937), 104–108. [16] J. L. Gross and T. W. Tucker, Topological Graph Theory, Courier Corporation, 2001. [17] B. Grünbaum, Graphs of polyhedra; polyhedra as graphs, Discrete Math. 307 (2007), 445– 463, doi:10.1016/j.disc.2005.09.037, https://doi.org/10.1016/j.disc.2005. 09.037. [18] N. W. Johnson, Convex polyhedra with regular faces, Can. J. Math. 18 (1966), 169–200, doi: 10.4153/cjm-1966-021-8, https://doi.org/10.4153/cjm-1966-021-8. [19] J. Kovič, T. Pisanski, A. T. Balaban and P. W. Fowler, On symmetries of benzenoid systems, MATCH Commun. Math. Comput. Chem. 72 (2014), 3–26. [20] B. Mohar, Face-width of embedded graphs, Math. Slovaca 47 (1997), 35–63. [21] B. Mohar and C. Thomassen, Graphs on Surfaces, volume 16, Johns Hopkins University Press Baltimore, 2001. [22] A. Orbanić, D. Pellicer and A. Ivić-Weiss, Map operations and k-orbit maps, J. Comb. The- ory, Ser. A 117 (2010), 411–429, doi:10.1016/j.jcta.2009.09.001, https://doi.org/10. 1016/j.jcta.2009.09.001. [23] T. Pisanski and M. Randic, Bridges between geometry and graph theory, MAA NOTES (2000), 174–194. [24] H. Van den Camp, The effect of local symmetry-preserving operations on the connectivity of embedded graphs, Master’s thesis, Ghent University, Belgium, 2020. [25] H. Van den Camp, The effect of symmetry-preserving operations on 3-connectivity, 2023, arXiv:2301.06913 [math.CO]. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.02 / 187–206 https://doi.org/10.26493/1855-3974.2896.7e6 (Also available at http://amc-journal.eu) Algebraic degrees of 2-Cayley digraphs over abelian groups* Yongjiang Wu, Jing Yang, Lihua Feng † School of Mathematics and Statistics, HNP-LAMA, Central South University, Changsha, Hunan, 410083, P.R. China Received 3 June 2022, accepted 22 March 2023, published online 8 September 2023 Abstract A digraph Γ is called a 2-Cayley digraph over a group G if there exists a 2-orbit semiregular subgroup of Aut(Γ) isomorphic to G. In this paper, we completely deter- mine the algebraic degrees of 2-Cayley digraphs over abelian groups. This generalizes the main results of Lu and Mönius in 2023. As applications, we consider the algebraic degrees of Cayley digraphs over finite groups admitting an abelian subgroup of index 2. Special attention is paid to the algebraic degrees of Cayley (di)graphs over generalized dihedral groups, generalized dicyclic groups and semi-dihedral groups. Keywords: Algebraic degree, 2-Cayley digraph, Abelian group. Math. Subj. Class. (2020): 05C25, 05C50 1 Introduction A digraph Γ consists of a finite set V (Γ) of vertices and a set E(Γ) of directed edges, where E(Γ) ⊆ V (Γ)× V (Γ). If (u, v) ∈ E(Γ) implies (v, u) ∈ E(Γ), then Γ is said to be undirected. For a digraph Γ on n vertices, its adjacency matrix A = (auv)n×n is defined as auv = { 1, if (u, v) ∈ E(Γ), 0, otherwise. The characteristic polynomial of Γ is the characteristic polynomial of A. The eigenvalues of A are called the eigenvalues of Γ. The collection of eigenvalues of Γ together with their *This research was supported by NSFC (Nos. 12271527, 12071484), Hunan Provincial Natural Science Foun- dation (2020JJ4675, 2018JJ2479). The authors would like to express their sincere thanks to the referee for the valuable suggestions which greatly improved the presentation of the original manuscript. †Corresponding author. E-mail addresses: su15273815046@163.com (Yongjiang Wu), yj1147943429@163.com (Jing Yang), fenglh@163.com (Lihua Feng) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 188 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 multiplicities is called the spectrum of Γ, denoted by Spec(Γ). Note that A is not always symmetric, so the eigenvalues of Γ need not be real numbers. Let G be a finite group and S ⊆ G \ {e} , where e is the identity. The Cayley digraph Γ = Cay(G,S) of G with respect to S is defined by V (Γ) = G and E(Γ) = {(g, sg) | g ∈ G, s ∈ S}. If S = S−1, then Γ = Cay(G,S) is called a Cayley graph. For a digraph Γ, the set of all permutations of V (Γ) that preserve the adjacency relation of Γ forms a group, called the automorphism group of Γ, and is denoted by Aut(Γ). By a theorem of Sabidussi [15], a digraph Γ is a Cayley digraph over G if and only if there exists a regular subgroup of Aut(Γ) isomorphic to G. As a generalization of Sabidussi’s Theorem [1], a digraph Γ is called a 2-Cayley digraph over G if there exists a 2-orbit semiregular subgroup of Aut(Γ) isomorphic to G. A 2-Cayley graph is also termed as a semi-Cayley graph in [5, 6]. A special 2-Cayley graph is called a bi-Cayley graph in [19]. For a digraph Γ, its splitting field SF(Γ) is the smallest field extension of Q which contains all eigenvalues of the adjacency matrix of Γ. The extension degree [SF(Γ) : Q] is called the algebraic degree of Γ, denoted by deg(Γ). A digraph Γ is called integral if all the eigenvalues of the adjacency matrix of Γ are integers. A digraph Γ is called algebraically integral over a number field K if all the eigenvalues of the adjacency matrix of Γ are algebraic integers of K. There is a close connection between the splitting field and the algebraic integrality of a digraph. For example, for any number field K, SF(Γ) ⊆ K if and only if Γ is algebraically integral over K. Integral graphs and algebraically integral graphs have been extensively studied in the literature [2, 3, 4, 8, 9, 11]. In recent years, the splitting field and algebraic degree have attracted much attention. In 2020, Mönius [13] studied the algebraic degrees of circulant graphs Cay (Zp, S) for a prime number p. In 2022, Mönius [14] generalized those results in [13] by determining the splitting fields and the algebraic degrees of circulant graphs Cay (Zn, S) for arbitrary n. Based on Mönius’s work, in 2022, Huang et al. [18] determined the splitting fields and algebraic degrees of mixed Cayley graphs over abelian groups. Lu et al. [12] determined the splitting fields of Cayley graphs over abelian groups and dihedral groups. They also gave bounds for the algebraic degrees of Cayley graphs over dihedral groups. Also in 2022, Sripaisan et al. [16] studied the algebraic degrees of Cayley hypergraphs. For more details, one may refer to the comprehensive survey [10] in this subject. In this paper, inspired by the above mentioned results, we completely determine the splitting fields and algebraic degrees of 2-Cayley digraphs over abelian groups in Section 3, which generalizes the main results of [12]. From computational viewpoints, we also derive sharp upper and lower bounds for their algebraic degrees. As applications, in Section 4, we consider the algebraic degrees of Cayley digraphs over finite groups admitting an abelian subgroup of index 2. Furthermore, we consider the algebraic degrees of Cayley graphs over generalized dihedral groups and generalized dicyclic groups, and get improved upper bounds. Finally, we determine the algebraic degrees of Cayley digraphs over semi-dihedral groups. 2 Preliminaries Let G be a finite group. A representation of G is a homomorphism ρ : G → GL(V ) for some n-dimensional vector space over the complex field C, where GL(V ) denotes the group of automorphisms of V . The dimension of V is called the degree of ρ . Two representations ρ1 and ρ2 of G on V1 and V2 respectively are equivalent if there is an Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 189 isomorphism T : V1 → V2 such that Tρ1(g) = ρ2(g)T for all g ∈ G. Let ρ : G → GL(V ) be a representation. The character χρ : G → C of ρ is defined by setting χρ(g) = Tr(ρ(g)) for g ∈ G, where Tr(ρ(g)) is the trace of the representation matrix of ρ(g) with respect to a specified basis of V . By the degree of χρ we mean the degree of ρ, which is simply χρ(1). If W is a ρ(g)-invariant subspace of V for each g ∈ G, then we call W a ρ(G)-invariant subspace of V . If the only ρ(G)-invariant subspace of V are {0} and V , we call ρ an irreducible representation of G, and the corresponding charac- ter χρ an irreducible character of G. We denote by IRR(G) and Irr(G) the complete set of non-equivalent irreducible representations of G and the complete set of non-equivalent irreducible characters of G, respectively. For any subset X ⊆ G, we denote by δX = (δg)g∈G the characteristic vector of X over G, where δg = 1 if g ∈ X and δg = 0 if g /∈ X . For any multi-subset X ⊆ G, we denote by δ′X = ( δ′g ) g∈G the characteristic vector of X over G, where δ ′ g = k if g appears k times in X and δ′g = 0 if g /∈ X . Throughout this paper, we use X = [x | x ∈ X] to denote the multi-set X , and φ(n) to denote the Euler totient function of a natural number n (it is the number of the positive integers which are smaller than n and coprime to n). Firstly, we state an equivalent definition of 2-Cayley digraphs. Lemma 2.1 ([1]). A digraph Γ is a 2-Cayley digraph over G if and only if there exist subsets Tij of G, where 1 ⩽ i, j ⩽ 2, such that Γ is isomorphic to a digraph Υ with V (Υ) = G× {1, 2}, E(Υ) = ⋃ 1⩽i,j⩽2 {((g, i), (tg, j)) | g ∈ G and t ∈ Tij} . By Lemma 2.1, a 2-Cayley digraph is characterized by a group G and four subsets Tij of G. Thus we denote a 2-Cayley digraph with respect to four subsets Tij by Γ = Cay (G;Tij | 1 ⩽ i, j ⩽ 2). Note that V (Γ) = G × {1, 2}, (g, i) ∼ (h, j) if and only if hg−1 ∈ Tij , and Γ is undirected if and only if for all 1 ⩽ i, j ⩽ 2, Tij = T−1ji . Note also that Γ is a digraph without loops if and only if Tii ⊆ G\{e}, for all 1 ⩽ i ⩽ 2. Let ωn = exp ( 2πi n ) be the primitive n-th root of unity. We consider an abelian group G of order n. It is well known that G ∼= Zn1 ⊕ · · · ⊕ Znr , where n = ∏r i=1 ni, and ni is a prime power for 1 ≤ i ≤ r. Without loss of generality, we assume that G = Zn1 ⊕ · · · ⊕ Znr and 0 = (0, . . . , 0) ∈ G is the identity of G. Lemma 2.2 ([17]). Let G = Zn1⊕· · ·⊕Znr be an abelian group of order n. Then Irr(G) = {χl | l ∈ G}, where χl(g) = ∏r i=1 ω ligi ni for all l = (l1, . . . , lr) , g = (g1, . . . , gr) ∈ G, and ωni = exp( 2πi ni ). For simplicity, for any (multi-)subset S of G, we denote χl(S) = ∑ s∈S χl(s). Arezoomand [1] obtained the following result. Lemma 2.3 ([1]). Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G = Zn1 ⊕ · · · ⊕ Znr of order n. Then Γ has eigenvalues χl(T11) + χl(T22)± √ (χl(T11)− χl(T22))2 + 4χl(T21)χl(T12) 2 , l ∈ G. 190 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 Let K be a field. In what follows, we will refer to the subgroup K×2 = { x2 : x ∈ K } ⊂ K×, where K× = K \ {0}. More precisely, we shall encounter quite often the quotient K×/K×2. The image of x ∈K×in K×/K×2 will be denoted by [x]K . Lemma 2.4 ([7, Corollary 1.23]). Suppose K is a field containing a primitive 2-th root of unity, and let F = K [√ a1, . . . , √ ak ] , where ai ∈ K. Then Gal(F/K) is isomorphic to the subgroup of K×/K×2 generated by [a1]K , . . . , [ak]K . 3 2-Cayley digraphs over abelian groups In this section, we always assume that G = Zn1 ⊕· · ·⊕Znr is an abelian group of order n. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over G. For any two subsets X and Y of G, we define the multi-set X + Y = [x+ y | x ∈ X, y ∈ Y ]. For any (multi-)set X and k ∈ N, k ∗X denotes the multi-set in which each element of X appears k times. For example, if X = [1, 1, 2, 2, 2, 3, 4] and k = 2, then k ∗X = 4 ∗ {1} ∪ 6 ∗ {2} ∪ 2 ∗ {3, 4}, with duplicate elements allowed in the union. For two multi-sets U = k1 ∗ {g1} ∪ k2 ∗ {g2} ∪ . . . ∪ ks ∗ {gs} and V = q1 ∗ {g1} ∪ q2 ∗ {g2} ∪ . . . ∪ qt ∗ {gt} ∪ ks+1 ∗ {gs+1} ∪ . . . ∪ ks+m ∗ {gs+m}, where ki ≥ qi, s > t and gi, 1 ≤ i ≤ s + m are pairwise distinct, we define U \ V = (k1 − q1) ∗ {g1} ∪ (k2 − q2) ∗ {g2} ∪ . . . ∪ (kt − qt) ∗ {gt} ∪ kt+1 ∗ {gt+1} ∪ . . . ∪ ks ∗ {gs}, V \ U = ks+1 ∗ {gs+1} ∪ . . . ∪ ks+m ∗ {gs+m}. Using the symbols in Lemma 2.3, we let I1 = [t | t ∈ T11 or t ∈ T22] , I2 = [t | t ∈ (T11 + T11) or t ∈ (T22 + T22) or t ∈ 4 ∗ (T12 + T21)], I3 = [t | t ∈ 2 ∗ (T11 + T22)], where I1, I2 and I3 are multi-sets. For example, for the group G = Z4, if T11 = {1, 2}, T12 = {1}, T21 = {2} and T22 = {3}, then I1 = {1, 2, 3}, I2 = 6 ∗ {3} ∪ 2 ∗ {2} ∪ {0} and I3 = 2 ∗ {0, 1}. By Lemma 2.3, we have the following result. Lemma 3.1. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. Then Γ has eigenvalues χl(I1)± √ χl(I2 \ I3)− χl(I3 \ I2) 2 , l ∈ G, where I1, I2, I3 are described as above. Proof. Firstly, we have χl(T11) + χl(T22) = χl(I1). In addition, (χl(T11)− χl(T22))2 + 4χl(T21)χl(T12) =χl(T11 + T11) + χl(T22 + T22) + χl (4 ∗ (T12 + T21))− χl (2 ∗ (T11 + T22)) =χl(I2)− χl(I3) =χl(I2 \ I3)− χl(I3 \ I2), Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 191 so the result follows from Lemma 2.3. Using the symbols in Lemma 3.1, for l ∈ G, let βl = χl(I1) and γl = χl(I2 \ I3)− χl(I3 \ I2). (3.1) As βl, γl ∈ Q (ωn) , where n = |G|, without loss of generality, we assume that K is a field such that Q ⊆ K ⊆ Q (ωn). Therefore, Gal (Q (ωn) /K)) ≤ Gal (Q (ωn) /Q) ∼= Z∗n = {k ∈ Zn | gcd(k, n) = 1}. Let η : Gal (Q (ωn) /Q) → Z∗n be the isomorphism such that σ (ωn) = ω η(σ) n , where σ ∈ Gal (Q (ωn) /Q). Let H = η (Gal (Q (ωn) /K)) . Then H is a subgroup of Z∗n. We consider the action of Z∗n on G = Zn1 ⊕ · · · ⊕ Znr by setting kg = k (g1, . . . , gr) = (kg1, . . . , kgr) for any k ∈ Z∗n and g ∈ G. Then σ ( ωlini ) = σ ( ωnli/nin ) = ωη(σ)·nli/nin = ω η(σ)li ni , where li ∈ Zni(1 ≤ i ≤ r). Note that for any σ ∈ Gal (Q (ωn) /Q), we have σ (βl) = σ (χl(I1)) = σ (∑ t∈I1 r∏ i=1 ωlitini ) = ∑ t∈I1 r∏ i=1 σ ( ωlitini ) = ∑ t∈I1 r∏ i=1 ωη(σ)litini = χl(η(σ)I1), where η(σ)I1 = {(η(σ)t1, . . . , η(σ)tr) | (t1, . . . , tr) ∈ I1}. Similarly, we have σ (γl) = σ (χl(I2 \ I3)− χl(I3 \ I2)) = χl (η(σ)(I2 \ I3))− χl (η(σ)(I3 \ I2)) , where I2 \ I3 and I3 \ I2 are multi-sets as stated in Lemma 3.1. We first prove the following results. Proposition 3.2. For the symbols in (3.1), we have βl ∈ K for all l ∈ G if and only if hI1 = I1 for all h ∈ H , where H = η (Gal (Q (ωn) /K)). Proof. Assume that hI1 = I1 for all h ∈ H . Then for any σ ∈ Gal (Q (ωn) /K), we have η(σ) ∈ H . Thus for any l ∈ G, we have σ (βl) = χl(η(σ)I1) = χl(I1) = βl. It follows that βl ∈ K for all l ∈ G. Conversely, assume that βl ∈ K for all l ∈ G. For any h ∈ H , there exists some σ ∈ Gal (Q (ωn) /K) such that η(σ) = h. Then χl(hI1) = χl(η(σ)I1) = σ (βl) = βl = χl(I1). 192 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 Let M = (χl (g))l,g∈G. We get Mδ′hI1 = Mδ ′ I1 . Note that M is invertible by the orthogonal relations of irreducible characters of G. So we have δ′hI1 = δ ′ I1 . This implies hI1 = I1. Since h is arbitrary, the result follows. Proposition 3.3. For the symbols in (3.1), we have γl ∈ K for all l ∈ G if and only if h(I2 \ I3) = I2 \ I3, h(I3 \ I2) = I3 \ I2 for all h ∈ H , where H = η (Gal (Q (ωn) /K)). Proof. Assume that h(I2 \ I3) = I2 \ I3, h(I3 \ I2) = I3 \ I2 for all h ∈ H . Then for any σ ∈ Gal (Q (ωn) /K), we have η(σ) ∈ H . Thus, for any l ∈ G, we have σ (γl) = χl (η(σ)(I2 \ I3))− χl (η(σ)(I3 \ I2)) = χl(I2 \ I3)− χl(I3 \ I2) = γl. It follows that γl ∈ K for all l ∈ G. Conversely, assume that γl ∈ K for all l ∈ G. For any h ∈ H , there exists some σ ∈ Gal (Q (ωn) /K) such that η(σ) = h. Then χl (η(σ)(I2 \ I3))− χl (η(σ)(I3 \ I2)) = σ (γl) = γl = χl(I2 \ I3)− χl(I3 \ I2). This means that χl (h(I2 \ I3))− χl (h(I3 \ I2)) = χl(I2 \ I3)− χl(I3 \ I2). Let M = (χl (g))l,g∈G. We get Mδ′h(I2\I3) −Mδ ′ h(I3\I2) = Mδ ′ I2\I3 −Mδ ′ I3\I2 . Note that M is invertible and I2 \I3 is disjoint with I3 \I2. So we have h(I2 \I3) = I2 \I3 and h(I3 \ I2) = I3 \ I2. As h is arbitrary, the result follows. Note that β0, γ0 ∈ Z, so we let L = K = Q (βl, γl | l ∈ G \ {0}) (3.2) and H ′ = {h ∈ Z∗n | hI1 = I1, h(I2 \ I3) = I2 \ I3, h(I3 \ I2) = I3 \ I2} . (3.3) Then we have the following result. Proposition 3.4. Using the symbols in (3.2) and (3.3), we have H ′ = η (Gal (Q (ωn) /L)). Proof. By Propositions 3.2 and 3.3, it is clear that η (Gal (Q (ωn) /L)) ⊆ H ′. Now we prove H ′ ⊆ η (Gal (Q (ωn) /L)) . Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 193 For each h′ ∈ H ′, let σ = η−1(h′). It follows that h′ = η(σ). For any l ∈ G, we have σ (βl) = χl(η(σ)I1) = χl(h ′I1) = χl(I1) = βl. Similarly, for any l ∈ G, σ (γl) = χl (η(σ)(I2 \ I3))− χl (η(σ)(I3 \ I2)) = χl(I2 \ I3)− χl(I3 \ I2) = γl. Hence σ ∈ Gal (Q (ωn) /L) and h′ = η(σ) ∈ η (Gal (Q (ωn) /L)) . Thus the result follows. Since H ′ is a subgroup of Z∗n, by Proposition 3.4, we have L = Q (ωn)η −1(H′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′)}. (3.4) Considering H ′ acting on G, assume that H ′g(1), H ′g(2), . . . ,H ′g(k) are all distinct orbits of H ′ on G, where g(i) ∈ G. Let C = { g(i) | G ∩H ′g(i) ̸= ∅ } . (3.5) Let M be the subgroup of L×/L×2 generated by all [γl]L for l ∈ C. Explicitly, M = ⟨[γl]L | l ∈ C⟩ . (3.6) Now we are ready to prove our main result. Theorem 3.5. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. Then the splitting field of Γ is L (√ γl | l ∈ C ) and the algebraic degree of Γ satisfies deg(Γ) = φ(n)|M | |H ′| , where γl, H ′, L, C,M are given in (3.1) and (3.3) – (3.6), respectively. Proof. If a, b are in the same orbit H ′g(i), then there exists h ∈ H ′ such that b = ha. It follows that γb = χb(I2 \ I3)− χb(I3 \ I2) = χha(I2 \ I3)− χha(I3 \ I2) = χa (h(I2 \ I3))− χa (h(I3 \ I2)) = χa(I2 \ I3)− χa(I3 \ I2) = γa. Therefore, there are at most |C| different elements in {γl | l ∈ G}. Set F = L (√ γl | l ∈ C ) . Note that F = Q ( βl + √ γl, βl − √ γl | l ∈ G ) . So the first assertion follows. By Lemma 2.4, deg(Γ) = [F : Q] = [F : L][L : Q] = [Q (ωn) : Q] [F : L] [Q (ωn) : L] = φ(n)|M | |H ′| . This completes the proof. 194 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 It is not easy to calculate |M |, but apparently 1 ≤ |M | ≤ 2|C|, thus we have Corollary 3.6. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. Then the algebraic degree of Γ satisfies φ(n) |H ′| ≤ deg(Γ) ≤ φ(n)2 |C| |H ′| , where H ′, C are given in (3.3) and (3.5), respectively. Remark 3.7. Theorem 3.5 and Corollary 3.6 still hold for a 2-Cayley graph. Indeed, we just need to restrict Tij = T−1ji for all 1 ⩽ i, j ⩽ 2, and modify the associated multi-sets I1, I2 and I3. The next two examples tell us that both the lower and upper bound in Corollary 3.6 are sharp. Example 3.8. Let Γ = Cay (Z3, Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley graph over G = Z3. Let T11 = T22 = {1, 2} and T12 = {1} and T21 = {2}. Then I1 = 2∗{1, 2}, I2 \I3 = 4∗{0} and I3 \ I2 = ∅. It follows that H ′ = {1, 2} = Z∗3, L = Q and γl = 4 for all l ∈ Z3. Thus |M | = 1 and deg(Γ) = φ(3)|H′| = 1. In fact, Spec(Γ) = 2 ∗ {−2, 0} ∪ {1, 3}. Example 3.9. Let Γ = Cay (Z4, Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over G = Z4. Let T11 = {1, 2} and T12 = {1}. Let T21 = {2} and T22 = {3}. Then I1 = {1, 2, 3}, I2 \ I3 = 6 ∗ {3} ∪ 2 ∗ {2} and I3 \ I2 = 2 ∗ {1} ∪ {0}. It follows that H ′ = {1} and C = Z4. By Corollary 3.6, deg(Γ) ≤ 25 = 32. In fact, L = Q (i) and F = L (√ 5, √ −8i− 3, √ −3, √ 8i− 3 ) . Obviously, deg(Γ) = 32. Observe that L = Q if and only if |H ′| = φ(n), as an application of Theorem 3.5, the next corollary provides a class of integral 2-Cayley digraphs over abelian groups. Corollary 3.10. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. If H ′ = Z∗n and γl is a square of an integer for each l ∈ C, where γl, H ′, C are given in (3.1), (3.3) and (3.5), respectively, then Γ is integral. Sometimes, we need not to compute |M | in Theorem 3.5. Corollary 3.11. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. If T11 = T22 and T12 = T−112 = T21, then the splitting field of Γ satis- fies SF(Γ) = Q (ωn)η −1(H′′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′′)}, the algebraic degree of Γ satisfies deg(Γ) = φ(n) |H ′′| , where H ′′ = {h ∈ Z∗n | hT11 = T11, hT12 = T12}. Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 195 Proof. Since I1 = [t | t ∈ 2 ∗ T11], I2 \ I3 = [ t | t ∈ 4 ∗ (T12 + T−112 ) ] and I3 \ I2 = ∅, we have γl = 4χl(T12 + T−112 ) = 4|χl(T12)|2 = 4χl(T12)2. Note that T12 = T −1 12 . So χl(T12) is a real number. It follows that √ γl = 2χl(T12) or √ γl = −2χl(T12). The rest of the proof is similar to that of Theorem 3.5. Corollary 3.12. Let Γ = Cay (G,Tij | 1 ⩽ i, j ⩽ 2) be a 2-Cayley digraph over an abelian group G of order n. If T11 = T22, T12 = T−112 = T21, and hT11 = T11, hT12 = T12 for all h ∈ Z∗n, then Γ is integral. 4 Some applications 4.1 Cayley digraphs over groups admitting an abelian subgroup of index 2 A Cayley digraph over a finite group G with a subgroup of index 2 is a 2-Cayley digraph, as the following result shows. Lemma 4.1 ([1]). Let Γ = Cay(G,S) be a Cayley (di)graph. Suppose that there exists a subgroup N of G with index 2. If {x1, x2} is a left transversal to N in G, then Γ ∼= Cay (N,Sij | 1 ⩽ i, j ⩽ 2), where Sij = {a ∈ N | x−1j axi ∈ S} = N ∩ xjSx −1 i . Let A be a finite abelian group of order n ≥ 3. Let f ∈ Aut(A) be of order 2. Let y ∈ A be such that f(y) = y. Let G be a non-abelian finite group admitting an abelian subgroup A of index 2. Then G admits a presentation G = 〈 A, x | x2 = y, xax−1 = f(a), a ∈ A 〉 . Observe that G = A ∪ xA and B = { f(a)a−1 | a ∈ A } is a subgroup of A. In particular, if f(a) = a−1 for a ∈ A, then B = A2 and y2 = e, where e is the identity of A. If y = e, then G is the generalized dihedral group Dih(A), with the presentation Dih(A) = 〈 A, x | x2 = e, xax−1 = a−1, a ∈ A 〉 . If y ̸= e (and so n = |A| is even), then G is the generalized dicyclic group Dic(A, y), with the presentation Dic(A, y) = 〈 A, x | x2 = y, xax−1 = a−1, a ∈ A 〉 . As the group operation here is multiplication, we assume that A = ⟨a1⟩n1⊗· · ·⊗⟨ar⟩nr and Irr(A) = {χl | (al11 , . . . , alrr ) ∈ A}, where l = (l1, . . . , lr). In this subsection, we always assume that G is a group admitting an abelian subgroup A of order n and of index 2. As an application of Theorem 3.5, we consider the algebraic degree of the Cayley digraph Γ = Cay(G,S). Note that A ∼= A′ = Zn1 ⊕ · · · ⊕ Znr . It is worth pointing out that the group operation here should correspond to the addition in Section 3. By Lemmas 2.3 and 4.1, we get the following result. Lemma 4.2. Let Γ = Cay(G,S) be a Cayley digraph and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. Then Γ has eigenvalues χl(T11) + χl(T22)± √ (χl(T11)− χl(T22))2 + 4χl(T21)χl(T12) 2 , l = (l1, . . . , lr) ∈ A′, 196 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 where n = ∏r i=1 ni, Tij = { t = (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xjSx−1i } and A′ = Zn1 ⊕ · · · ⊕ Znr . Using the symbols in Lemma 4.2, in a similar way as in Section 3, we define I1 = [t | t ∈ T11 or t ∈ T22] , I2 = [t | t ∈ (T11 + T11) or t ∈ (T22 + T22) or t ∈ 4 ∗ (T12 + T21)], I3 = [t | t ∈ 2 ∗ (T11 + T22)]. Let βl = χl(I1) and γl = χl(I2 \ I3)− χl(I3 \ I2). (4.1) Let η : Gal (Q (ωn) /Q) → Z∗n be the isomorphism such that σ (ωn) = ω η(σ) n , where σ ∈ Gal (Q (ωn) /Q). Let L = Q (βl, γl | l ∈ A′ \ {0}) (4.2) and H ′ = {h ∈ Z∗n | hI1 = I1, h(I2 \ I3) = I2 \ I3, h(I3 \ I2) = I3 \ I2} . (4.3) Since Γ ∼= Cay (A′, Tij | 1 ⩽ i, j ⩽ 2), by Proposition 3.4, we have the following result. Proposition 4.3. Using the symbols in (4.2) and (4.3), we have H ′ = η (Gal (Q (ωn) /L)). Now we consider H ′ acting on A′ = Zn1⊕· · ·⊕Znr . Assume that H ′a(1), H ′a(2), . . . , H ′a(k) are all distinct orbits of H ′ on A′, where a(i) ∈ A′. Let C = { a(i) | A′ ∩H ′a(i) ̸= ∅ } (4.4) and M = ⟨[γl]L | l ∈ C⟩ . (4.5) Since Γ ∼= Cay (A′, Tij | 1 ⩽ i, j ⩽ 2), using the conclusions in Section 3, we imme- diately get the following results. Theorem 4.4. Let Γ = Cay(G,S) be a Cayley digraph, and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. Then the splitting field of Γ is L (√ γl | l ∈ C ) and the algebraic degree of Γ satisfies deg(Γ) = φ(n)|M | |H ′| , where L = Q (ωn)η −1(H′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′)} and γl, H ′, C,M are given in (4.1) and (4.3) – (4.5), respectively. Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 197 Corollary 4.5. Let Γ = Cay(G,S) be a Cayley digraph, and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. Then the algebraic degree of Γ satisfies φ(n) |H ′| ≤ deg(Γ) ≤ φ(n)2 |C| |H ′| , where H ′, C are given in (4.3) and (4.4), respectively. Corollary 4.6. Let Γ = Cay(G,S) be a Cayley digraph, and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. If H ′ = Z∗n and γl is a square of an integer for each l ∈ C, where γl, H ′, C are given in (4.1), (4.3) and (4.4), respectively, then Γ is integral. Corollary 4.7. Let Γ = Cay(G,S) be a Cayley digraph, and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. Let Tij = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xjSx−1i } . If T11 = T22 and T12 = T−112 = T21, then the splitting field of Γ satisfies SF(Γ) = Q (ωn)η −1(H′′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′′)}, the algebraic degree of Γ satisfies deg(Γ) = φ(n) |H ′′| , where H ′′ = {h ∈ Z∗n | hT11 = T11, hT12 = T12}. Corollary 4.8. Let Γ = Cay(G,S) be a Cayley digraph, and A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr be an abelian subgroup of G of order n and of index 2 with left transversal {x1, x2}. Let Tij = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xjSx−1i } . If T11 = T22, T12 = T−112 = T21, and hT11 = T11, hT12 = T12 for all h ∈ Z∗n, then Γ is integral. 4.2 Cayley graphs over generalized dihedral groups In the following two subsections, we consider Cayley graphs but not digraphs. The gener- alized dihedral group Dih(A) is given by the following presentation Dih(A) = 〈 A, x | x2 = e, xax−1 = a−1, a ∈ A 〉 . Let Γ = Cay(Dih(A), S) be a Cayley digraph. Using the symbols in Subsection 4.1, note that A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr , and |A| = n, so |Dih(A)| = 2n. Without loss of generality, let x1 = e and x2 = x. Then T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ S } , T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } , T21 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ Sx } , T22 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xSx } . (4.6) For the algebraic degree of the digraph Γ, we just need to replace Tij given in Subsec- tion 4.1 with Tij given in (4.6), so we omit the details here. 198 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 We are now interested in the algebraic degree of the undirected Cayley graph Γ = Cay(Dih(A), S). Using the symbols in (4.6), as S = S−1, we have T−122 = T22 = T11 = T−111 and T −1 12 = T21. Let t = (t1, . . . , tr). In a similar way as in Subsection 4.1, we define I1 = [t | t ∈ T11 or t ∈ T22] , I2 = [t | t ∈ (T11 + T11) or t ∈ (T22 + T22) or t ∈ 4 ∗ (T12 + T21)], I3 = [t | t ∈ 2 ∗ (T11 + T22)]. It follows that I1 = [t | t ∈ 2 ∗ T11], I2 \ I3 = [ t | t ∈ 4 ∗ (T12 + T−112 ) ] and I3 \ I2 = ∅. In fact, by Lemma 4.2, the eigenvalues of the Cayley graph Γ = Cay(Dih(A), S) are χl(T11)± |χl(T12)| , l ∈ A′, where A′ = Zn1 ⊕ · · · ⊕ Znr . Note that |χl(T12)| = √ χl(T12)χl(T −1 12 ) =√ χl(T12 + T −1 12 ), the multi-sets I1 and I2 \I3 can be reduced to I ′1 = T11 and (I2 \I3)′ =[ t | t ∈ T12 + T−112 ] . Let βl = χl(I ′ 1) and γl = χl((I2 \ I3)′). (4.7) Let η : Gal (Q (ωn) /Q) → Z∗n be the isomorphism such that σ (ωn) = ω η(σ) n , where σ ∈ Gal (Q (ωn) /Q). Let L = Q (βl, γl | l ∈ A′ \ {0}) (4.8) and H ′ = {h ∈ Z∗n | hI ′1 = I ′1, h(I2 \ I3)′ = (I2 \ I3)′} . (4.9) Note that I1 = 2 ∗ I ′1 and I2 \ I3 = 4 ∗ (I2 \ I3)′. So we have the following result by Proposition 4.3. Proposition 4.9. Using the symbols in (4.8) and (4.9), we have H ′ = η (Gal (Q (ωn) /L)). Similarly, we consider H ′ acting on A′. Assume that H ′a(1), H ′a(2), . . . H ′a(k) are all distinct orbits of H ′ on A′, where a(i) ∈ A′. Let B = {x ∈ A′ | 2x = 0} and A′ = B ∪ E ∪ E−1, where B,E,E−1 are disjoint. Let C ′ = { a(i) | (B ∪ E) ∩H ′a(i) ̸= ∅ } (4.10) and M = ⟨[γl]L | l ∈ C ′⟩ . (4.11) Then the following results hold. Theorem 4.10. Let Γ = Cay(Dih(A), S) be a Cayley graph over the generalized dihe- dral group Dih(A) of order 2n. Then the splitting field of Γ is L (√ γl | l ∈ C ′ ) and the algebraic degree of Γ satisfies deg(Γ) = φ(n)|M | |H ′| , Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 199 where L = Q (ωn)η −1(H′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′)} and γl, H ′, C ′,M are given in (4.7) and (4.9) – (4.11), respectively. Proof. Since ((I2 \ I3)′)−1 = (I2 \ I3)′, it follows that γl = χl((I2 \ I3)′) = χl(((I2 \ I3)′)−1) = χ−l((I2 \ I3)′) = γ−l. Then the result follows from Theorem 4.4. Corollary 4.11. Let Γ = Cay(Dih(A), S) be a Cayley graph over the generalized dihedral group Dih(A) of order 2n. Then the algebraic degree of Γ satisfies φ(n) |H ′| ≤ deg(Γ) ≤ φ(n)2 |C′| |H ′| , where H ′, C ′ are given in (4.9) and (4.10), respectively. Corollary 4.12. Let Γ = Cay(Dih(A), S) be a Cayley graph over the generalized dihedral group Dih(A) of order 2n. If H ′ = Z∗n and γl is a square of an integer for each l ∈ C ′, where γl, H ′, C ′ are given in (4.7), (4.9) and (4.10), respectively, then Γ is integral. Corollary 4.13. Let Γ = Cay(Dih(A), S) be a Cayley graph over the generalized dihedral group Dih(A) of order 2n. Let T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ S } and T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } . If T12 = T−112 , then the splitting field of Γ satisfies SF(Γ) = Q (ωn)η −1(H′′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′′)}, the algebraic degree of Γ satisfies deg(Γ) = φ(n) |H ′′| , where H ′′ = {h ∈ Z∗n | hT11 = T11, hT12 = T12}. Corollary 4.14. Let Γ = Cay(Dih(A), S) be a Cayley graph over the generalized dihedral group Dih(A) of order 2n. Let T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ S } and T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } . If T12 = T−112 and for all h ∈ Z∗n, hT11 = T11 and hT12 = T12, then Γ is integral. In particular, let Dih(A) = D2n = ⟨a, b | an = b2 = e, bab = a−1⟩ be the dihedral group of order 2n. Then A′ = Zn, I ′1 = T11 = {t | at ∈ S}, T12 = {t | bat ∈ S} and (I2 \ I3)′ = [ t | t ∈ T12 + T−112 ] . Note that βl = χl(I ′ 1) and γl = χl((I2 \ I3)′), (4.12) where χl(t) = ωltn and 0 ≤ l ≤ n− 1. Furthermore, L = Q (βl, γl | 1 ≤ l ≤ n− 1) . We first try to simplify the expression of L. 200 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 Lemma 4.15. Let K be a field such that Q ⊆ K ⊆ Q (ωn). If β1, γ1 ∈ K, then βl, γl ∈ K for 1 ≤ l ≤ n− 1. Proof. For 1 ≤ l ≤ n − 1, let σl : Q (ωn) → Q (ωn) be defined by σl (ωn) = ωln. It is clear that σl is a homomorphism and βl = σl (β1) , γl = σl (γ1). Thus, for any σ ∈ Gal (Q (ωn) /K), we have σ (βl) = σ (σl (β1)) = σ σl ∑ t∈I′1 ωtn  = ∑ t∈I′1 ωη(σ)tln = σl (σ (β1)) = σl (β1) = βl. Similarly, σ (γl) = γl. Therefore, βl, γl ∈ K. By Lemma 4.15, we have L = Q (β1, γ1) . Let H ′ = {h ∈ Z∗n | hI ′1 = I ′1, h(I2 \ I3)′ = (I2 \ I3)′} . (4.13) Then H ′ = η (Gal (Q (ωn) /L)). Note that C ′ = { a(i) | {0, 1, . . . , ⌊n/2⌋} ∩H ′a(i) ̸= ∅ } (4.14) and M = ⟨[γl]L | l ∈ C ′⟩ , (4.15) where H ′a(1), H ′a(2), . . . H ′a(k) are all distinct orbits of H ′ on Zn. Consequently, we have the following corollaries. Corollary 4.16. Let Γ = Cay(D2n, S) be a Cayley graph over the dihedral group D2n. Then the splitting field of Γ is L (√ γl | l ∈ C ′ ) and the algebraic degree of Γ satisfies deg(Γ) = φ(n)|M | |H ′| , where L = Q (ωn)η −1(H′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′)} and γl, H ′, C ′,M are given in (4.12) – (4.15), respectively. Corollary 4.17. Let Γ = Cay(D2n, S) be a Cayley graph over the dihedral group D2n. Then the algebraic degree of Γ satisfies φ(n) |H ′| ≤ deg(Γ) ≤ φ(n)2 |C′| |H ′| , where H ′, C ′ are given in (4.13) and (4.14), respectively. Corollary 4.18. Let Γ = Cay(D2n, S) be a Cayley graph over the dihedral group D2n. Let T11 = {t | at ∈ S} and T12 = {t | bat ∈ S}. If T12 = T−112 , then the splitting field of Γ satisfies SF(Γ) = Q (ωn)η −1(H′′) = {x ∈ Q (ωn) | σ(x) = x for all σ ∈ η−1(H ′′)}, Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 201 the algebraic degree of Γ satisfies deg(Γ) = φ(n) |H ′′| , where H ′′ = {h ∈ Z∗n | hT11 = T11, hT12 = T12}. There are results similar to Corollaries 4.12 and 4.14 as well, we omit them here. We end this subsection with the following example. Example 4.19. Let D16 = ⟨a, b | a8 = b2 = e, bab = a−1⟩ be the dihedral group of order 16 and S = { a, a7, b } . We consider the algebraic degree of Γ = Cay(D16, S). Then T11 = {1,−1} and T12 = {0} = T−112 . It follows that H ′′ = {1,−1} ≤ Z∗8. By Corollary 4.18, SF(Γ) = Q (ω8)η −1(H′′) = Q( √ 2) and deg(Γ) = φ(8)|H′′|=2. In fact, Spec(Γ) = 2 ∗ { √ 2 + 1, √ 2− 1,− √ 2 + 1,− √ 2− 1} ∪ 3 ∗ {1,−1} ∪ {−3, 3}. 4.3 Cayley graphs over generalized dicyclic groups For the generalized dicyclic group Dic(A, y), it has the following presentation Dic(A, y) = 〈 A, x | x2 = y, xax−1 = a−1, a ∈ A 〉 . We put our focus on the algebraic degree of the Cayley graph Γ = Cay(Dic(A, y), S). Using the symbols in Subsection 4.1, since A = ⟨a1⟩n1 ⊗ · · · ⊗ ⟨ar⟩nr , and |A| = n is even, say n = 2m, then |Dic(A, y)| = 4m. Let x1 = e and x2 = x. Then T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ S } , T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } , T21 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ Sx−1 } , T22 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xSx−1 } . Since S = S−1, we have T−122 = T22 = T11 = T −1 11 and T −1 12 = T21. Let t = (t1, . . . , tr). By similar arguments as those in Subsection 4.2, we just need to consider I ′1 = T11 and (I2 \ I3)′ = [ t | t ∈ T12 + T−112 ] . Let βl = χl(I ′ 1) and γl = χl((I2 \ I3)′). (4.16) Let η : Gal (Q (ω2m) /Q) → Z∗2m be the isomorphism such that σ (ω2m) = ω η(σ) 2m , where σ ∈ Gal (Q (ω2m) /Q). Let L = Q (βl, γl | l ∈ A′ \ {0}) and H ′ = {h ∈ Z∗2m | hI ′1 = I ′1, h(I2 \ I3)′ = (I2 \ I3)′} . (4.17) By Proposition 4.3, we have H ′ = η (Gal (Q (ω2m) /L)). Also, we consider H ′ acting on A′ = Zn1 ⊕· · ·⊕Znr . Assume that H ′a(1), H ′a(2), . . . H ′a(k) are all distinct orbits of H ′ on A′, where a(i) ∈ A′. Let B = {x ∈ A′ | 2x = 0} and A′ = B ∪ E ∪ E−1, where B,E,E−1 are disjoint. Let C ′ = { a(i) | (B ∪ E) ∩H ′a(i) ̸= ∅ } (4.18) 202 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 and M = ⟨[γl]L | l ∈ C ′⟩ . (4.19) In a similar way as in Theorem 4.10, we get the following result. Theorem 4.20. Let Γ = Cay(Dic(A, y), S) be a Cayley graph over Dic(A, y) of order 4m. Then the splitting field of Γ is L (√ γl | l ∈ C ′ ) , and the algebraic degree ofΓ satisfies deg(Γ) = φ(2m)|M | |H ′| , where L = Q (ω2m)η −1(H′) = {x ∈ Q (ω2m) | σ(x) = x for all σ ∈ η−1(H ′)} and γl, H ′, C ′,M are given in (4.16) – (4.19), respectively. Corollary 4.21. Let Γ = Cay(Dic(A, y), S) be a Cayley graph over Dic(A, y) of order 4m. Then the algebraic degree of Γ satisfies φ(2m) |H ′| ≤ deg(Γ) ≤ φ(2m)2 |C′| |H ′| , where H ′, C ′ are given in (4.17) and (4.18), respectively. Corollary 4.22. Let Γ = Cay(Dic(A, y), S) be a Cayley graph over Dic(A, y) of order 4m. If H ′ = Z∗2m and γl is a square of an integer for each l ∈ C ′, where γl, H ′, C ′ are given in (4.16) – (4.18), respectively, then Γ is integral. Corollary 4.23. Let Γ = Cay(Dic(A, y), S) be a Cayley graph over Dic(A, y) of order 4m. Let T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈S } and T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } . If T12 = T−112 , then the splitting field of Γ satisfies SF(Γ) = Q (ω2m)η −1(H′′) = {x ∈ Q (ω2m) | σ(x) = x for all σ ∈ η−1(H ′′)}, the algebraic degree of Γ satisfies deg(Γ) = φ(2m) |H ′′| , where H ′′ = {h ∈ Z∗2m | hT11 = T11, hT12 = T12}. Corollary 4.24. Let Γ = Cay(Dic(A, y), S) be a Cayley graph over Dic(A, y) of order 4m. Let T11 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ S } and T12 = { (t1, . . . , tr) | ( at11 , . . . , a tr r ) ∈ xS } . If T12 = T−112 and for all h ∈ Z∗2m, hT11 = T11 and hT12 = T12, then Γ is integral. Furthermore, for the dicyclic group Dic4m = 〈 a, b | a2m = e, am = b2, b−1ab = a−1 〉 , as direct consequences of Theorem 4.20 and Corollaries 4.21 – 4.24, we have similar re- sults, so we omit the details here. Example 4.25. Let Dic12 = ⟨a, b | a6 = e, a3 = b2, b−1ab = a−1⟩ be the dicyclic group of order 12 and S = { a, a5, ab, a2b, a4b, a5b } . We consider the algebraic degree of Γ = Cay(Dic12, S). Then T11 = {1, 5} and T12 = {1, 2, 4, 5} = T−112 . It follows that H ′′ = {1, 5} = Z∗6. By Corollary 4.24, deg(Γ) = 1. In fact, Spec(Γ) = 4 ∗ {−1, 1} ∪ 3 ∗ {−2} ∪ {6}. Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 203 4.4 Cayley digraphs over semi-dihedral groups For the semi-dihedral group SD8m, it has the following presentation SD8m = ⟨a, b | a4m = b2 = e, bab = a2m−1⟩. We now consider the algebraic degree of the Cayley digraph Γ = Cay(SD8m, S). Using the symbols in Subsection 4.1, it follows that A = ⟨a⟩4m and A′ = Z4m. Let x1 = e and x2 = b. Then T11 = { t | at ∈ S } , T12 = { t | bat ∈ S } , T21 = { t | atb ∈ S } , T22 = { t | a(2m−1)t ∈ S } . In a similar way as in Subsection 4.1, we define I1 = [t | t ∈ T11 or t ∈ T22] , I2 = [t | t ∈ (T11 + T11) or t ∈ (T22 + T22) or t ∈ 4 ∗ (T12 + T21)], I3 = [t | t ∈ 2 ∗ (T11 + T22)]. Let βl = χl(I1) and γl = χl(I2 \ I3)− χl(I3 \ I2), (4.20) where χl(t) = ωlt4m and 0 ≤ l ≤ 4m − 1. Let K be a field such that Q ⊆ K ⊆ Q (ω4m). Then Gal (Q (ω4m) /K)) ≤ Gal (Q (ω4m) /Q) ∼= Z∗4m. Let η : Gal (Q (ω4m) /Q) → Z∗4m be the isomorphism such that σ (ω4m) = ω η(σ) 4m , where σ ∈ Gal (Q (ω4m) /Q). Let L = Q (βl, γl | 1 ≤ l ≤ 4m− 1) . The following lemma helps to simplify the expression of L. Lemma 4.26. If β1, γ1 ∈ K, then βl, γl ∈ K for 1 ≤ l ≤ 4m− 1. Proof. The proof is similar to that of Lemma 4.15. By Lemma 4.26, we have L = Q (β1, γ1) . Let H ′ = {h ∈ Z∗4m | hI1 = I1, h(I2 \ I3) = I2 \ I3, h(I3 \ I2) = I3 \ I2} . (4.21) By Proposition 4.3, we have H ′ = η (Gal (Q (ω4m) /L)). Assume that H ′a(1), H ′a(2), . . . H ′a(k) are all distinct orbits of H ′ on Z4m. Let C = {a(i) | Z4m ∩H ′a(i) ̸= ∅} (4.22) and M = ⟨[γl]L | l ∈ C⟩ . (4.23) By Theorem 4.4 and Corollaries 4.5 – 4.8, we have 204 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 Theorem 4.27. Let Γ = Cay(SD8m, S) be a Cayley digraph over the semi-dihedral group SD8m. Then the splitting field of Γ is L (√ γl | l ∈ C ) and the algebraic degree of Γ satisfies deg(Γ) = φ(4m)|M | |H ′| , where L = Q (ω4m)η −1(H′) = {x ∈ Q (ω4m) | σ(x) = x for all σ ∈ η−1(H ′)} and γl, H ′, C,M are given in (4.20) – (4.23), respectively. Corollary 4.28. Let Γ = Cay(SD8m, S) be a Cayley digraph over the semi-dihedral group SD8m. Then the algebraic degree of Γ satisfies φ(4m) |H ′| ≤ deg(Γ) ≤ φ(4m)2 |C| |H ′| , where H ′, C are given in (4.21) and (4.22), respectively. Corollary 4.29. Let Γ = Cay(SD8m, S) be a Cayley digraph over the semi-dihedral group SD8m. If H ′ = Z∗4m and γl is a square of an integer for each l ∈ C, where γl, H ′, C are given in (4.20) – (4.22), respectively, then Γ is integral. Corollary 4.30. Let Γ = Cay(SD8m, S) be a Cayley digraph over the semi-dihedral group SD8m. If T11 = T22 and T12 = T−112 = T21, then the splitting field of Γ satisfies SF(Γ) = Q (ω4m)η −1(H′′) = {x ∈ Q (ω4m) | σ(x) = x for all σ ∈ η−1(H ′′)}, the algebraic degree of Γ satisfies deg(Γ) = φ(4m) |H ′′| , where H ′′ = {h ∈ Z∗4m | hT11 = T11, hT12 = T12}. Corollary 4.31. Let Γ = Cay(SD8m, S) be a Cayley digraph over the semi-dihedral group SD8m. If T11 = T22, T12 = T−112 = T21 and for all h ∈ Z∗4m, hT11 = T11 and hT12 = T12, then Γ is integral. We end this paper with the following example. Example 4.32. Let SD16 = ⟨a, b | a8 = b2 = e, b−1ab = a3⟩ be the semi-dihedral group of order 16 and S = { a2, a6, ba, ba5 } . We consider the algebraic degree of Γ = Cay(SD16, S). Then T11 = T22 = {2, 6} and T12 = {1, 5}, T−112 = T21 = {3, 7}. Thus, I1 = 2 ∗ {2, 6}, I2 \ I3 = 8 ∗ {0, 4} and I3 \ I2 = ∅. It follows that H ′ = {1, 3, 5, 7} = Z∗8. Since γ1 = 8[1 + (−1)l] is a square of integer for each l ∈ Z8, by Corollary 4.29, deg(Γ) = 1. In fact, Spec(Γ) = 10 ∗ {0} ∪ 2 ∗ {−2, 1, 4}. ORCID iDs Lihua Feng https://orcid.org/0000-0003-4144-1649 Y. Wu et al.: Algebraic degrees of 2-Cayley digraphs over abelian groups 205 References [1] M. Arezoomand and B. Taeri, On the characteristic polynomial of n-Cayley digraphs, Electron. J. Comb. 20 (2013), research paper p57, 14, doi:10.37236/3105, https://doi.org/10. 37236/3105. [2] A. Behajaina and F. Legrand, On integral mixed cayley graphs over non-abelian finite groups admitting an abelian subgroup of index 2, 2022, arXiv:2203.08793 [math.CO]. [3] T. Cheng, L. Feng and H. Huang, Integral Cayley graphs over dicyclic group, Linear Alge- bra Appl. 566 (2019), 121–137, doi:10.1016/j.laa.2019.01.002, https://doi.org/10. 1016/j.laa.2019.01.002. [4] T. Cheng, L. Feng, W. Liu, L. Lu and D. Stevanović, Distance powers of integral Cayley graphs over dihedral groups and dicyclic groups, Linear Multilinear Algebra 70 (2022), 1281–1290, doi:10.1080/03081087.2020.1758609, https://doi.org/10. 1080/03081087.2020.1758609. [5] X. Gao, H. Lü and Y. Hao, The Laplacian and signless Laplacian spectrum of semi-Cayley graphs over abelian groups, J. Appl. Math. Comput. 51 (2016), 383–395, doi:10.1007/ s12190-015-0911-9, https://doi.org/10.1007/s12190-015-0911-9. [6] X. Gao and Y. Luo, The spectrum of semi-Cayley graphs over abelian groups, Linear Alge- bra Appl. 432 (2010), 2974–2983, doi:10.1016/j.laa.2009.12.040, https://doi.org/10. 1016/j.laa.2009.12.040. [7] P. Guillot, A gentle course in local class field theory. Local number fields, Brauer groups, Ga- lois cohomology, Cambridge University Press, Cambridge, 2018, doi:10.1017/9781108377751, https://doi.org/10.1017/9781108377751. [8] F. Li, Circulant digraphs integral over number fields, Discrete Math. 313 (2013), 821–823, doi: 10.1016/j.disc.2012.12.025, https://doi.org/10.1016/j.disc.2012.12.025. [9] F. Li, A method to determine algebraically integral Cayley digraphs on finite abelian group, Contrib. Discrete Math. 15 (2020), 148–152, doi:10.11575/cdm.v15i2.62327, https:// doi.org/10.11575/cdm.v15i2.62327. [10] X. Liu and S. Zhou, Eigenvalues of Cayley graphs, Electron. J. Comb. 29 (2022), research paper p2.9, 164, doi:10.37236/8569, https://doi.org/10.37236/8569. [11] L. Lu, Q. Huang and X. Huang, Integral Cayley graphs over dihedral groups, J. Algebr. Comb. 47 (2018), 585–601, doi:10.1007/s10801-017-0787-x, https://doi.org/10. 1007/s10801-017-0787-x. [12] L. Lu and K. Mönius, Algebraic degree of Cayley graphs over abelian groups and dihedral groups, J. Algebr. Comb. (2023), doi:10.1007/s10801-022-01190-7, https://doi.org/ 10.1007/s10801-022-01190-7. [13] K. Mönius, The algebraic degree of spectra of circulant graphs, J. Number Theory 208 (2020), 295–304, doi:10.1016/j.jnt.2019.08.002, https://doi.org/10.1016/j.jnt.2019. 08.002. [14] K. Mönius, Splitting fields of spectra of circulant graphs, J. Algebra 594 (2022), 154– 169, doi:10.1016/j.jalgebra.2021.11.036, https://doi.org/10.1016/j.jalgebra. 2021.11.036. [15] G. Sabidussi, Vertex-transitive graphs, Monatsh. Math. 68 (1964), 426–438, http:// eudml.org/doc/177267. [16] N. Sripaisan and Y. Meemark, Algebraic degree of spectra of Cayley hypergraphs, Discrete Appl. Math. 316 (2022), 87–94, doi:10.1016/j.dam.2022.03.029, https://doi.org/10. 1016/j.dam.2022.03.029. 206 Ars Math. Contemp. 24 (2024) #P2.02 / 187–206 [17] B. Steinberg, Representation Theory of Finite Groups, Springer, New York, 2009. [18] L. L. X.Y. Huang and K. Mönius, Splitting fields of mixed cayley graphs over abelian groups, 2022, arXiv:2202.00987 [math.CO]. [19] H. Zou and J. Meng, Some algebraic properties of Bi-Cayley graphs, Acta Math. Sin., Chin. Ser. 50 (2007), 1075–1080. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.03 / 207–230 https://doi.org/10.26493/1855-3974.2894.b07 (Also available at http://amc-journal.eu) A non-associative incidence near-ring with a generalized Möbius function* John Johnson †, Max Wakefield ‡ US Naval Academy, 572-C Holloway Rd, Annapolis MD, 21402 USA This paper is dedicated to the memory of John Johnson. Received 1 June 2022, accepted 27 February 2023, published online 20 September 2023 Abstract There is a convolution product on 3-variable partial flag functions of a locally finite poset that produces a generalized Möbius function. Under the product this generalized Möbius function is a one sided inverse of the zeta function and satisfies many generaliza- tions of classical results. In particular we prove analogues of Phillip Hall’s Theorem on the Möbius function as an alternating sum of chain counts, Weisner’s Theorem, and Rota’s Crosscut Theorem. A key ingredient to these results is that this function is an overlapping product of classical Möbius functions. Using this generalized Möbius function we define analogues of the characteristic polynomial and Möbius polynomials for ranked lattices. We compute these polynomials for certain families of matroids and prove that this generalized Möbius polynomial has -1 as root if the matroid is modular. Using results from Ardila and Sanchez we prove that this generalized characteristic polynomial is a matroid valuation. Keywords: Incidence algebra, matroid, Möbius function, valuation. Math. Subj. Class. (2020): 37K15, 42A99, 60E05, 05A17 *The authors are very thankful for detailed comments by the reviewer. The reviewers suggestions have signif- icantly improved the article. The authors are thankful for discussions with Carolyn Chun, Joel Lewis, and Will Traves. The authors are also thankful to George Andrews for help on Lemma 6.14. Frederico Ardila and Mario Sanchez significantly helped with the material on valuations for which the authors are very thankful. Also, Jose Bastidas made multiple excellent comments for which the authors are very thankful. The authors would like to thank the US Naval Academy trident program for support during this project. †Supported by the US Naval Academy as a Trident Scholar. ‡Corresponding author. E-mail addresses: m213162@usna.edu (John Johnson), wakefiel@usna.edu (Max Wakefield) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 208 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 1 Introduction Combinatorial invariants in incidence algebras play a central role in many areas of com- binatorics as well as in number theory, algebraic topology, algebraic geometry, and repre- sentation theory. In particular, the Möbius function appears in the inverse of the Riemann zeta function as well as the coefficients of the chromatic polynomial for graphs. In this note we study a generalization of the classical incidence algebra by looking at three vari- able incidence functions. A large portion of this study is focussed on studying a 3-variable generalized Möbius function inside this generalized incidence structure. Incidence algebras and Möbius functions were popularized by Rota in [26]. Rota char- acterized the classical Möbius function from number theory (see [20] and [14]) as the inverse of the constant function 1 on the intervals of the poset which is called the zeta function. In [26] Rota gives many results on the Möbius function, including his Crosscut Theorem. Since then, many advances can be attributed to Möbius functions. Of particular importance are the counting theorems of Zaslavsky in [33] and Terao’s factorization the- orem (see [29]) using the Möbius function in the form of the characteristic polynomial of a hyperplane arrangement. The main motivation for this work is to build invariants which are finer than the classical Möbius function and characteristic polynomial to obtain more information about the underlying combinatorial structure. More recently, there has been considerable developments in understanding of some classical invariants on matroids. One generalization came from Krajewski, Moffatt, and Tanasa who built Tutte polynomials from a Hopf algebra in [18]. Taking this a little fur- ther, in [11] Dupont, Fink and Moci construct a categorical framework to view various combinatorial invariants and they prove some convolution formulas. The work of Aguiar and Ardila in [1] framed many combinatorial structures like matroids in terms of general- ized permutahedra, where there is a natural Hopf monoid governing classical operations. One possible starting place for this study could be the work of Joni and Rota in [16]. Then, in [6], Ardila and Sanchez use this Hopf monoid structure to build a concrete method for investigating valuations on many combinatorial structures. Another aim of this study is to add another invariant to the list of valuations. Concretely, we use the methods of Ardila and Sanchez to show that one of our invariants is a valuation on matroids. One view that one can take for many combinatorial structures is that of posets (e.g. matroids are geometric lattices) and this is the view that we take here. The starting point for our study is the collection of 3-variable functions on ordered triples of elements in a poset. The set of these 3-variable functions also appears in the book [2] by Aguiar and Mahajan in Appendix C4 where they study 2-cochains and 2-cocycles. We differ from the work in Appendix C4 [2] by equipping this set of functions with a spe- cial convolution product. The motivation for this product comes from trying to symmetrize a more natural convolution product that was studied by the second author in [30] as well as making new invariants with special properties. This product provides a 3-variable Möbius function which is a sort of left inverse of the 3-variable analogue of the zeta function. We call this function the J-function and study many of its properties. It turns out that it is essentially a staggered product of the classical Möbius functions and hence satisfies gen- eralizations of many of the classical theorems on the classical Möbius function. To prove these results we develop and use certain operations and formulas these 3-variable func- tions satisfy that give maps between various different types of incidence algebras. In [8] Jose Bastidas studies Type B Hopf monoids and defines an antipode via some convolution formulas which seem to have some similar properties to the work presented here. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 209 As an application we build two different polynomials from the J-function: a general- ized characteristic polynomial and a generalized Möbius polynomial (see [17] and [21] for Möbius polynomials). It turns out that these polynomials have some interesting properties that are not apparent from the surface. In the case of matroids, the generalized character- istic polynomial has positive coefficients. We then compute these polynomials for certain families of matroids and find special roots. Of particular interest is that the generalized Möbius function has −1 as a root for modular matroids, which mimics Theorem 1 in [21]. However, we show that the converse is not true and so one is led to question what do these polynomial count? Could there be some chromatic generalization for the generalized poly- nomials, or some lattice point or finite field counting formula for these polynomials (like [9] or [7])? Also, in [8], Bastidas defines some polynomial invariants via characters of a Hopf monoid. Can the polynomials we define here be put in the framework of [8]? We finish by employing the methods of Ardila and Sanchez in [6] to show that our generalized characteristic polynomial is a matroid valuation. This follows from the fact that the J-function splits as a product of Möbius functions. In the case of the Möbius polynomial, we are not sure whether or not it is a valuation, yet we show that it does have a decomposition in terms of the classical characteristic polynomials. We find it interest- ing that this decomposition looks very similar to the recursive definition of the matroid Kazhdan-Lusztig polynomial originally defined in [12]. We begin this study with reviewing classical results on incidence algebras and Möbius functions in Section 2. Then we define our 3-variable incidence structure in Section 3. There we show that this structure has some interesting properties but that it is neither as- sociative nor distributive. However, in Section 4 we develop multiple operations which give nice formulas between these different kinds of incidence functions. Using these for- mulas we define a generalized Möbius function, the J-function, and study its properties in Section 5. Finally in Section 6 we define our generalized characteristic and Möbius polynomials. 2 Incidence Algebras Let R be a commutative ring and P be a locally finite poset. We follow [28] and [4] for combinatorics on posets. For the remainder of this note we refer to the order in P by ≤. Also, for n ∈ N let [n] = {1, 2, 3, . . . , n}. In this section we review basic material of incidence algebras where we follow [27]. First we define the poset of partial flags. Definition 2.1. The poset of partial flags of length k on P is F lk(P) = { (x1, x2, . . . , xk) ∈ Pk| x1 ≤ x2 ≤ · · · ≤ xk } with order given by (x1, . . . , xk) ⪯ (y1, . . . , yk) if and only if for all i ∈ [k] we have xi ≤ yi. Now we define the classical incidence algebras. Definition 2.2. The incidence algebra on P is the set I(P, R) = Hom(F l2(P), R) where R is a commutative ring. Addition in I(P, R) is given by (f + g)(x, y) = f(x, y) + g(x, y), 210 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 the multiplication is given by convolution (f ∗ g)(x, y) = ∑ x≤a≤b f(x, a)g(a, y), and the scalar product is given by (rf)(x, y) = rf(x, y) for all r ∈ R. In this note, we will examine multiple different operations on functions on posets. For this reason we will reserve juxtaposition only for products of elements in the ring R. Oth- erwise we will denote products of functions with specific operation names like ∗. It turns out that I(P, R) is a non-commutative R-algebra with identity element given by the Kronecker delta function δ(x, y) = { 1 if x = y, 0 else. There are two other very important elements in I(P, R). Definition 2.3. The zeta function ζ ∈ I(P, R) is defined as the constant function on F l2(P) ζ(x, y) = 1 for all (x, y) ∈ F l2(P). The Möbius function µ ∈ I(P, R) is defined by∑ x≤a≤y µ(x, a) = ∑ x≤a≤y µ(a, y) = δ(x, y) for all (x, y) ∈ F l2(P). The Möbius function was originally defined by Möbius (see [20]) on the poset of the natural numbers ordered by division for the purpose of inverting the Riemann zeta function. Since then the Möbius function has been used in many different contexts and broadened by the work of Rota in [26]. For our discussion, it is important to note that µ is the multiplica- tive inverse of the zeta function µ ∗ ζ = ζ ∗ µ = δ. Now we review how the incidence algebra functor factors over products. Recall that for posets P and Q the product poset is P ×Q with order given by (x1, x2) ≤ (y1, y2) if and only if x1 ≤ y1 and x2 ≤ y2. Proposition 2.4 (Proposition 2.1.12 [27]). If P and Q are locally finite posets then I(P, R)⊗R I(Q, R) ∼= I(P ×Q, R). Because of Proposition 2.4 we define the following operation on functions. In order to the make the exposition clear in the case when we are dealing with functions over different posets, we will put the poset in the subscript. For fP ∈ I(P, R) and gQ ∈ I(Q, R) define fP × gQ ∈ I(P ×Q, R) by (fP × gQ)((x1, x2), (y1, y2)) = fP(x1, y1)gQ(x2, y2). We will use this notation and the following consequence of Proposition 2.4 in our study in Section 5. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 211 Corollary 2.5. If P and Q are locally finite posets then µP × µQ = µP×Q. Next we recall how the Möbius function counts chains (or is an Euler characteristic for the order complex). For (x, y) ∈ F l2(P) let ci(x, y) = ∣∣{(a0, . . . , ai) ∈ F li+1 : ∀k, ak < ak+1 and a0 = x and ai = y}∣∣ be the number of chains of length i between x and y. Theorem 2.6 (Phillip Hall’s Theorem [13]; Proposition 3.8.5 [28]). If P is a locally finite poset and (x, y) ∈ F l2(P) then µ(x, y) = ∑ i (−1)ici(x, y). Now we review Rota’s Crosscut Theorem. Let L be a finite lattice with 0̂ the minimum element and 1̂ the maximum element. Usually, Rota’s Crosscut Theorem is stated globally in the lattice giving a formula for µ(0̂, 1̂). However, for our generalization we will need a local version. Definition 2.7. Let (x, y) ∈ F l2(L). A lower crosscut of the interval [x, y] = {a ∈ L|x ≤ a ≤ y} is a set Sx,y ⊆ [x, y]\{x} such that if b ∈ [x, y]\(Sx,y ∪ {x}) then there is some a ∈ Sx,y with a < b. A upper crosscut of the interval [x, y] is a set Tx,y ⊆ [x, y]\{y} such that if a ∈ [x, y]\(Tx,y ∪ {y}) then there is some b ∈ Tx,y with a < b. This definition gives Rota’s famous Crosscut Theorem which we state in the style of Lemma 2.35 in [22] for use in arrangement theory. Theorem 2.8 ([26, Theorem 3]). If L is a lattice, (x, y) ∈ F l2(L), and Sx,y is a lower crosscut of [x, y] then µ(x, y) = ∑ A⊆Sx,y∨ A=y (−1)|A|. Dually, if Tx,y is an upper crosscut of [x, y] then µ(x, y) = ∑ B⊆Tx,y∧ B=x (−1)|B|. Next we consider Weisner’s Theorem (see [31]). Theorem 2.9 ([28, Weisner’s Theorem, Corollary 3.9.3]). If L is a finite lattice with at least two elements and 1̂ ̸= a ∈ L then∑ x∈L x∧a=0̂ µ(x, 1̂) = 0. Now we recall one more result that follows from this classical result for matroids: the Mobius function of the lattice of flats of a matroid alternates in sign. Lemma 2.10. If L is a finite semimodular lattice then sgn(µ(x, y)) = (−1)rk(x)+rk(y). 212 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 3 A 3-variable incidence non-associative near-ring In this section, we define the algebraic structures where our invariants live. It turns out that these algebraic structures support various operations that can yield nice formulas. Later these formulas will be used to show certain formulas and relations on our new invariants. Definition 3.1. Let R be a commutative ring and P be a locally finite poset. Define the 3-variable incidence left near-ring as J(P, R) = Hom(F l3(P),R) with binary operations as follows: • For f, g ∈ J(P, R) we define addition by (f + g)(x, y, z) = f(x, y, z) + g(x, y, z). • For f, g ∈ J (P, R) we define a multiplication by (f  g)(x, y, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)g(a, y, b)f(b, b, z) where the juxtaposition in each term is multiplication in the ring R and (a, b) ⊴ (x, y, z) means x ≤ a ≤ y ≤ b ≤ z in P . First we show that J(P, R) is indeed left distributive. Proposition 3.2. If P is any poset then the multiplication  in J(P, R) is left distributive. Proof. Let f, g, h ∈ J(P, R) and (x, y, z) ∈ F l3(P). Then (f  (g + h))(x, y, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)(g + h)(a, y, b)f(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)(g(a, y, b) + h(a, y, b))f(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a, a)g(a, y, b)f(b, b, z) + ∑ (a,b)⊴(x,y,z) f(x, a, a)h(a, y, b)f(b, b, z) = (f  g)(x, y, z) + (f  h)(x, y, z). Remark 3.3. With this + the set J(P, R) is an abelian group. It would be convenient if J(P, R) were naturally an R-algebra. However, this is far from the case as we will see. Even the natural action of R on J(P, R) is flawed. Let r ∈ R and f, g ∈ J(P, R) then r · (f  g) = f  (r · g) but (r · f)  g = r2 · (f  g). Fortunately, though, there are a few special functions in J(P, R) that provide substantial information. We will use these to study the structure of J(P, R) and define other special elements later. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 213 Definition 3.4. Assume that 1 is the multiplicative identity and 0 is the additive identity in R. • Define δ3 ∈ J(P, R) by δ3(x, y, z) = { 1 if x = y = z 0 otherwise • Define ζ3 ∈ J(P, R) by setting ζ3(x, y, z) = 1 for all (x, y, z) ∈ F l3(P). With these functions we can investigate basic properties of J(P, R). Proposition 3.5. The element δ3 ∈ J(P, R) is a left multiplicative identity. Proof. Let f ∈ J(P, R) and (x, y, z) ∈ F l3(P). Then (δ3  f)(x, y, z) = ∑ (a,b)⊴(x,y,z) δ3(x, a, a)f(a, y, b)δ3(b, b, z) = δ3(x, x, x)f(x, y, z)δ3(z, z, z) = f(x, y, z). In the next three propositions we note that in general J(P, R) is not commutative, as- sociative, or right distributive. We could do this with a single example, however these propositions show that J(P, R) is basically never commutative, associative, or right dis- tributive. Proposition 3.6. If P is a non-trivial poset (it has at least two comparable elements) or the base ring is not Boolean (not idempotent), then the multiplication  in J(P, R) is non- commutative and δ3 is not a right multiplicative identity. Proof. Let (x, y, z) ∈ F l3(P) and suppose that either x < y or that y < z in P or that R is not Boolean. Under these assumptions we can construct a function f ∈ J(P, R) that has f(x, y, z) ̸= f(x, y, y)f(y, y, z). Then from Proposition 3.5 we have (δ3  f)(x, y, z) = f(x, y, z) but (f  δ3)(x, y, z) = f(x, y, y)f(y, y, z). The proof for the next fact is very similar. Proposition 3.7. If P is a poset with three elements x, y, z satisfying x < y < z or the base ring is not Boolean (not idempotent), then the multiplication  in J(P, R) is non- associative. Proof. Let (x, y, z) ∈ F l3(P) be three elements satisfying x < y < z in P or that R is not Boolean. Under these assumptions we can construct a function f ∈ J(P, R) that has f(x, y, y)f(y, y, y)2f(y, y, z) ̸= f(x, y, y)f(y, y, z). Compute ((f  δ3)  δ3))(x, y, z) = ∑ (a,b)⊴(x,y,z) (f  δ3)(x, a, a)δ3(a, y, b)(f  δ3)(b, b, z) = [(f  δ3)(x, y, y)][(f  δ3)(y, y, z)] = [f(x, y, y)f(y, y, y)][f(y, y, y)f(y, y, z)]. Then from Proposition 3.5 we have (f  (δ3  δ3))(x, y, z) = (f  δ3)(x, y, z) = f(x, y, y)f(y, y, z) which is different from ((f  δ3)  δ3))(x, y, z) by our assumption on f . 214 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 Proposition 3.8. If P is a non-trivial poset (it has at least two comparable elements) and R is any non-trivial commutative ring, then the multiplication  in J(P, R) is not right distributive. Proof. Let (x, y, z) ∈ F l3(P) and f ∈ J(P, R) be any function such that f(x, y, y) + f(y, y, z) ̸= 0. Then ((f + ζ3)  δ3)(x, y, z) = ∑ (a,b)⊴(x,y,z) (f + ζ3)(x, a, a)δ3(a, y, b)(f + ζ3)(b, b, z) = [(f + ζ3)(x, y, y)][(f + ζ3)(y, y, z)] = f(x, y, y)f(y, y, z) + f(x, y, y) + f(y, y, z) + 1. On the other hand we have ((f  δ3) + (ζ3  δ3))(x, y, z) = f(x, y, y)f(y, y, z) + ζ3(x, y, y)ζ3(y, y, z) = f(x, y, y)f(y, y, z) + 1 which by the hypothesis on f we have the right distributive property not holding. With Propositions 3.5, 3.6, 3.7, 3.2, and 3.8 we conclude that J(P, R) is a left only unital, non-commutative, non-associative, near-ring (see [25] for this terminology). Also, note that there is the zero function Z ∈ J(P, R) which satisfies Z  f = f  Z = Z for all f ∈ J(P, R). Further note that addition in J(P, R) is abelian. Hence J(P, R) is an abelian, zero-symmetric, left only unital, non-commutative, non-associative, near-ring. It is worth noting that in general J(P, R) is not even close to being associative on both sides and is not an alternative algebra or any similar generalization. Now we look at a few special cases that do not satisfy the hypothesis of some of these propositions. Example 3.9. Let P = B0 = {0} be the poset with just one element and R any commu- tative ring. Then as a set J(B0, R) = R, but multiplication is given by a  b = aba = a2b. If R is Boolean then J(B0, R) ∼= R. Otherwise, this near-ring is not associative, not com- mutative, and is only left unital. Example 3.10. Let P = B1 = {0, 1} be the Boolean poset of rank 1 and R be any Boolean ring (one example would be F2). Then the hypothesis of Proposition 3.7 is not satisfied and the non-equality f(x, y, y)f(y, y, y)2f(y, y, z) ̸= f(x, y, y)f(y, y, z) used in the proof is always equal. It turns out that in this case J(B1, R) is associative and we prove this now. In order to shorten the calculation we will denote (0, 0, 0) by 0⃗ and (1, 1, 1) by 1⃗. First we see that ((f  g)  h)(⃗0) = f (⃗0)g(⃗0)h(⃗0) = (f  (g  h))(⃗0). Then for the non-trivial tuple (0, 0, 1) we compute ((f  g)  h)(0, 0, 1) =(f  g)(⃗0)h(⃗0)(f  g)(0, 0, 1) + (f  g)(⃗0)h(0, 0, 1)(f  g)(⃗1) =f (⃗0)g(⃗0)h(⃗0)[f (⃗0)g(⃗0)f(0, 0, 1) + f (⃗0)g(0, 0, 1)f (⃗1)] + f (⃗0)g(⃗0)h(0, 0, 1)f (⃗1)g(⃗1) =f (⃗0)g(⃗0)h(⃗0)f(0, 0, 1) + f (⃗0)g(⃗0)h(⃗0)g(0, 0, 1)f (⃗1) + f (⃗0)g(⃗0)h(0, 0, 1)f (⃗1)g(⃗1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 215 Then the other side of the associative identity is (f  (g  h))(0, 0, 1) =f (⃗0)(g  h)(⃗0)f(0, 0, 1) + f (⃗0) [ (g  h)(0, 0, 1) ] f (⃗1) =f (⃗0)g(⃗0)h(⃗0)f(0, 0, 1) + f (⃗0) [ g(⃗0)h(⃗0)g(0, 0, 1) + g(⃗0)h(0, 0, 1)g(⃗1) ] f (⃗1) =((f  g)  h)(0, 0, 1). Hence J(B1, R) is associative. This example does satisfy the hypothesis of Proposition 3.8. Hence J(B1, R) is a (associative) left abelian (addition is commutative) near-ring. That’s about as good as it gets though. For example, if R = F2 then J(B1,F2) is not a near-field because any function with f (⃗0) = 0 and f(0, 0, 1) = 1 does not have an inverse. For exactly the same reason δ3 ∈ J(B1,F2) is still not a right identity element. 4 Operations on incidence functions In this section we look at a relationship between the classical incidence algebra I(P, R) and J(P, R). For f, g ∈ I(P, R) we define f♢g ∈ J(P, R) by setting (f♢g)(x, y, z) = f(x, y)g(y, z). We can use the ♢ operation to construct interesting elements in J(P, R). There are rela- tionships between the operations ∗ in I(P, R),  in J(P, R), and ♢. Proposition 4.1. If f, g, r, s ∈ I(P, R) and f(b, b)g(a, a) = 1 for all a, b ∈ P then (f♢g)  (r♢s) = (f ∗ r)♢(s ∗ g). Proof. Let (x, y, z) ∈ F l3(P) and f, g, r, s ∈ I(P, R). Then ((f♢g)  (r♢s))(x, y, z) = ∑ (a,b)⊴(x,y,z) (f♢g)(x, a, a)(r♢s)(a, y, b)(f♢g)(b, b, z) = ∑ (a,b)⊴(x,y,z) f(x, a)g(a, a)r(a, y)s(y, b)f(b, b)g(b, z) =  ∑ x≤a≤y f(x, a)r(a, y)  ∑ y≤b≤z s(y, b)g(b, z)  = [(f ∗ r)(x, y)] [(s ∗ g)(y, z)] =((f ∗ r)♢(s ∗ g))(x, y, z) where the third equality only holds due the the assumption. One can see from the proof that without the hypothesis on f and g that the equality will not hold. Hence there is no hope for this to give any kind of near-ring homomorphism from a twisted product version of I(P, R)× I(P, R). Also, the natural addition homomorphism assumption does not hold. Instead we have the following proposition which does not have special hypothesis on the functions. For this proposition there are two different additions, for I(P, R) and J(P, R), which for brevity we use the same addition symbol. 216 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 Proposition 4.2. If f, g, r, s ∈ I(P, R) then (f + g)♢(r + s) = (f♢r) + (f♢s) + (g♢r) + (g♢s). Proof. For all (x, y, z) ∈ F l3(P) ((f + g)♢(r + s))(x, y, z) =(f(x, y) + g(x, y))(r(y, z) + s(y, z)) =f(x, y)r(y, z) + f(x, y)s(y, z) + g(x, y)r(y, z)+ g(x, y)s(y, z) =((f♢r) + (f♢s) + (g♢r) + (g♢s))(x, y, z) which is the identity we are looking for. We can also define products of functions on products of posets over 3-flags. We prefer to limit our study of J(P, R) to this product definition since the technicalities of tensor products over non-associative near-rings would present significant and unnecessary com- plications. Definition 4.3. Let P and Q be locally finite posets, fP ∈ J(P, R), and gQ ∈ J(Q, R). Define fP × gQ ∈ J(P ×Q, R) by (fP × gQ)((x1, x2), (y1, y2), (z1, z2)) = fP(x1, y1, z1)gQ(x2, y2, z2). Now we show how the ♢ operation is compatible with products of posets. Proposition 4.4. If P and Q are locally finite posets, fP , gP ∈ I(P, R), and rQ, sQ ∈ I(Q, R) then (fP♢gP)× (rQ♢sQ) = (fP × rQ)♢(gP × sQ). Proof. Let ((x1, x2), (y1, y2), (z1, z2)) ∈ F l3(P ×Q). Then ((f♢g)× (r♢s))((x1, x2), (y1, y2), (z1, z2)) = [(f♢g)(x1, y1, z1)] [(r♢s)(x2, y2, z2)] = [f(x1, y1)g(y1, z1)] [r(x2, y2)s(y2, z2)] = [f(x1, y1)r(x2, y2)] [g(y1, z1)s(y2, z2)] = [(f × r)((x1, y1), (x2, y2))] [(g × s)((y1, z1), (y2, z2))] = ((f × r)♢(g × s))((x1, x2), (y1, y2), (z1, z2)) which completes the proof. As in Proposition 4.4 we will now show how the operations × and  factor over prod- ucts of posets. We use subscripts on these operations to keep track of which poset the operation is applied. Proposition 4.5. If P and Q be locally finite posets, fP , gP ∈ J(P, R), and rQ, sQ ∈ J(Q, R) then (fP P gP)× (rQ Q sQ) = (fP × rQ) P×Q (gP × sQ). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 217 Proof. Let x = (x1, x2), y = (y1, y2), z = (z1, z2) ∈ P ×Q so that (x, y, z) ∈ F l3(P × Q) and a = (a1, a2), b = (b1, b2) ∈ P ×Q so that (a, b) ∈ F l2(P ×Q). Then ((fP×rQ) P×Q (gP × sQ))(x, y, z) = ∑ (a,b)⊴(x,y,z) (fP × rQ)(x, a, a)(gP × sQ)(a, y, b)(fP × rQ)(b, b, z) = ∑ (a1,b1) ∑ (a2,b2) fP(x1, a1, a1)gP(a1, y1, b1)fP(b1, b1, z1) rQ(x2, a2, a2)sQ(a2, y2, b2)rQ(b2, b2, z2) = [(fP  gP)(x1, y1, z1)] [(rQ  sQ)(x2, y2, z2)] = ((fP P gP)× (rQ Q sQ))(x, y, z) which is the required identity. 5 The J-function Let P be a locally finite poset. In this section we define the central invariant of this note which we call the J function. This function is a generalization of the classical Möbius function µ. We show that it satisfies generalizations of the classical theorems on µ. A key ingredient for these results is the operation ♢. Definition 5.1. Define J : F l3(P) → Z for all fixed (x, y, z) ∈ F l3(P ) by∑ (a,b)⊴(x,y,z) J(a, y, b) = δ3(x, y, z). This function is well defined because either x = y = z with J(x, y, z) = 1 or otherwise all of the following summations are finite J(x, y, z) =− ∑ x 2. The lattice Pn consists of 0̂, 1̂, and n atoms α1, . . . , αn. Now JPn(0̂, 0̂, 1̂) = n − 1 and JPn(0̂, 1̂, 1̂) = n − 1 are the only JPn values that do not have αn as an entry and incorporate αn in it’s recursive definition. So, JPn(0̂, 0̂, 1̂) = JPn−1(0̂, 0̂, 1̂) + 1 and similarly for (0̂, 1̂, 1̂). Incorporating this difference into the calculation we get that M(Pn, t) =M(Pn−1, t) + t4 + t2 + J(0̂, 0̂, αn)t5 + J(0̂, αn, αn)t4 + J(0̂, αn, 1̂)t3 + J(αn, αn, αn)t 3 + J(αn, αn, 1̂)t 2 + J(αn, 1̂, 1̂)t =(t2 − (n− 1)t+ 1)(t+ 1)2(t− 1)2 − (t5 − 2t3 + t) =(t2 − nt+ 1)(t+ 1)2(t− 1)2 which is the desired formula. Now we consider a decomposition of M(L, t) for a finite lattice L. If L is a finite lattice then Lop is the same underlying set as L but with the order reversed (i.e. x ≤op y in Lop if and only if x ≥ y in L). Also for y ∈ L let Ly = {x ∈ L|x ≤ y} and Ly = {x ∈ L|x ≥ y}. Now we can state the result. Proposition 6.10. If L is a finite ranked lattice then M(L, t) = trk(L) ∑ y∈L tcrk(y)χ(Ly, t)χ((Lop)y, t−1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 223 Proof. First we note that for x ≤ y ∈ L the Möbius function on Lop has µop(y, x) = µ(x, y) and that rank is corank in Lop. Then again using Theorem 5.2 we compute M(L, t) = ∑ (x,y,z)∈Fl3(P) J(x, y, z)tρ(x,y,z) = ∑ y∈L ∑ x≤y ∑ z≥y µ(x, y)µ(y, z)tcrk(x)+crk(y)+crk(z) = ∑ y∈L tcrk(y) ∑ x≤y µ(x, y)tcrk(x) ∑ z≥y µ(y, z)tcrk(z) = ∑ y∈L tcrk(y)χ(Ly, t) ∑ x≤y µ(x, y)trk(L)−rk(x) = ∑ y∈L tcrk(y)χ(Ly, t)trk(L) ∑ x≥opy µop(y, x)t−rk(x) =trk(L) ∑ y∈L tcrk(y)χ(Ly, t)χ((Lop)y, t−1). We can use Proposition 6.10 to compute M(P, t) for cases where χ(P, t) is well known. Let Lnq be the modular lattice of all subspaces in Fnq , a vector space of dimen- sion n over a field with q elements. The Möbius function and the characteristic polynomial of Lnq are well known. Proposition 6.11 ([34, Proposition 7.5.3]). In Lnq we have µ(0̂, 1̂) = (−1)nq( n 2) and χ(Lnq , t) = n−1∏ i=0 (t− qi). Using this we can get a nice formulation for M(Lnq , t). First we need to recall some terminology from q-series. Let[ n k ] q = (qn − 1) · · · (q − 1) (qk − 1) · · · (q − 1) · (qn−k − 1) · · · (q − 1) be the q-binomial coefficient (aka Gaussian coefficient). Also, we denote by[ n k1, k2, . . . , km ] q = [ n k1 ] q [ n− k1 k2 ] q · · · [ n− (k1 + · · · km−1) km ] q the q-multinomial coefficient. We also use the q-Pochhammer symbol (a; q)n = n−1∏ i=0 (1− aqi). 224 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 We use [3] for a general reference for q-series. Using Proposition 6.11 we get the following. Proposition 6.12. If Lnq is the modular lattice of subspaces of Fnq then M(Lnq , t) = ∑ 0≤i≤j≤k≤n (−1)k−i [ n i, j − i, k − j, n− k ] q q( j−i 2 )+( k−j 2 )t3n−i−j−k. Proof. Use that [ n k ] q counts the number of subspaces of dimension k in Fnq and apply Theorem 5.2 to J in M(Lnq , t) together with Proposition 6.11. Now we can reformulate Proposition 6.12 using Proposition 6.10 together with Propo- sition 6.11 to get a nice identity in q-series. Proposition 6.13. If Lnq is the modular lattice of subspaces of Fnq , then M(Lnq , t) = tn ∑ 0≤k≤n tn−k [ n k ] q n−k−1∏ i=0 (t− qi) k−1∏ j=0 (t− qj). It turns out that −1 is a root of M(Lnq , t). We need a few results in order to prove this. First we present a formula or q-identity which seems to be a kind of q-generalized binomial theorem (the authors could not find it in the literature). It’s interesting that in the odd case the sum trivially collapses but not for the even case. Lemma 6.14. If n > 0 then n∑ k=0 (−1)k [ n k ] q (−1 : q)n−k(−1; q)k = 0. Proof. Let S(n) = n−1∑ k=0 (−1)k [ n k ] q (−1; q)n−k(−1; q)k (−1; q)n which is the left hand side up to the n − 1 term divided by the nth term. Using tech- niques from [24] and Mathematica [15] we build a recursion for S(n). We compute J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 225 (1 + qn−1)S(n) = n−1∑ k=0 (−1)k ( qk [ n− 1 k ] q + [ n− 1 k − 1 ] q ) (−1; q)n−k(−1; q)k (−1; q)n−1 = n−1∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 (qk + qn−1) + n−1∑ k=1 (−1)k [ n− 1 k − 1 ] q (−1; q)n−k(−1; q)k (−1; q)n−1 =(−1)n−12qn−1 + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qk + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qn−1 + n−2∑ k=0 (−1)k+1 [ n− 1 k ] q (−1; q)n−(k+1)(−1; q)k+1 (−1; q)n−1 =(−1)n−12qn−1 + n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 qk + qn−1S(n− 1)− n−2∑ k=0 (−1)k [ n− 1 k ] q (−1; q)n−k−1(−1; q)k (−1; q)n−1 (1 + qk) =(−1)n−12qn−1 + qn−1S(n− 1)− S(n− 1). Now we prove with induction that S(n) = (−1)n−1. First we see that S(1) = 1. Then using the recursion above we have (1 + qn−1)S(n) = (−1)n−12qn−1 − qn−1(−1)n−1 + (−1)n−1 = (−1)n−1(qn−1 + 1) which finishes the proof. Proposition 6.15. If Lnq is the modular lattice of subspaces of Fnq then M(Lnq ,−1) = 0. Proof. Evaluate the expression in Proposition 6.13 and apply Lemma 6.14. Now we can prove the main result of this section. Theorem 6.16. If L is a modular geometric lattice (modular matroid) then M(L,−1) = 0. Proof. Use the classical result that a modular geometric lattice is product of Boolean and projective spaces (see 12.1 Theorem 4 in [32] or Proposition 6.9.1 in [23]). Then the result follows from Propositions 6.15, 6.8, and 6.6. Remark 6.17. The proof of Theorem 6.16 is done in cases. It would be interesting if there was a case free proof just using the modular property. 226 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 Remark 6.18. At first when looking at examples of M on the lattice of flats L(M) of a matroid M it seems that the converse of Theorem 6.16 might be true. As for even the simplest non-modular matroid U3,4 has M polynomial M(L(U3,4), t) = (t− 1)(t8 − 3t7 − t6 + 12t5 − 2t4 − 12t3 + 3t2 + 5t− 1) which does not have a factor of (t + 1). However, the converse is false, but the example seems rather special. Using the SageMath computer algebra system [10] we compute M(L(M∗(K3,3)), t) = (t10 − 9t9 + 22t8 + 12t7 − 81t6 + 21t5 + 69t4 − 18t3 − 34t2 + 15t− 1)(t+ 1)(t− 1) where M∗(K3,3) is the dual matroid of the graphic matroid corresponding to the complete bipartite graph K3,3. Since M∗(K3,3) is a connected non-modular matroid (it does not have a modular direct summand) this example gives a connected non-modular matroid that has −1 as a root of M. This example and Theorem 6.16 motivate a few questions. Question 6.19. Is there a rank 3 non-modular connected matroid M such that M(M,−1) = 0? Question 6.20. Is there a classification of all matroids whose M polynomial has -1 as a root? Question 6.21. Is there a nice enumerative combinatorial interpretation for M(M,−1) where M is a matroid (i.e. what does it count)? 6.1 No Deletion-Contraction We now show that J and M are not some evaluation of the Tutte polynomial for matroids. We first recall the following definition. Definition 6.22. We say that a function f from matroids to a ring R is a generalized Tutte- Grothendieck invariant (following [4] Sec 1.8.6) if there exists a, b ∈ R such that for every matroid M and element of the ground set e ∈ M f(M) =  f(M\e)f(L) if e is a loop f(M/e)f(C) if e is a coloop af(M\e) + bf(M/e) otherwise. where L is the matroid consisting of exactly one loop and C is the matroid consisting of exactly one coloop. Let Ur,n be the uniform matroid of rank r on n elements and recall that Ur,r ∼= Br are Boolean or free matroids. Then, a direct computation gives J (B1, t) = t+ 1 and J (U2,n, t) = (n− 1)t2 + nt+ n− 1. Hence J(U2,3, t) = 2t2 + 3t + 2. Then any deletion is U2,3\e ∼= B2 and any contraction is U2,3/e ∼= U1,2. Putting this together with Definition 6.22 and assuming that J is a Tutte-Grothendieck invariant 2t2 + 3t+ 2 = a(t2 + 2t+ 1) + b(t+ 1). J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 227 However, this is a contradiction since t+ 1 is not a factor of the left hand side. The same result for M needs two more steps. Looking at the same matroid and using Proposition 6.9 we get M(U2,3, t) = (t2 − 3t+ 1)(t+ 1)2(t− 1)2 = a(t+ 1)2(t− 1)4 + b(t+ 1)(t− 1)2 which reduces to b = (t+ 1)(t2 − 3t+ 1)− a(t+ 1)(t− 1)2. Then we look at U2,4 and again assume M is a Tutte-Grothendieck invariant M(U2,4, t) = (t2−4t+1)(t+1)2(t−1)2 = a(t2−3t+1)(t+1)2(t−1)2+b(t+1)(t−1)2. Inserting the above value for b and reducing we get t2 − 4t+ 1 = a(t2 − 3t+ 1) + (t2 − 3t+ 1)− a(t− 1)2 which gives a = 1 and makes b = −t(t+ 1). But then M(U3,4, t) = (t− 1)(t8 − 3t7 − t6 + 12t5 − 2t4 − 12t3 + 3t2 + 5t− 1) which does not have a factor of t+ 1. This is a contradiction since the right hand side M(U3,4\e, t)− t(t+ 1)M(U3,4/e, t) = M(U3,3, t)− t(t+ 1)M(U2,3, t) does have a t+ 1 factor. 6.2 Valuations Here we study the invariant M over matroid subdivisions. One could focus on a wider range combinatorial objects like posets but we are motived by applications to matroid the- ory. First we recall the basis matroid polytope (using [6] as our general reference for this material). A matroid M can be defined via its set of bases B(M) which are all the inde- pendent sets of M whose size is the rank of M . Then, the matroid polytope of M is P (M) = Conv{eB |B ∈ B(M)} where eB = ei1 + · · ·+ eir with B = {i1, . . . , ir}. Now we need a few key definitions to state our main result. Definition 6.23. A matroid polyhedral subdivision of a matroid polytope P (M) is a col- lection of polyhedra {Pi} such that ⋃ Pi = P (M), each Pi is a matroid polytope whose vertices are vertices of P (M), and for i ̸= j if Pi ⋂ Pj ̸= ∅, then Pi ⋂ Pj is a proper face of both Pi and Pj . Now we want to know how invariants decompose across subdivisions which gives rise to valuations. We will use what is called a weak valuation in [6] but we follow [5] and just say valuation. This makes sense since by Theorem 4.2 in [6] for matroids weak valuations are actually strong valuations. 228 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 Definition 6.24. Let P be the collection of matroid polytopes and R a commutative ring. A function f : P → R is a (weak) valuation if for any matroid polytope P (M) and any matroid polyhedral subdivision with maximal pieces {P (M1), . . . , P (Mk)} we have that f(∅) = 0 and f(P (M)) = ∑ {j1,...,ji}⊆[k] (−1)if(P (Mj1) ∩ · · · ∩ P (Mji)). Finally we can state the result for the invariant J in terms of valuations. Proposition 6.25. The polynomial J is a valuation on matroids. Proof. Using the decomposition of the J-function given in Theorem 5.2 we know that J (M, t) = (−1)rk(M) ∑ X∈L(M) µ(∅, X)µ(X, 1̂)trk(X) where 1̂ is the maximal flat of M . Hence as a function from the collection of matroids to Z[t] we can represent the function J as J = (±1) ∑ f1 ⋆ f2 where f1 ⋆ f2 = m ◦ (f1 ⊗ f2) ◦ ∆S,T from the notation in Theorem C in [6] and f1 = χM (0) and f2 = χM (0)trk(M). Since f1 and f2 are both Tutte-Grothendieck invariants for matroids and are evaluations of the Tutte polynomial we can conclude that f1 and f2 are both valuations from Proposition 7.5 in [6]. Finally putting it all together Theorem C in [6] finished the result. We conclude with a natural question. The polynomial M(L, t) is slightly more com- plicated but has promising properties that seems to imply it should be a valuation. Question 6.26. Is the polynomial M a matroid valuation? It seems that Proposition 6.10 with Proposition 7.5 and Theorem C in [6] is essentially the proof. However that would use that the characteristic polynomial on the flipped lattice of flats L(M)op is a matroid valuation. References [1] M. Aguiar and F. Ardila, Hopf monoids and generalized permutahedra, 2017, arXiv:1709.075048 [math.CO]. [2] M. Aguiar and S. Mahajan, Topics in Hyperplane Arrangements, volume 226 of Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2017, doi:10.1090/ surv/226, https://doi.org/10.1090/surv/226. [3] G. E. Andrews, q-series: Their Development and Application in Analysis, Nnumber Theory, Combinatorics, Physics, and Computer Algebra, volume 66 of CBMS Regional Conference Series in Mathematics, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1986. [4] F. Ardila, Algebraic and geometric methods in enumerative combinatorics, in: Handbook of enumerative combinatorics, CRC Press, Boca Raton, FL, Discrete Math. Appl., pp. 3–172, 2015. J. Johnson et al.: A non-associative incidence near-ring with a generalized Möbius function 229 [5] F. Ardila, A. Fink and F. Rincón, Valuations for matroid polytope subdivisions, Canad. J. Math. 62 (2010), 1228–1245, doi:10.4153/cjm-2010-064-9, https://doi.org/10. 4153/cjm-2010-064-9. [6] F. Ardila and M. Sanchez, Valuations and the Hopf monoid of generalized permutahedra, Int. Math. Res. Not. IMRN (2023), 4149–4224, doi:10.1093/imrn/rnab355, https://doi.org/ 10.1093/imrn/rnab355. [7] C. A. Athanasiadis, Characteristic polynomials of subspace arrangements and finite fields, Adv. Math. 122 (1996), 193–233, doi:10.1006/aima.1996.0059, https://doi.org/10.1006/ aima.1996.0059. [8] J. Bastidas, Species and hyperplane arrangements, Ph.D. thesis, Cornell University, 2021. [9] A. Cameron and A. Fink, The Tutte polynomial via lattice point counting, J. Comb. Theory Ser. A 188 (2022), Paper No. 105584, doi:10.1016/j.jcta.2021.105584, https://doi.org/ 10.1016/j.jcta.2021.105584. [10] T. S. Developers, Sage Mathematics Software (Version 8.1), 2020, http://www. sagemath.org. [11] C. Dupont, A. Fink and L. Moci, Universal Tutte characters via combinatorial coalgebras, Al- gebr. Comb. 1 (2018), 603–651, doi:10.5802/alco, https://doi.org/10.5802/alco. [12] B. Elias, N. Proudfoot and M. Wakefield, The Kazhdan-Lusztig polynomial of a matroid, Adv. Math. 299 (2016), 36–70, doi:10.1016/j.aim.2016.05.005, https://doi.org/10.1016/ j.aim.2016.05.005. [13] P. Hall, A Contribution to the Theory of Groups of Prime-Power Order, Proc. London Math. Soc. (2) 36 (1934), 29–95, doi:10.1112/plms/s2-36.1.29, https://doi.org/10.1112/ plms/s2-36.1.29. [14] G. H. Hardy and E. M. Wright, An Introduction to the Theory of Numbers, Oxford University Press, Oxford, sixth edition, 2008. [15] W. R. Inc., Mathematica, Version 12.3.1, Champaign, IL, 2021, https://www.wolfram. com/mathematica. [16] S. A. Joni and G.-C. Rota, Coalgebras and bialgebras in combinatorics, Stud. Appl. Math. 61 (1979), 93–139, doi:10.1002/sapm197961293, https://doi.org/10.1002/ sapm197961293. [17] R. Jurrius, Relations between Möbius and coboundary polynomials, Math. Comput. Sci. 6 (2012), 109–120, doi:10.1007/s11786-012-0117-6, https://doi.org/10.1007/ s11786-012-0117-6. [18] T. Krajewski, I. Moffatt and A. Tanasa, Hopf algebras and Tutte polynomials, Adv. in Appl. Math. 95 (2018), 271–330, doi:10.1016/j.aam.2017.12.001, https://doi.org/10. 1016/j.aam.2017.12.001. [19] J. L. Martin, Lecture notes on algebraic combinatorics, 2012 . [20] A. F. Möbius, Über eine besondere Art von Umkehrung der Reihen, J. Reine Angew. Math. 9 (1832), 105–123, doi:10.1515/crll.1832.9.105, https://doi.org/10.1515/crll. 1832.9.105. [21] W. Murray, Möbius polynomials, Math. Mag. 85 (2012), 376–383, doi:10.4169/math.mag.85. 5.376, https://doi.org/10.4169/math.mag.85.5.376. [22] P. Orlik and H. Terao, Arrangements of Hyperplanes, volume 300 of Grundlehren der Math- ematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Springer- Verlag, Berlin, 1992, doi:10.1007/978-3-662-02772-1, https://doi.org/10.1007/ 978-3-662-02772-1. 230 Ars Math. Contemp. 24 (2024) #P2.03 / 207–230 [23] J. Oxley, Matroid Theory, volume 21 of Oxford Graduate Texts in Mathematics, Oxford Uni- versity Press, Oxford, 2nd edition, 2011, doi:10.1093/acprof:oso/9780198566946.001.0001, https://doi.org/10.1093/acprof:oso/9780198566946.001.0001. [24] P. Paule and A. Riese, A Mathematica q-analogue of Zeilberger’s algorithm based on an alge- braically motivated approach to q-hypergeometric telescoping, in: Special functions, q-series and related topics (Toronto, ON, 1995), Amer. Math. Soc., Providence, RI, volume 14 of Fields Inst. Commun., pp. 179–210, 1997. [25] G. Pilz, Near-Rings, volume 23 of North-Holland Mathematics Studies, North-Holland Pub- lishing Co., Amsterdam, 2nd edition, 1983, the theory and its applications. [26] G.-C. Rota, On the foundations of combinatorial theory. I. Theory of Möbius functions, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 2 (1964), 340–368 (1964), doi:10.1007/ bf00531932, https://doi.org/10.1007/bf00531932. [27] E. Spiegel and C. J. O’Donnell, Incidence Algebras, volume 206 of Monographs and Textbooks in Pure and Applied Mathematics, Marcel Dekker, Inc., New York, 1997. [28] R. P. Stanley, Enumerative Combinatorics. Volume 1, volume 49 of Cambridge Studies in Ad- vanced Mathematics, Cambridge University Press, Cambridge, 2nd edition, 2012. [29] H. Terao, Free arrangements of hyperplanes and unitary reflection groups, Proc. Japan Acad. Ser. A Math. Sci. 56 (1980), 389–392, http://projecteuclid.org/euclid.pja/ 1195516722. [30] M. Wakefield, Partial flag incidence algebras, 2022, arXiv:1605.01685 [math.CO]. [31] L. Weisner, Abstract theory of inversion of finite series, Trans. Am. Math. Soc. 38 (1935), 474–484, doi:10.2307/1989808, https://doi.org/10.2307/1989808. [32] D. J. A. Welsh, Matroid Theory, Academic Press [Harcourt Brace Jovanovich, Publishers], London-New York, 1976, l. M. S. Monographs, No. 8. [33] T. Zaslavsky, Facing up to arrangements: face-count formulas for partitions of space by hyper- planes, Mem. Am. Math. Soc. 1 (1975), vii+102. [34] T. Zaslavsky, The Möbius function and the characteristic polynomial, in: Combinatorial Ge- ometries, Cambridge Univ. Press, Cambridge, volume 29 of Encyclopedia Math. Appl., pp. 114–138, 1987. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.04 / 231–271 https://doi.org/10.26493/1855-3974.2968.d23 (Also available at http://amc-journal.eu) Valuations and orderings on the real Weyl algebra Lara Vukšić * Institute of Mathematics, Physics, and Mechanics, Ljubljana, Slovenia and Faculty of Mathematics and Physics, University of Ljubljana, Slovenia Received 19 September 2022, accepted 12 February 2023, published online 20 September 2023 Abstract The first Weyl algebra A1(k) over a field k is the k-algebra with two generators x, y subject to [y, x] = 1 and was first introduced during the development of quantum mechan- ics. In this article, we classify all valuations on the real Weyl algebra A1(R) whose residue field is R. We then use a noncommutative version of the Baer-Krull theorem to classify all orderings on A1(R). As a byproduct of our studies, we settle two open problems in real algebraic geometry. First, we show that not all orderings on A1(R) extend to an ordering on a larger ring R[y; δ], where R is the ring of Puiseux series, introduced by Marshall and Zhang in 2000, and characterize the orderings that do have such an extension. Second, we show that for valuations on noncommutative division rings, Kaplansky’s theorem that extensions by limits of pseudo-Cauchy sequences are immediate fails in general. Keywords: Weyl algebra, noncommutative valuations, skew polynomial rings, orderings, extensions of valuations, extensions of orderings. Math. Subj. Class. (2020): 16W60, 06F25, 13J30, 14A22, 16S3 1 Introduction Valuation theory was first developed for commutative fields in the context of number theory and was first defined by József Kürschák [12] in 1913. For modern treatments, we refer to the books of Engler and Prestel [5] or Kuhlmann [10]. Oscar Schilling wrote the first major work on valuations on (noncommutative) division rings in 1945 [21]. A valuation on a division ring D is a map v : D → Γ ∪ {∞}, where Γ is an ordered group written additively and ∞ ̸∈ Γ,∞ > γ for each γ ∈ Γ, with the following properties: *I would like to thank my advisor Igor Klep for his guidance and many helpful comments and suggestions. E-mail address: lara.vuksic@fmf.uni-lj.si (Lara Vukšić) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 232 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 1. ∀x ∈ D : v(x) = ∞ ⇔ x = 0, 2. ∀x, y ∈ D : v(xy) = v(x) + v(y), 3. ∀x, y ∈ D : v(x+ y) ≥ min{v(x), v(y)}. It follows that v is a homomorphism from D∗ to Γ. The set Ov := {x ∈ D | v(x) ≥ 0} is called the valuation ring associated to v, and Mv := {x ∈ M | v(x) > 0} is its maximal ideal. The division ring D := Ov/Mv is called the associated residue division ring. Since v is a group homomorphism, the subgroup O∗v is normal in D∗. Several alternative ap- proaches to noncommutative valuations, where v does not define a group homomorphism, were introduced and studied recently by Nicolai Ivanovich Dubrovin in [7] and [4] (see also [17] for a more thorough treatment), and by Jean-Pierre Tignol and Adrien Wadsworth in [23]. Suppose F is a field. Then all valuations on the field or rational functions F (x) with residue field F are well-known, namely, the p-adic valuations for irreducible polynomials p(x) ∈ F [x], and the vdeg valuation, defined by vdeg( p q ) := deg(q)− deg(p). The description of all valuations on the field of rational functions in several variables with residue field equal to the base field is much more involved. There are many descriptions of constructions of such valuations in the literature. Among famous examples of such descriptions are the one given by Saunders MacLane in [13] and the one given by Franz- Viktor Kuhlmann in [11]. As valuations on Ore extensions uniquely extend to their quotient division ring, the description of all valuations on Ore division rings is equivalent to the description of all valuations on corresponding quotient division rings. The description of all valuations on noncommutative Ore extensions R[x;σ, δ] where R is a domain, σ : R → R is a ring homomorphism and δ : R → R a σ-derivation is even more complex than in the commu- tative case. Additional difficulties arise from the fact that [f, g] = 0 does not hold for all f, g ∈ R[x;σ, δ]. Granja, Martı́nez, and Rodrı́guez have shown in [6] that the set of all real valuations extending to the skew polynomial ring has the structure of a parameterized complete non-metric tree. Further recent progress on valuations on Ore extensions is given by Onay in [18] and Rohwer in his PhD thesis [20]. 1.1 Results Our main goal is to classify all orderings and real valuations on the real Weyl algebra A1(R) or, equivalently, its quotient division ring D1(R). The Weyl algebra is the noncommutative algebra generated by two elements x, y satisfying [y, x] = 1. Hence its elements are all of the form ∑ i,j αi,jx iyj , αi,j ∈ R. Because of this, our approach to constructing valuations on A1(R) is inspired by classical constructions of valuations on commutative rational functions in two unknowns mentioned above. However, the relation [y, x] = 1 gives rise to additional constraints and far fewer valuations than in the commutative case. L. Vukšić: Valuations and orderings on the real Weyl algebra 233 As we will show, the valuations on A1(R) we are interested in all satisfy v[a, b] > v(ab) for all nonzero a, b. We call such valuations strongly abelian. They have an abelian value group and commutative residue field. In Section 2 we give some properties of strongly abelian valuations. We show that if a valuation v on a division ring D satisfies D = Z(D) and the value group is of rational rank one, then v is strongly abelian. Under additional constraints on the residue field and the value group we extend this statement to valuations of higher rational rank. In Section 3 we give a characterization of all valuations v on the real Weyl algebra A1(R) with residue field R in the spirit of MacLane [13]. The construction is inspired by the outline given by Shtipelman in [22] for valuations on the complex Weyl algebra A1(C). We also explicitly describe the associated value groups and show that they are all isomorphic to subgroups of Q or Q× Z. In their attempt to describe all orderings on A1(R) in [16], Murray Marshall and Yufei Zhang introduced the Ore extensions R[y; δ] and R̃[y; δ], with R := { ∑ k≥m akx − kn | ak ∈ R,m ∈ Z, n ∈ N}, R̃ := { ∑ q∈A aqx −q | A ⊂ Q is well-ordered} and δ(p(x)) = p′(x). As is often done in real algebraic geometry, all orderings are de- scribed by classifying all real valuations via the Baer-Krull theorem. Marshall and Zhang described almost all valuations v on R[y; δ] with residue field R; in one case, they did not prove that v is a valuation. In Section 4, we complete their characterization. Marshall and Zhang also conjectured that all valuations on A1(R) with residue field R extend to a valuation on R[y; δ] with the same residue field. We refute their conjecture in Section 4. Further, we combine our classification of valuations on A1(R) with Marshall and Zhang’s description of valuations on R[y; δ] to characterize the valuations on A1(R) with residue field R that extend to a valuation R[y, δ] with the same residue field. All such extensions are again strongly abelian. In Section 5, we show that all valuations on R[y; δ] with residue field R uniquely extend to a strongly abelian valuation on R̃[y; δ] with the same residue field. We also show that the value group of such an extension is not of rational rank one. As a byproduct of our investigations, we show that Kaplansky’s theorem that all ex- tensions by limits of pseudo-Cauchy sequences are immediate (in particular, they do not change the rational rank of the value group) fails for noncommutative division rings. As Marshall and Zhang observe in [15], all strongly abelian valuations v on a division ring D with a formally real residue field are compatible with an order on D. In Section 6, we describe all v-compatible orders on A1(R) for every valuation v on A1(R) constructed in Section 3 using a noncommutative version of the Baer-Krull theorem as given in [2] (see also [1, 3, 24] and [9] for modern treatments and extensions). We also characterize the v-compatible orders on A1(R) that extend to an order on R[y; δ] compatible with v’s extension to R[y; δ]. 2 Strongly abelian valuations We present some properties of valuations on noncommutative division rings which we will use later to describe order-compatible valuations on the real Weyl algebra A1(R) and some of its ring extensions. First, we define a property of valuations on division rings. 234 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 Definition 2.1. Suppose v is a valuation on a division ring D. We say v is strongly abelian if v[a, b] > v(ab) holds for all nonzero a, b ∈ D. Any valuation on a field is strongly abelian. In this section, we describe a sufficient condition for a valuation v to be strongly abelian. This property will be important for us for two reasons. Firstly, it is obvious that if a valuation v on a division ring D is strongly abelian, then the associated value group and residue division ring are commutative. Sec- ondly, we are particularly interested in order-compatible valuations on A1(R); minimal such have residue field R, as it was shown in [16]. It follows from Theorem 2.5 of [15] that a strongly abelian valuation v on a division ring D with a formally real residue field is compatible with an order on D by the noncommutative version of the Baer-Krull theorem as given in [24]. Proposition 2.2. Let v be a valuation on a division ring D such that D = Z(D). Let a, b ∈ D∗ be such that v(a) and v(b) are rationally dependent. Then v[a, b] > v(ab). Proof. Since v(a) and v(b) are rationally dependent, v(ab) = v(ba) ≤ v[a, b]. Suppose v[a, b] = v(ab). In particular, a and b do not commute. Let β := aba−1b−1 ∈ D. We have β = aba−1b−1 = ([a, b] + ba)a−1b−1 = [a, b]a−1b−1 + 1 ̸= 1. Let v(b) = − ℓkv(a) for ℓ, k ∈ Z, ℓ and k coprime. It follows that v[a, b] = k−ℓ k v(a). Define γ := (ba)kaℓ−k ∈ D. Let β′, γ′ ∈ Z(D) be such that v(β′) = v(γ′) = 0, β′ = β and γ′ = γ. Then on one hand, v(a(ba)k − γ′ak−ℓ+1) = v(a((ba)kaℓ−k − γ′)ak−ℓ) > (k − ℓ+ 1)v(a), and on the other, v(a(ba)k − γ′ak−ℓ+1) = v((ab)ka− γ′ak−ℓ+1) = v(((ab)kaℓ−k − γ′)ak−ℓ+1) = (k − ℓ+ 1)v(a) since (ab)kaℓ−k − γ′ = (aba−1b−1ba)kaℓ−k − γ = βk(ba)kaℓ−k − γ ̸= 0. In the last equation, we used that (aba−1b−1ba)kaℓ−k = (β′ba)kaℓ−k = β′k(ba)kaℓ−k = βk(ba)kaℓ−k since D = Z(D) as presumed. This holds if βk ̸= 1. If βk = 1, we repeat our calculations with roles of a and b interchanged. The new β will now be the inverse value of the former and since gcd(ℓ, k) = 1, β−ℓ ̸= 1. In either case, we get a contradiction from which we deduce v[a, b] > v(ab). Remark 2.3. The condition D = Z(D) is fulfilled by every valuation on an algebra over a field that is isomorphic to the residue field. In particular, this holds for minimal order- compatible valuations on R-algebras. Corollary 2.4. Let v be a valuation on a division ring D such that D = Z(D). If the value group has rational rank one, then v is strongly abelian. L. Vukšić: Valuations and orderings on the real Weyl algebra 235 Lemma 2.5. Let v be a valuation on a division ring D such that the value group is abelian and D = Z(D). Then for all x, y ∈ D \ {0}: (1) If v[x, y] > v(xy), then v[xm, y] > v(xmy) for all m ∈ Z. (2) Suppose v[x, y] = v(xy). Then v[x−1, y] = v(x−1y) and for each m ∈ N, v[xm, y] > v(xmy) holds if and only if α := y−1x−1yx satisfies 1 + α + · · · + αm−1 = 0 in D. Proof. To prove (1), first observe [x−1, y] = x−1y − yx−1 = x−1(yx− xy)x−1, so if v[x, y] > v(x, y), then v[x−1, y] = v[x, y]− 2v(x) > v(xy)− 2v(x) = v(x−1y). Suppose m ∈ N. Then [xm, y] = m∑ ℓ=1 xm−ℓ[x, y]xℓ−1 and since the value group is commutative, v(xm−ℓ[x, y]xℓ−1) = (m− 1)v(x) + v[x, y] > v(xmy) for each 1 ≤ ℓ ≤ m. Item (1) is thus proved. To prove (2), suppose v[x, y] = v(xy). Then v[x−1, y] = v(x−1y) is proved as for the first case. Since the value group is abelian, v[xm, y] ≥ v(xmy) holds, so we can observe y−1x−m[xm, y] = y−1x−m m∑ ℓ=1 xm−ℓ[x, y]xℓ−1 = m∑ ℓ=1 y−1x−ℓ[x, y]xℓ−1. For each 1 ≤ ℓ ≤ m, y−1x−ℓ[x, y]xℓ−1 = (αx−1)ℓ−1y−1x−1[x, y]xℓ−1 = αℓ−1x−ℓ+1y−1x−1[x, y]xℓ−1 = y−1x−1[x, y]αℓ−1. We can change the order of α and x−1 by Proposition 2.2 since v(α) = 0. The last equation follows from v(y−1x−1[x, y]) = 0 and Proposition 2.2. So now we have y−1x−m[xm, y] = y−1x−1[x, y] m−1∑ ℓ=0 αℓ, which proves the equivalence in (2). Proposition 2.6. Let v be a valuation on a division ring D such that the value group is abelian and D = Z(D). Suppose the residue field is formally real and suppose v[x, y] = v(xy) for some x, y ∈ D. Then v[xm, y] = v(xmy) for all odd m > 2. If v[xm, y] > v(xmy) for some even m, then v[x2, y] > v(x2y) and there is no a ∈ D such that v(a2) = v(x). 236 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 Proof. Suppose v[x, y] = v(xy). If m is odd, ∑m−1 ℓ=0 α ℓ = 0 does not have a solution in the residue field. By Lemma 2.5, it follows that v[xm, y] = v(xmy). Now consider the case for even m. If v[xm, y] > v(xmy), then α = −1 by Lemma 2.5 and v[x2, y] > v(x2y). Suppose a ∈ D satisfies v(a2) = v(x). We will first show that v[a2, y] = v(a2y). Assume that v[a2, y] > v(a2y). Then on one hand, a−2x = a−2xy−1y = y−1a−2xy, since v(a−2x) = 0. On the other hand, a−2x = y−1ya−2x = y−1a−2yx, where the last equation follows from v[a2, y] > v(a2y), or, by Lemma 2.5 equivalently, v[a−2, y] > v(a−2y). From x−1a−2(xy − yx) = 0 we conclude v[x, y] > v(xy), which is a contradiction. So v[a2, y] = v(a2y) and v[a, y] = v(ay) follows from Lemma 2.5. Now we show v[a4, y] > v(a4y). On one hand, we can write a−4x2 = y−1a−4x2y = y−1a−4yx2 since v[x2, y] > v(x2y). On the other hand, a−4x2 = y−1ya−4x2, so we conclude v[a4, y] > v(a4y). But by Lemma 2.5, v[a4, x] > v(a4x) gives us xax−1a−1 = −1. But then, again by Lemma 2.5, v[a2, x] > v(a2x). The proposition is thus proved. Proposition 2.7. Let v be a valuation on a division ring D such that such that D = Z(D), the value group is abelian and 2-divisible and the residue field is formally real. Suppose the value group of v is of rational rank 2 and suppose there are x, y ∈ D∗ such that v(x) and v(y) are rationally independent with v[x, y] > v(xy). Then v is strongly abelian. Proof. Suppose a, b ∈ D. Since the value group is abelian, v[a, b] ≥ v(ab). Suppose v[a, b] = v(ab). Then v(ak1) = v(x−m1y−n1) and v(bk2) = v(x−m2y−n2) for some ki,mi, ni ∈ Z, i = 1, 2. We conclude from Lemma 2.5 that v[ak1 , bk2 ] = v(ak1bk2). This is immediate if k1 and k2 are both odd. If k1 or k2 is even, v[ak1 , bk2 ] = v(ak1bk2) follows from the 2-divisibility of the value group and Proposition 2.6. Let c := xm1yn1 and d := xm2yn2 . Then on the one hand, ak1cbk2d = cak1dbk2 = dcak1bk2 since v(ak1) and v(c) are rationally dependent, v(bk2) and v(d) are rationally dependent and v(bk2d) = v(ak1c) = 0. On the other hand, ak1cbk2d = bk2dak1c = dbk2cak1 = cdbk2ak1 = dcbk2ak1 . Here, the last equality follows from v[x, y] > v(xy) and Lemma 2.5. Thus we have v(dc(ak1bk2 − bk2ak1)) > 0, so we get v[ak1 , bk2 ] > v(ak1bk2) which contradicts our assumption v[a, b] = v(a, b). We conclude v[a, b] > v(ab). The proof of the following proposition is the same as the proof of Proposition 2.7. L. Vukšić: Valuations and orderings on the real Weyl algebra 237 Proposition 2.8. Let v be a valuation on a division ring D such that D = Z(D), the value group is abelian and the residue field is formally real. Suppose the value group of v is of rational rank 2 and suppose there are x, y ∈ D∗ such that for every z ∈ D, v(zk) = v(x−my−n) holds for some k,m, n ∈ Z where k is odd. Then v is strongly abelian. We will later use this result to show that all valuations v on A1(R) with residue field R are strongly abelian. Propositions 2.7 and 2.8 can be easily generalized to higher rational ranks of the value group. The proofs are analogous. Corollary 2.9. Let v be a valuation on a division ring D such that D = Z(D). Suppose the value group is abelian and 2-divisible of rational rank n and that there are x1, . . . , xn ∈ D such that v(x1), . . . , v(xn) are rationally independent with v[xi, xj ] > v(xixj) for all i, j. Then v is strongly abelian. Corollary 2.10. Let v be a valuation on division ring D such that D = Z(D). Suppose the value group is abelian and of rational rank n and that there are x1, . . . , xn ∈ D such that for every z ∈ D, v(zk) = v(xm11 · · ·xmnn ) for some k,m1, . . . ,mn ∈ Z with k odd. Then v is strongly abelian. 3 Valuations on A1(R) We now describe the construction of all valuations on A1(R) with residue field R that was sketched in [22] over the ground field of C. Since every f ∈ A1(R) can be written as∑ m,n≥0 αm,nx myn, the construction will be similar to the construction of all valuations on the field of rational functions R(x, y) with residue field R (examples of constructions of such valuations can be found in [11] or [13]), but with some additional constraints arising from the fact that the generators x, y ∈ A1(R) satisfy [y, x] = 1. We first note that it follows from Theorem 5.3 of [16] that the value group of any valuation on A1(R) is commutative. Also, since every valuation v on A1(R) can be uniquely extended to its quotient division ring D1(R), our construction will take place in the quotient ring as we will use inverses. To construct a valuation v trivial on R with residue field R, we compare v(x) and v(y). It is easy to show, as it was done in [16], that v(xy) = v(yx) < 0, so v(x) or v(y) will be less than zero. Without loss of generality, we can set v(x) = −1 ∈ Q and compare it to v(y). If v(y) ̸∈ Q, then we get v( ∑ m,n≥0 αm,nx myn) = min m,n {mv(x) + nv(y)} for all elements of A1(R). Otherwise, v(y) = m1n1 ∈ Q. It follows that x m1yn1 = β1 ∈ R, so v(xm1yn1 − β1) > 0. Set ω1 := x m1yn1 − β1 and as before compare v(ω1) to v(x) in terms of rational dependence. If v(ω1) = m2n2 ∈ Q, then xm2ωn21 = β2 for some β2 ∈ R. Hence ω2 := x m2ωn21 − β2 238 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 also has value greater than zero. We continue this procedure. If we additionally define ω−1 = x and ω0 = y, we thus get a sequence (ωi)i≥−1, ωi ∈ A1(R) which ends with ωn for some n ∈ N if v(ωn) ̸∈ Q or is infinite otherwise. By the end of this section, we will prove a necessary and sufficient condition for the possibility to extend v from (ωi)i≥−1 to a valuation on A1(R) with residue field R. Every such extension from (ωi)i≥−1 to A1(R) will be uniquely determined. We will also show that every valuation on A1(R) with residue field R is strongly abelian. 3.1 Properties of the sequence (ωi)i≥−1 associated to a valuation on A1(R) Thorough this subsection let v be a valuation on A1(R). Lemma 3.1. Suppose (ωi)i≥−1 ⊆ A1(R) is a sequence as described above, with ω−1 = x, ω0 = y, ωi = xmiωnii−1 − βi for all i ≥ 0. Then [ωi, x] equals xmi ni∑ ℓi=1 ωni−ℓii−1 ( xmi−1 ni−1∑ ℓi−1=1 ω ni−1−ℓi−1 i−2( · · · ( xm2 n2∑ ℓ2=1 ωn2−ℓ21 n1x m1yn1−1ωℓ21 ) ωℓ32 ) · · · ) ωℓii−1 for each i ≥ 1. Proof. We prove the lemma by induction on i. If i = 1, [ω1, x] = [x m1yn1 − β1, x] = xm1 n1∑ ℓ1=1 yn1−ℓ1 [y, x]yℓ1−1 = n1x m1yn1−1. Now suppose that the equality holds for [ωi, x]. Then we have [ωi+1, x] = [x mi+1ω ni+1 i − βi+1, x] = x mi+1 [ω ni+1 i , x] = xmi+1 ni+1∑ ℓi+1=1 ω ni+1−ℓi+1 i [ωi, x]ω ℓi+1−1 i We can then proceed by the induction hypothesis. Before proving the next lemma, we define an equivalence relation between nonzero elements of A1(R) that have the same v-value, but their difference does not. For any a, b ∈ A1(R) \ {0}, we write a ∼ b if v(a) = v(b) < v(a− b). This is also a congurence relation, as ac ∼ bc and ca ∼ cb holds for all a, b, c ∈ A1(R) \ {0} with a ∼ b. Lemma 3.2. Suppose v is a valuation on A1(R) with residue field R and suppose (ωi)i≥−1 is a sequence such that ω−1 = x, ω0 = y, ωi = xmiωnii−1 − βi, v(x) = −1 and v(ωi) = mi+1 ni+1 for all i ≥ 0 up to either some n ≥ 0 in which case v(ωn) ̸∈ Q, or up to infinity. Then v[ωj , ωi] > v(ωiωj) for all i, j ≤ k if and only if v(Πkℓ=−1ωℓ) < 0, where k ≤ n in case v(ωn) ̸∈ Q for some n ≥ 0. L. Vukšić: Valuations and orderings on the real Weyl algebra 239 If any and hence both sides of the equivalence hold, then v[ωj , ωi] = −v(xyω1 · · ·ωi−1ωi+1 · · ·ωj−1) for all i < j ≤ k. Proof. Suppose v is a valuation on A1(R) and (ωi)i≥−1 is a sequence as described in the lemma. So v(ωi) ∈ Q either for all i ≥ 0 or for all 0 ≤ i < n for some n ≥ 0 and v(ωn) ̸∈ Q. It follows from Proposition 2.2 that v[ωi, ωj ] > v(ωiωj) for all i, j < n since v(ωi) and v(ωj) are rationally dependent. We shall use this fact to evaluate v[ωi, ωj ] for all i, j ≤ k ≤ n. It follows from Lemma 3.1 that [ωk, x] is a sum of products P , all equal to yn1−1xm1ωn2−11 x m2 · · ·ωnk−1k−1 xmk up to the order of factors. Since v[ωi, ωj ] > v(ωiωj) for all i, j ≤ k − 1, P ∼ yn1−1xm1ωn2−11 xm2 · · ·ω nk−1 k−1 x mk holds for every product P of the sum. Since v(yn1−1xm1ωn2−11 x m2 · · ·ωnk−1k−1 x mk) + v(yω1 · · ·ωk−1) = k∑ i=1 v(xmiωnii−1) = 0, we can conclude v[ωk, x] = v(yn1−1xm1ωn2−11 x m2 · · ·ωnk−1k−1 xmk) = −v(yω1 · · ·ωk−1). It follows that v[x, ωk] > v(xωk) if and only if v(xyω1 · · ·ωk) < 0. We will now prove that v[ωi+k, ωi] = −v(xyω1 · · ·ωi−1ωi+1 · · ·ωi+k−1) by induction on k, 1 ≤ k ≤ n − i. It will then follow that v[ωi, ωj ] > v(ωiωj) for all i, j ≤ n if and only if v(Πnℓ=−1ωℓ) < 0. If k = 1, then [ωi+1, ωi] = [x mi+1ω ni+1 i − βi+1, ωi] = [x mi+1 , ωi]ω ni+1 i = ( mi+1∑ ℓ=1 xmi+1−ℓ[x, ωi]x ℓ−1)ω ni+1 i , and since [x, ωi] is a sum of products all equal to yn1−1xm1ωn2−11 x m2 · · ·ωni−1i−1 xmi up to the order of factors, we can, using v[ωi, ωj ] > v(ωiωj) for j < i, deduce that v[ωi+1, ωi] = −v(xyω1 · · ·ωi−1) just like we did when evaluating v[ωi, x]. For k > 1, we have [ωi+k, ωi] = [x mi+kω ni+k i+k−1 − βi+k, ωi] = [x mi+k , ωi]ω ni+k i+k−1 + x mi+k [ω ni+k i+k−1, ωi] = ( mi+k∑ ℓ=1 xmi+k−ℓ[x, ωi]x ℓ−1)ω ni+k i+k−1+ xmi+k( ni+k∑ ℓ=1 ω ni+k−ℓ i+k−1 [ωi+k−1, ωi]ω ℓ i+k−1) ∼ mi+kxmi+k−1ω ni+k i+k−1[x, ωi] + ni+kx mi+kω ni+k−1 i+k−1 [ωi+k−1, ωi] and using both Lemma 3.1 and induction on k, we see that the first sum has v-value equal to −v(xyω1 · · ·ωi−1) and the second has v-value equal to −v(xyω1 · · ·ωi+k−1). Since the latter is smaller, it is equal to v[ωi+k, ωi]. This proves the lemma. 240 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 It follows that if v can be extended from (ωi)i≥−1 to a valuation on A1(R),∑k i≥−1 v(ωi) must be strictly less than 0 for all k ≤ n in case v(ωn) ̸∈ Q for some n, and for all k ≥ 0 if v(ωi) ∈ Q for all i ∈ N. We will now describe a necessary condition for the residue field to be R and then proceed to show that if both conditions are fulfilled, v can be extended from (ωi)i≥−1 to a valuation on A1(R) with residue field R. To ensure that the residue field is R, it is obviously necessary that ωkii−1ω kj j−1 ∈ R holds for all ki, kj ∈ Z with ki mini + kj mj nj = 0. For given i, j ≥ 0, all solutions (ki, kj) ∈ Z2 to the diophantine equation kimjni + kjminj = 0 (3.1) are integer multiples of the pair (Ki,j ,−Kj,i) with Ki,j = mjni di,j , Kj,i = minj di,j , where di,j = gcd{mjni,minj}. So for all ki, kj ∈ Z with ki mini + kj mj nj = 0, we can write ωkii−1ω kj j−1 = ω nKi,j i−1 ω −nKj,i j−1 = (ω Ki,j i−1 ω −Kj,i j−1 ) n for some n ∈ Z, where we used Proposition 2.2 in the second equality. So for every ki, kj ∈ Z satisfying 3.1, ωkii−1ω kj j−1 is uniquely determined by ω Ki,j i−1 ω −Kj,i j−1 . For each i, j ≥ 0, we define αi,j = ω Ki,j i−1 ω −Kj,i j−1 . We immediately see that αj,i = α −1 i,j and hence αi,i = 1 for all i, j ≥ 0. As α di,j i,j = ω Ki,jdi,j i−1 ω −Kj,idi,j j−1 = ω mjni i−1 ω −minj j−1 = β mj i β −mi j for all i, j ≥ 0, αi,j is one of the possible di,j-th roots for β mj i β −mi j . If v is a valuation on D1(R) with residue field R, αi,j must be real for all i, j ≥ 0. For every i, j ≥ 0 with even di,j , this means that β mj i β −mi j > 0 must hold. In the next lemma, we present a necessary condition on the sequence (βi)i≥1 so that αi,j ∈ R can be chosen for all i, j. We also prove that if ni is odd, αi,j is uniquely determined for all j ≥ 0. Lemma 3.3. Let v be a valuation on D1(R) as in Lemma 3.2. Then the following holds: (1) If ni is odd, there is a unique possible choice for αi,j ∈ R for all j ≥ 0. (2) Only if sgn(βi) is constant on the set of all i ≥ 0 for which ni is even can we choose αi,j ∈ R for all i, j ≥ 0. Proof. Suppose ni is odd. Then for any j ≥ 0, let d̃i,j be the highest odd number dividing di,j . Since ni is odd, ℓ1 := d̃i,jmj di,j ∈ Z. If nj is odd as well, ℓ2 := d̃i,jmidi,j ∈ Z holds too. If nj is even, mj is odd, so di,j = ˜di,j as di,j divides mjni. In both cases, ℓ2 ∈ Z holds. L. Vukšić: Valuations and orderings on the real Weyl algebra 241 Then for ℓ := ℓ1mi = ℓ2mj = d̃i,jmimj di,j we can evaluate α d̃i,j i,j = ω Ki,j d̃i,j i−1 ω −Kj,id̃i,j j−1 = x ℓx−ℓωℓ1nii−1 ω −ℓ2nj j−1 = (x miωnii−1) ℓ1(xmjω nj j−1) −ℓ2 = βℓ1i β −ℓ2 j , and since d̃i,j is odd, αi,j ∈ R is uniquely determined. The first point of the lemma is thus proven. To prove the second point of the lemma, suppose i, j ≥ 0 are such that ni and nj are both even. As a consequence, both mi and mj are odd while di,j is even. So, provided αi,j ∈ R, we compute 1 = sgn(α di,j i,j ) = sgn(β mj i β −mi j ) = sgn(βiβj), which proves the second part of the lemma. For even di,j we have seemingly two choices for αi,j ∈ R – a positive and a nega- tive one. We will show that in most cases, we cannot choose sgn(αi,j) for all i, j ≥ 0 independently of each other. Before that, we observe that for any i, j ≥ 0, at most one of Ki,j and Kj,i is even. In fact, if at most one of ni and nj is odd, Ki,j is odd if and only if ni is divisible by the greatest power of two that divides nj . For each i ≥ 1, let 2hi be the biggest power of two that divides ni. Define also m0 = 1 and n0 = −1. Proposition 3.4. Let v be a valuation on D1(R) associated to a sequence (ωi)i≥−1 with v(ω−1) = v(x) = −1, v(ωi−1) = mini with gcd(mi, ni) = 1 and x miωnii−1 = βi ∈ R for each i ≥ 1. Suppose sgn(βi) is constant on the set of all i ≥ 0 for which ni is even. Suppose αi,j ∈ R is determined for all i, j ≥ 0. Then Πri=0ω ki i−1 ∈ R is uniquely determined for each set of integers k0, k1, . . . , kr ∈ Z with ∑r i=0 ki mi ni = 0 if and only if for each a, b, c ≥ 0, αa,bαa,cαb,c > 0 whenever ha = hb ≤ hc holds. Proof. To prove the necessity of the condition, suppose a, b, c ≥ 0 are such that ha = hb ≤ hc holds. Suppose Ka,b and Kb,c are both odd. Choose ka, kb, kc ∈ Z \ {0} such that kamana + kb mb nb + kc mc nc = 0 and that ka and kb are odd while kc is even. Then on one hand, sgn(ωkaa−1ω kb b−1ω kc c−1) = sgn(ω ka a−1ω kb b−1ω kc c−1 Ka,b ) = sgn(ωkaa−1ω kb b−1ω kc c−1 Ka,b ω −kaKb,a b−1 ω kaKb,a b−1 ) = sgn(αkaa,bω ℓb b−1ω ℓc c−1) with ℓb = kbKa,b + kaKb,a, ℓc = kcKa,b. As v(ωℓbb−1ω ℓc c−1) = 0, (ℓb, ℓc) = ℓ(Kb,c,−Kc,b) for some ℓ ∈ Z with −ℓKc,b = ℓc = kcKa,b. So we can conclude sgn(ωkaa−1ω kb b−1ω kc c−1) = sgn(α ka a,bα ℓ b,c). 242 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 On the other hand, we see by analogous computations that sgn(ωkaa−1ω kb b−1ω kc c−1) = sgn(ω ka a−1ω kb b−1ω kc c−1 Ka,c ) = sgn(ωkaa−1ω kb b−1ω kc c−1 Ka,c ω −kaKc,a c−1 ω kaKc,a c−1 ) = sgn(αkaa,cω ℓ′b b−1ω ℓ′c c−1) = sgn(α ka a,cα ℓ′ b,c) for some ℓ′b, ℓ ′ c, ℓ ′ ∈ Z with ℓ′b = kbKa,c = ℓ ′Kb,c, ℓ′c = kcKa,c + kaKc,a. We have chosen ka, kb odd and kc even. In this case, the greatest power of two that divides kc is 2hc−ha+1. On the other hand, the greatest power of two that divides Kc,a and Kc,b is 2hc−ha . We can thus conclude from ℓKc,b = −kcKa,b and ℓ′Kb,c = kbKa,c that ℓ is even while ℓ′ is odd since Ka,b,Ka,c and Kb,c are all odd. So we see that sgn(ωkaa−1ω kb b−1ω kc c−1) = sgn(αa,b) = sgn(αa,cαb,c), which proves the necessity of the condition. Now suppose sgn(αa,bαa,cαb,c) = 1 for all a, b, c ≥ 0 with ha = hb ≤ hc. Let K := (k0, . . . , kr) ∈ Zr+1 be such that ∑r i=0 ki mi ni = 0. Let suppK := {i | ki ̸= 0} and n := | suppK|. We prove that Πri=0ω ki i−1 ∈ R is uniquely determined by induction on n ≥ 2. We first suppose 0 ̸∈ suppK. We will deal with the case 0 ∈ suppK at the end of our proof. If n = 2, then Πri=0ω ki i−1 = ω ki i−1ω kj j−1 for some i, j > 0, and its value in the residue field is a power of αi,j . Now suppose n > 2. Take two distinct a, b ∈ suppK. As at least one of Ka,b and Kb,a is odd, so suppose Ka,b is odd. Then Πri=1ω ki i−1 Ka,b = Πri=1ω ki i−1 Ka,b ω −kaKb,a b−1 ω kaKb,a b−1 = Π r i=1ω ℓi,1 i−1α ka a,b with ℓa,1 = 0, ℓb,1 = kaKa,b + kbKb,a and ℓi,1 = kiKa,b for i ̸= a, b. Since |{i | ℓi,1 ̸= 0}| is strictly smaller than n, Πri=0ω ℓi,1 i−1 ∈ R is uniquely determined by the induction hypothesis. So we have determined Πri=0ω ki i−1 Ka,b ∈ R. As Ka,b is odd, Πri=0ω ki i−1 ∈ R is determined as well. We now need to show that in this way, Πri=1ω ki i−1 is uniquely determined, that is, if we choose another a′, b′ ∈ suppK instead of a, b, we get the same value for Πri=1ω ki i−1 ∈ R. We will show this by choosing c ∈ suppK \ {a, b} and proving that the evaluated value of Πri=1ω ki i−1 is the same whether we factor a power of αa,b as above, or αa,c or αb,c instead. By transitivity of the equality relation, this will imply that the obtained value of Πri=1ω ki i−1 is independent of the choice of a, b ∈ suppK. Suppose without loss of generality that ha ≤ hb ≤ hc and that Ka,b,Ka,c and Kb,c are odd. Above, we have evaluated Πri=1ω ki i−1 Ka,b = Πri=1ω ℓi,1 i−1α ka a,b L. Vukšić: Valuations and orderings on the real Weyl algebra 243 with ℓa,1 = 0, ℓb,1 = kaKa,b + kbKb,a and ℓi,1 = kiKa,b for i ̸= a, b. We proceed by evaluating, in the same way as before, Πri=1ω ℓi,1 i−1 Kb,c = Πri=1ω pi,1 i−1α ℓb,1 b,c with pa,1 = pb,1 = 0, pc,1 = ℓc,1Kb,c + ℓb,1Kc,b and pi,1 = ℓi,1Kb,c for i ̸= a, b, c. So Πri=1ω ki i−1 Ka,bKb,c = Πri=1ω pi,1 i−1α kaKb,c a,b α ℓb,1 b,c (3.2) with ℓi,1 and pi,1 for all 0 ≤ i ≤ r as above. In particular, we see that for i ̸= a, b, c, pi,1 = ℓi,1Kb,c = kiKa,bKb,c. Similarly, we can compute Πri=1ω ki i−1 Ka,cKb,c = (Πri=1ω ℓi,2 i−1α ka a,c) Kb,c = Πri=1ω pi,2 i−1α kaKb,c a,c α ℓb,2 b,c (3.3) with 1. ℓa,2 = 0, ℓc,2 = kaKa,c + kcKc,a ℓi,1 = kiKa,c for i ̸= a, c, and 2. pa,2 = pb,2 = 0, pc,2 = ℓb,2Kc,b + ℓc,2Kb,c, pi,2 = kiKa,cKb,c for i ̸= a, b, c. Let N := Ka,bKa,cKb,c. On one hand, we see from 3.2 that Πri=1ω ki i−1 N = Πri=1ω pi,1 i−1 Ka,c α kaKb,cKa,c a,b α ℓb,1Ka,c b,c , (3.4) and on the other hand, we see from 3.3 that Πri=1ω ki i−1 N = Πri=1ω pi,2 i−1 Ka,b α kaKb,cKa,b a,c α ℓb,2Ka,b b,c . (3.5) We need to show that in both equations, we get the same value. We first see that for all i ̸= c, pi,1Ka,c = pi,2Ka,b. So, given that r∑ i=1 pi,1 mi ni = r∑ i=1 pi,2 mi ni = 0, we can see pc,1Ka,c = pc,2Ka,b holds as well, and thus we conclude Πri=1ω pi,1 i−1 Ka,c = Πri=1ω pi,2 i−1 Ka,b . It then follows that α kaKa,cKb,c a,b α ℓb,1Ka,c b,c = α kaKb,cKa,b a,c α ℓb,2Ka,b b,c , since both sides of the equation are equal to ωNkaa−1 ω Nkb b−1 ω Nkc−pc,1Ka,c c−1 and the signs of αa,b, αa,c and αb,c were chosen so that the signs of both sides of the equality match. We conclude that the value of Πri=1ω ki i−1 N is the same in both 3.4 and 3.5. As N is odd (since Ka,b, Ka,c and Kb,c are all odd), we conclude that Πri=1ω ki i−1 is the same whether we factor a power of αa,b or αa,c. If we factored a power of αb,c, we would, as similar computations as above would show, get the same value for Πri=1ω ki i−1. 244 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 We have now shown that if the condition of the proposition is fulfilled, Πri=1ω ki i−1 is uniquely determined for all k1, . . . , kr ∈ Z with ∑r i=1 ki mi ni = 0. Now we consider the case 0 ∈ suppK. Let K = (k0, . . . , kr) ∈ Zr+1 be such that k0 ̸= 0 and ∑r i=0 ki mi ni = 0. Let N := gcd{ni | i ∈ suppK}. If N is odd, i.e., if ni is odd for every i ∈ suppK, then Πri=0ω ki i−1 is uniquely determined. This is because Πri=0ω ki i−1 N = (Πri=0x ki mi ni Πri=0ω ki i−1) N = Πi=1(xmiωnii−1) kici = Πri=1β kici i where ci := Nni for each i ≤ i ≤ r. We thus conclude Π r i=0ω ki i−1 ∈ R is the uniquely determined N -the real root of Πri=1β kici i . Now suppose nj is even for some j ∈ suppK. Then mj must be odd since gcd(mj , nj) = 1. Let k′j := kj − k0nj and k′i := ki for all i ∈ suppK \ {0, j}. Then ∑r i=1 k ′ i mi ni = 0 and Πri=0ω ki i−1 mj = (xmjωnj )k0(Πri=1ω k′i i−1) mj = βk0j (Π r i=1ω k′i i−1) mj . We evaluate (Πri=1ω k′i i−1) mj as above k0 = 0 and conclude that Πri=0ω ki i−1 ∈ R is the unique mj-th real root of βk0j (Π r i=1ω k′i i−1) mj . This concludes the proof of our proposition. In Lemma 3.6, we suppose that v is a valuation on D1(R) extended from (ωi)i≥−1 to D1(R) and compute the value of certain elements of D1(R) in this case. Lemma 3.5. Let D be a division ring endowed with a valuation v with an abelian value group and a commutative residue field with characteristic zero. Let a, b ∈ D be such that a ∼ b, v(a) = v(b) = 0 and v(ab) < v[a, b]. Then v(an − bn) = v(a − b) for all n ∈ Z \ {0}. If there exist c, d ∈ D such that cn = a, dn = b, c = d for some n ∈ N, then v(cm − dm) = v(a− b) for all m ∈ Z. Proof. For n ∈ N, write an − bn = (a− b) ∑n−1 i=0 a n−1−ibi + terms with higher v-value. Since an−1−ibi is the same for all 0 ≤ i ≤ n − 1, the v-value of the sum is equal to zero, proving the statement for positive integers n. For negative n ∈ Z, the statement follows from an − bn = −an(a−n − b−n)bn. The last statement of the lemma follows from a− b ∼ (c− d) ∑n−1 i=0 c n−1−ibi. Lemma 3.6. Suppose v is a valuation on A1(R) and suppose i1, i2, . . . , ir ∈ N and k0 ∈ Z, ki1 , ki2 , . . . , kir ∈ Z \ {0} are such that v(xk0ω ki1 i1−1 · · ·ω kir ir−1) = 0. If min1≤j≤r{v(ωij )} is achieved at exactly one j, then v(xk0ω ki1 i1−1 · · ·ω kir ir−1 − xk0ω ki1 i1−1 · · ·ω kir ir−1) = min{v(ωij ) | 1 ≤ j ≤ r}. Proof. Let n be the least common multiple of ni1 , . . . , nir and cij = n nij for each ij . Since (xk0ω ki1 i1−1 · · ·ω kir ir−1) n = Πrj=1x nkij mij nij (ω ki1 i1−1 · · ·ω kir ir−1) n = Πrj=1(x mijω nij ij−1) kij cij = Πrj=1β kij cij ij , L. Vukšić: Valuations and orderings on the real Weyl algebra 245 by Proposition 2.2, we can compute (xk0ω ki1 i1−1 · · ·ω kir ir−1) n −Πrj=1β kij cij ij ∼ r∑ j=1 (xmi1ω ni1 i1−1) ki1ci1 · · · (xmij−1ω nij−1 ij−1−1) kij−1cij−1 · ((xmijω nij ij−1) kij cij − β kij cij ij )β kij cij ij+1 · · ·β kij cij ir . For each j such that kij > 0, (xmijω nij ij−1) kij cij − β kij cij ij ∼ ωij kij cij−1∑ i=0 (xmijω nij ij−1) kij cij−iβiij , which gives us v((xmijω nij ij−1) kij cij − β kij cij ij ) = v(ωij ). In case kij < 0, we see that (xmijω nij ij−1) kij cij − β kij cij ij = − (xmijω nij ij−1) kij cij β kij cij ij ((xmijω nij ij−1) −kij cij − β −kij cij ij ), which again implies v((xmijω nij ij−1) kij cij − β kij cij ij ) = v(ωij ). We conclude, using Lemma 3.5, v(xk0ω ki1 i1−1 · · ·ω kir ir−1 − xk0ω ki1 i1−1 · · ·ω kir ir−1) = v((xk0ω ki1 i1−1 · · ·ω kir ir−1) n − (xk0ωki1i1−1 · · ·ω kir ir−1) n) = min{v(ωij ) | 1 ≤ j ≤ r}. With the help of Lemma 3.6, we will evaluate v(xk0ωki1i1−1 · · ·ω kir ir−1 − xk0ω ki1 i1−1 · · ·ω kir ir−1) when v(x k0ω ki1 i1−1 · · ·ω kir ir−1) = 0 in general. As in Lemma 3.6, we assume k0 ∈ Z, kij ∈ Z \ {0} for all 1 ≤ j ≤ r. This will be helpful when we will later construct a valuation v associated to a sequence (ωi)i≥−1. Let us assume for now that i1 < i2 < · · · < ir; at the end of the calculation we will see that the order of ij does not affect the v-value. To start, we introduce some abbreviations to make the written equations easier to read. Let n and cij for all j be as in the proof of Lemma 3.6, A0 := x k0ω ki1 i1−1 · · ·ω kir ir−1 B0 := A0 A (n) 0 := x nk0ω nki1 i1−1 · · ·ω nkir ir−1 B (n) 0 := A (n) 0 = Π r j=1β kij cij ij . Since B0 is in R, we can write A0 −B0 = (An0 −Bn0 )( n−1∑ i=0 An−i−10 B i 0) −1 ∼ (A(n)0 −B (n) 0 )( n−1∑ i=0 An−i−10 B i 0) −1. (3.6) 246 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 Since v( ∑n−1 i=0 A n−i−1 0 B i 0) = v(A n−i−1 0 B i 0) = 0 for all i, which holds due to A n−i−1 0 B i 0 = Bn−10 for all i, v(A0 −B0) = v(A (n) 0 −B (n) 0 ). To evaluate the right-hand side of (3.6), we first proceed as we have done in the proof of Lemma 3.6, so A (n) 0 −B (n) 0 ∼ r∑ j=1 Πj−1ℓ=1(x miℓω niℓ iℓ−1) kiℓciℓ · ((xmijω nij ij−1) kij cij − β kij cij ij )· Πrℓ=j+1β kij cij iℓ . If kij > 0, we proceed by (xmijω nij ij−1) kij cij − β kij cij ij ∼ ωij kij cij−1∑ p=0 (xmijω nij ij−1) kij cij−p−1βpij . For ij > 0, kij < 0, we can on the other hand write (xmijω nij ij−1) kij cij − β kij cij ij ∼ −ωij (x mijω nij ij−1) kij cij β kij cij ij −kij cij−1∑ p=0 (xmijω nij ij−1) −kij cij−p−1βpij . We now define, if kj > 0, Cj := kij cij−1∑ p=0 Πj−1ℓ=1(x miℓω niℓ iℓ−1) kiℓciℓ · (xmijω nij ij−1) kij cij−p−1βpij ·Π r ℓ=j+1β kiℓciℓ iℓ , and, if kj < 0, Cj := −(xmijω nij ij−1) kij cij β kij cij ij −kij cij−1∑ p=0 Πj−1ℓ=1(x miℓω niℓ iℓ−1) kiℓciℓ · (xmijω nij ij−1) −kij cij−p−1βpij ·Π r ℓ=j+1β kiℓciℓ iℓ for each 1 ≤ j ≤ r. Further, we define A1,j = ωijCj A (n) 0 −B (n) 0 ∼ r∑ j=1 ωijCj = r∑ j=1 A1,j for each 1 ≤ j ≤ r. Thus we can conclude that v(A1,j) = v(ωij ), since the image of Cj in the residue field is equal kijcij · Πrℓ=1β kiℓciℓ iℓ ̸= 0 if kij > 0 and −β 2kij cij ij |kijcij | · Πrℓ=1β kiℓciℓ iℓ ̸= 0 if kij < 0 , making v(Cj) = 0. We can now write A (n) 0 −B (n) 0 = r∑ j=1 A1,j +A = ∑ v(A1,j) is minimal A1,j + ∑ v(A1,j) is not minimal A1,j . L. Vukšić: Valuations and orderings on the real Weyl algebra 247 Here we note that the second of both finite sums on the right-hand side of this equation includes A, which denotes the sum of all terms obtained by changing the order of factors of the form ωiℓ (which was not explicitly written above). The fact that the v-value of these terms is higher than the v-value of the terms of the first sum (the ones with minimal v-value) follows from Lemma 3.2. If min1≤j≤r{v(ωij )} is achieved at more than one j, we take the sum of all ωijCj that have the minimal v-value, i.e., ∑ v(A1,j) is minimal A1,j , then factor ωi1 , so the sum now looks like∑ v(A1,j) is minimal A1,j = ωi1 ∑ v(A1,j) is minimal ω−1i1 A1,j = ωi1 ∑ v(A1,j) is minimal ω−1i1 ωijCj and, since v(ω−1i1 ωijCj) = 0, for each j, we can evaluate the sum of their images in the residue field. If this sum is not equal to zero, then v(A0 − B0) = min1≤j≤r{v(A1,j)}. Otherwise write ωi1 ∑ j ω−1i1 ωijCj = ∑ j ωi1(ω −1 i1 ωijCj − ω−1i1 ωijCj). For every j, we write ω−1i1 ωijCj − ω −1 i1 ωijCj as an R-linear sum of terms of the form Πℓω nℓ iℓ−1 (in the same way we did with A0 −B0). We sum all of the newly obtained terms, as well as the terms in ∑ v(A1,j) is not minimal A1,j , and relabel them as A2,j where j goes from 1 to the number of all terms. As A0 −B0 can be written in the form A0 −B0 = ( ∑ v(A2,j) is minimal A2,j + ∑ v(A2,j) is not minimal A2,j)D, where we use D as the label of the product of all terms of the form ( ∑ i A k−iBi)−1 where A = xk ′ 0Πrj=1ω k′ij ij−1 for some i1, . . . , ir ∈ N, k ′ 0, k ′ i1 , . . . , k′ir ∈ Z and B = A, that we factor out when we evaluate ω−1i1 ωijCj −ω −1 i1 ωijCj for each j. All terms ( ∑ i A k−iBi)−1 have v-value equal to zero and their image in the residue field, which is of the form (mBm−1)−1 ∈ R for some m ∈ N, is easy to determine. We repeat the described procedure, writing A0 − B0 = ( ∑ j Ak,j)D for increasing k. We stop when for some k, ∑ j,v(Ak,j) is minimal Ak,j is either composed of one single term or, after factoring out one of the terms, the image of the sum in the residue field is not zero. In this case, we conclude that v(A0 − B0) is equal to v(Ak,j) for any term of the sum∑ j,v(Ak,j) is minimal Ak,j . We must show that the process ends at some point even if the number of terms whose v- value we evaluate at each step is growing. We see that whenever we write xk0Πrℓ=1ω kℓ iℓ−1− xk0Πrℓ=1ω kℓ iℓ−1 as a sum of terms with strictly positive v-value, the value of each of these terms is v(ωiℓ) for some ℓ = 1, . . . , r. It follows that v(A0 − A0) is a sum of v(ωℓ) for some ℓ ≥ 1. If v(ωN ) is irrational for some N ≥ 0, the process either stops beforehand or, after k ≥ N − ir steps we get a unique term Ak,j that has v-value equal to v(ωirωir+1 · · ·ωN ). This is the term we get when we take the last term of A(n)0 − B (n) 0 , written as a sum of terms A1,j with higher v-value and in each of the following steps whenever the v-value of 248 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 this term is minimal, take the last term when Ak,j is written as a sum of terms with higher v-value. If on the other hand, v(ωk) ∈ Q for an infinite sequence (ωk)k≥−1, then limk→∞ v(ωk) = 0 since by Lemma 3.2, ∑k i≥−1 v(ωk) < 0 for all k ≥ −1 and v(ωi) > 0 for i ≥ 1. Then for some N ≥ 1, v(ωN ) < v(ωi) for all 1 ≤ i < N . The evaluation of v(A0 − B0) again either stops beforehand or we get a unique term Ak,j that has v-value equal to v(ωirωir+1 · · ·ωN ). As in the first case, this term is the one we get when we take the last term of A(n)0 − B (n) 0 , written as a sum of terms A1,j with higher v-value and in each of the following steps whenever the v-value of this term is minimal, take the last term when Ak,j is written as a sum of terms with higher v-value. In both cases, the value of this term, v(ωirωir+1 · · ·ωN ) is strictly smaller than the value of all additional terms we get when we change the order of the factors in a product. It follows that given the v-values of (ωi)i≥−1, v(A0 − B0) is the same as it would be if all elements of the sequence (ωi)i≥−1 commuted. This follows from the fact that when we change the order of factors ωi and ωj in some Ak,j1 , the term we obtain has v-value greater by v[ωi, ωj ] − v(ωiωj) = −v(xyω1 · · ·ωj) for −1 ≤ i < j ≤ N while v(Ak+1,j1) − v(Ak,j2) is for any j1, j2 equal to v(ωiℓ+1) for some iℓ such that Ak,j2 contains a power of ωiℓ . Since, as we have shown in the proof of Lemma 3.2, ∑N i=−1 v(ωi) < 0, it follows that the terms we obtain by changing the order of the factors have v-value greater than v(A0 − B0). And since v(xk0ω ki1 i1−1 · · ·ω kir ir−1 − xk0ω ki1 i1−1 · · ·ω kir ir−1) is higher than the v-value of any terms we get when we change the order of factors, the order of ωi1 , . . . , ωir does not matter. 3.2 Extending v from the sequence (ωi)i≥−1 to A1(R) We can now prove that every v associated to either a finite or an infinite sequence (ωi)i≥−1 can be extended to a valuation on D1(R). Lemma 3.7. For every r ≥ 0, there exists a finite number M of elements of the form ai := ω ki,−1 −1 ω ki,0 0 ω ki,1 1 · · ·ω ki,r−1 r−1 for some ki,1, . . . , ki,r−1 ∈ Z such that v(ai) = 0 for all 1 ≤ i ≤ M and every ωℓ−1−1 · · ·ω ℓr−1 r−1 ∈ D1(R) with v-value zero is ∼-equivalent to a product of positive integer powers of a1, . . . , aM . Proof. Since v(ω−1) = −1 and v(ωj) = mj+1nj+1 ∈ Q for j ≥ 0, the problem translates to finding general classes of solutions to the diophantine equation x0a0 + · · ·xkak = 0 with a0 = −Πkj=1ni and ai = mini Π k j=1ni for all 0 ≤ i ≤ k. Theorem 3.8. Let v and (ωi)i≥−1 be as described in the beginning of the section, i.e., ω−1 = x, ω0 = y, v(ω−1) = −1, v(ωi) = mi+1ni+1 ∈ Q, x mi+1ω ni+1 i = βi+1 ∈ R, ωi+1 = x mi+1ω ni+1 i − βi+1 for i, 0 ≤ i ≤ N − 1 and v(ωN ) ̸∈ Q for some N ≥ 0 or v(ωi) ∈ Q for infinitely many i, that ∑k i=−1 v(ωi) < 0 for all k ≥ −1. Suposse that sgnβi is constant on the set of all i for which ni is even. Then v can be extended to a valuation on D1(R) with residue field R. The valuation is unique for every choice of {αi,j}i,j≥0 where αi,j = ω Ki,j i−1 ω Kj,i j−1 ,Ki,j = mjni di,j ,Kj,i = −minjdi,j with di,j = gcd{mjni,minj}. The associated value group is group-isomorphic to a subgroup of Q × Z generated by {v(ωi)}i≥−1. L. Vukšić: Valuations and orderings on the real Weyl algebra 249 Proof. The following construction of the valuation v associated to the sequence (ωi)i≥−1 was first sketched in [22]. Here we present it in full detail. Before we begin with the construction of the v-value for an arbitrary element of D1(R), we define it for some specific elements of D1(R). 1. Since we have defined v(ωi) for all −1 ≤ i ≤ N , v(ΠNi=−1ω ki i ) = ∑N i=−1 kiv(ωi) must hold for all k−1, . . . , kN ∈ Z. 2. Since we supposed ∑k i=−1 v(ωi) < 0 for all k ≥ −1, it follows from Lemma 3.2 that v[ωi, ωj ] = −v(ω−1ω0 · · ·ωi−1ωi+1 · · ·ωj−1), which must be strictly greater than v(ωiωj) for all i < j. 3. We also see that if v(ωk0−1 · · ·ω kr r−1) = 0 for some k0, . . . , kr ∈ Z, then ω k0 −1 · · ·ω kr r−1 is uniquely determined by {αi,j}i,j≥0 as in shown in Proposition 3.4. In this case, v(ωk0−1 · · ·ω kr r−1−ω k0 −1 · · ·ω kr r−1) must be equal to the value determined in Lemma 3.6 and the discussion following it. In all three cases, the chosen values were the only possible extensions of v from (ωi)i≥−1 if we want v to be a valuation. To determine v(F ) for any F ∈ A1(R), we first note that F can be written as a finite sum F = ∑ ℓ αℓx iℓyjℓ , αℓ ∈ R. Let F1 be the sum of all terms αℓxiℓyjℓ such that iℓv(x) + jℓv(y) is equal to u := minℓ{iℓv(x) + jℓv(y)}. If F1 consists of only one such term, then we define v(F ) = u; this is obviously always the case whenever v(y) ̸∈ Q. Otherwise, we factor out xi1yj1 with the smallest power of x and get F1 ∼ xi1yj1 ∑ ℓ, iℓv(x)+jℓv(y)=u αℓx iℓ−i1yjℓ−j1 . Since (iℓ − i1)v(x) + (jℓ − j1)v(y) = 0, for each ℓ in the sum, iℓ−i1jℓ−j1 = Kℓ m1 n1 for some Kℓ ∈ Z. We can write F1 ∼ xi1yj1f(xm1yn1) where f(t) is a polynomial in R[t]. Since we know v(ω1) > 0 and v(α) = 0 for α ∈ R∗, v is uniquely determined on R[ω1]. From this, it follows that v(f(xm1yn1)) = 0 if and only if f(β1) ̸= 0 since xm1yn1 = ω1 + β1. In this case, v(F1) = u and since all terms in F − F1 have v-value strictly greater than u, v(F ) = u must hold. Since v(u) is a sum of integer powers of v(x) and v(y), v(F ) is in 250 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 the abelian group, generated by {v(ωi)}i≥−1. If u = 0, F = βk1f(β1) ∈ R. If on the other hand f(β1) = 0, write f(t) = g1(t)(t− β1)k1 with g1(β1) ̸= 0 and we have F1 ∼ xi1yj1g1(xm1yn1)ωk11 . We set v(F1) = u + k1v(ω1) and add all terms we get from exchanging the order of factors, whose v-value can be lower than the newly set v(F1), although still strictly higher than u due to v[x, y] > v(xy), to F − F1. It is immediate that v(F1) is in the subgroup of Γ, generated by v(x), v(y) and v(ω1) and that if v(F1) = 0, F1 ∈ R. It is important to note that in both cases, we consider F1 as a single term. It follows that during our transformation, the number of terms (if we ignore the ones we got when we changed the order of factors in a product) is strictly smaller than before (unless, of course, F1 was just a single term in the beginning and we get v(F ) = v(F1)). We now consider the values of the terms in F − F1. If all of them have v-value strictly greater than that of F1, we conclude v(F ) = v(F1). Otherwise, we take all terms of F − F1 = ∑ ℓ′, iℓ′v(x)+jℓ′v(y)>u αℓ′x iℓ′ yjℓ′ , αℓ′ ∈ R for which iℓ′v(x) + jℓ′v(y) = u′ := minℓ′{iℓ′v(x) + jℓ′v(y)} and then as before define F2 = x i2yj2 ∑ ℓ′, iℓ′v(x)+jℓ′v(y)=u ′ αℓ′x iℓ′−i2yjℓ′−j2 . As above, we write F2 ∼ xi2yj2g2(xm1yn1)(x − β1)k2 , k2 ≥ 0, g2(β1) ̸= 0 and add all the terms we get when we change the order of factors to F − F1 − F2. Their v-value is strictly greater than u′ due to v[x, y] > v(xy). We continue this process, defining F1, F2, . . . , Fk until all terms in F − F1 − · · · − Fk have v-value strictly greater than min{F1, . . . , Fk}. Note that it is possible that Fk consists of only one term from F − F1 − · · · − Fk−1. Afterwards, we sum together all those Fi for 1 ≤ i ≤ k for which v(Fi) = u1 := min{v(F1), . . . , v(Fk)}. If the minimum is achieved at exactly one such Fi, we set v(F ) = v(Fi). This is always the case whenever v(ω1) ̸∈ Q. Otherwise we can relabel the terms so the minimum is achieved at F1, . . . , Fr for some r ≤ k. As we have shown, each Fi can be written as Fi = xiiyjigi(xm1yn1)ωki1 . We sum the terms together, factor out x i1yj1ωk11 , the term that has, written as a polynomial in x and y, the lowest power of x, and label the new sum F1,1. To evaluate v(F1,1), we follow a procedure similar to the one evaluating v(F1). After factoring xi1yj1ωk11 , we are left with F1,1 ∼ xi1yj1ωk11 (g1(xm1yn1) + g2(xm1yn1)xi2−i1yj2−j1ω k2−k1 1 + · · · + gr(x m1yn1)xir−i1yjr−j1ωkr−k11 ) ∼ xi1yj1ωk11 R∑ J=1 αJx iJ yjJωkJ1 , with αJ ∈ R, iJ , jJ , kj ∈ Z, R ≥ 1. Each term in the sum has v-value zero. Let a1, a2, · · · , aℓ be the terms such that each product of the form xiyjωk1 , i, j, k ∈ Z that fulfills the condition v(xiyjωk1 ) = 0 is a L. Vukšić: Valuations and orderings on the real Weyl algebra 251 product of positive integer powers of some of ai up to the order of factors x, y and ω1. The existence of a1, . . . , aℓ is assured by Lemma 3.7. We can then write F1,1 ∼ xi1yj1ωk11 R∑ J=1 γJa m1,J 1 a m2,J 2 · · · a mℓ,J ℓ ∼ xi1yj1ωk11 g(a1, · · · , aℓ), g ∈ R[t1, · · · , tℓ] (3.7) with γJ ∈ R, mI,J ∈ Z for all 1 ≤ I ≤ ℓ and 1 ≤ J ≤ R. As before, we add all terms we get when we change the order of multiplication of x, y or ω1 in a product to F − F1,1 since the value of its terms is strictly greater than u1. Since, as we have determined in the beginning, each term in the sum (3.7) has v-value equal to zero and we know what ai ∈ R is for each 1 ≤ i ≤ ℓ, v(g(a1, . . . , aℓ)) will have to be greater than or equal to zero, we can define g(a1, . . . , aℓ) = g(a1, . . . , aℓ). If g(a1, . . . , aℓ) ̸= 0, then we set v(g(a1, . . . , aℓ)) = 0 and v(F ) = v(F1,1) = i1v(x) + j1v(y) + k1v(ω1). Otherwise write g(t1, . . . , tℓ) = h(t1 − a1, . . . , tℓ − aℓ) = L∑ i=1 Πℓj=1(tj − aj)mi,jhi(t1, . . . , tℓ) with h, h1, . . . , hL ∈ R[t1, . . . , tn] and hi(a1, . . . , aℓ) ̸= 0 for all i. We factor out the Πj(tj − aj)mi,j for those i for which ∑ j v(aj − aj)mi,j is minimal. Then g(t1, . . . , tℓ) = Πj(tj − aj)mi,j g̃(t1, . . . , tk), with g̃ ∈ R[t1, . . . , tℓ]. If g̃(a1, . . . , ak) ̸= 0, we set v(F1,1) = v(x i1yj1) + ∑ j mi,jv(aj − aj). If on the other hand, g̃(a1, . . . , ak) = 0, we do the same thing as we did with g. The process cannot go on indefinitely since g is a polynomial and hence of finite degree. All terms we get when we exchange the order of x, y and ω1 are added to F − F1,1. Their v-value must be strictly greater than u1. It follows from the construction that v(F1,1) must be in the group generated by {v(ωi)}i≥−1 since this holds for v(ai − ai) for all i and that if v(F1,1) = 0, F1,1 ∈ R. Since v(ai − ai) is, as we have shown in Lemma 3.6 and the discussion following it, a sum of v(ωj) and thus v(Πjω−1j (ai − ai)) = 0, we can write F1,1 as one term of the form Πni=−1ω ki i g(a1, a2, . . . , aℓ) with n ∈ N and g ∈ R[t1, . . . , tℓ] and g(a1, a2, . . . , aℓ) ̸= 0. After v(F1,1) is set, we compare it to both v(Fi) for all Fi that are not part of F1,1 and the terms of F − F1 − · · · − Fk − F1,1. If all of these terms have v-value strictly greater than v(F1,1), then we can set v(F ) = v(F1,1). Otherwise, we collect all terms with minimal v-value in a sum which we label F1,2. We determine v(F1,2) in the same way we determined F1,1 and then sum all of the remaining terms that have v-value less or equal to min{v(F1,1), v(F1,2)} to a sum labeled F1,3. We repeat the process until for some k, min{v(F1,1), · · · , v(F1,k)} is strictly smaller than the v-value of any of the remaining terms. If min{v(F1,1), · · · , v(F1,k)} is achieved at exactly one i, we set v(F ) = v(F1,i). Otherwise we sum all the terms with the minimal v-value and label the sum F2,1. We 252 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 evaluate v(F2,1) in the same way we evaluated v(F1,1). We repeat the process, defining Fi,j and determining its v-value in the same way as above. We point out that after v(Fi,j) is defined, we regard Fi,j as one single term in future evaluations. Now we must show that at one point, the process ends, i.e., that for some i, j, v(Fi,j) is strictly smaller than the v-value of all other terms. This holds because each time we define Fi,j for some i, j, we sum a number of different terms into one single term and because whenever we change the order of factors in a term, the degree of x and y in the difference is strictly smaller. This means that we eventually run out of terms. We have thus defined v for an arbitrary polynomial F ∈ A1(R). What we essentially did was that we wrote F = F̃ + F̃1 where F̃ is written as a single term, v(F̃ ) is computed as if x and y commuted and the v-value of each term of F̃1 is strictly greater than v(F̃ ). For another G ∈ A1(R), we can write FG = F̃G+ ˜(FG)1 and since we evaluate v(F̃ ) and v(G̃) as if x and y commuted, v(F̃G) = v(F̃ ) + v(G̃). We use the same reasoning to show v(F +G) ≥ min{v(F ), v(G)}. It follows from the construction that for each i, j, v(Fi,j) is a linear combination of {v(ωi)}i≥−1 and that in case v(Fi,j) = 0, Fi,j ∈ R. Theorem 3.9. Let v be a valuation on A1(R) trivial on R with residue field R. Then v is strongly abelian. Proof. If v’s value group is Q, then the theorem follows from Corollary 2.4. Otherwise, v(ωN ) ̸∈ Q for some N by our construction. But as we have shown in Lemma 3.2, v[ωN , x] = −v(yω1 · · ·ωN−1). If v(ωNx) = v[ωN , x], it follows that v(ωN ) ∈ Q, a contradiction. Since the value group is generated by {v(ωi)}i≥−1, it follows from Propo- sition 2.7 that v is strongly abelian. 4 Valuations on R[y; δ] In this section, we explain a construction of valuations on the ring R[y; δ] with R := { ∑ k≥m akx − kn | ak ∈ R,m ∈ Z, n ∈ N} and δ(p(x)) = p′(x). This construction, which was first introduced in [16], will, as we will see in this section, give us all valuations on R[y; δ] with residue field R. Then, we will prove exactly which valuations on A1(R) with residue field R extend to a valuation on R[y; δ] with the same residue field, answering the question posed by Marshall and Zhang in [16]. We will see the extensions of valuations on R[y; δ] are strongly abelian. Every valuation on R[y; δ] can be uniquely extended to its quotient ring, which we label as D, because R[y; δ] is an Ore domain. Since [y, x] = 1 as before, v(xy) < 0 must hold. We set v(x) = −1, z0 := y and consider v(y). If v(y) ̸∈ Q, then v( n∑ i=0 pi(x)y i) = min 0≤i≤n {v(pi(x)) + iv(y)} L. Vukšić: Valuations and orderings on the real Weyl algebra 253 for any ∑n i=0 pi(x)y i ∈ R[y; δ]. Otherwise v(y) = r1 ∈ Q and hence v(y − γ1x−r1) > v(y) for some γ1 ∈ R. If v(z1) = r2 ∈ Q for z1 := y − γ1x−r1 , we proceed to find z2 = z1 − γ2x−r2 such that v(z2) is greater than r2. We repeat this process to construct a sequence (zi)i≥0. If v(zk) ̸∈ Q for some k ∈ N, then we can write every f ∈ R[y; δ] as ∑n i=0 pi(x)z i k and deduce v(f) = min 0≤i≤n {v(pi(x)) + iv(zk)}. The value group is then group-isomorphic to Q×Z. Since v[x, zk] = v[x, y] = 0 > v(xzk), v is strongly abelian by Proposition 2.7. Otherwise, the sequence (zi)i≥0, v(zi) = ri+1 ∈ Q is infinite. We take note of the fact that v(zi+1) > v(zi) and since [zi, x] = [y, x] = 1 for all i, v(zi) < 1 for all i. We define r := limi→∞ ri ≤ 1. 4.1 Case r < 1 If r < 1, it has been shown in [16] that v can be extended to a valuation on R[y; δ] with residue field R. We first extend v from R to R̃ = { ∑ q∈A aqx −q | aq ∈ R, A ⊂ Q is well-ordered} in a natural way, i.e., by defining v( ∑ q∈A aqx −q) = minA for each ∑ q∈A aqx −q ∈ R̃[y; δ]. Then for every f(t) ∈ R[t], define v(f(y)) = v(f(z)) with z := ∑∞ i≥1 aix −ri and f(y) = ∑n i=0 piy i for f(t) = ∑n i=0 pit i. This gives rise to a valuation on R[y; δ]. However if r = 1, we cannot define a valuation in this way. Let k ∈ N be such that 2rk > 1 + r1, which exists since r = 1, and ak = y − zk = ∑k i=1 γix −ri . Let f(t) = (t− ak)(t− ak) = t2 − 2tak + a2k ∈ R[t]. On one hand, v(f(y)) = v(f(z)) = 2v(z − ak) = 2rk+1. On the other, 2rk+1 = v((y−ak)(y−ak)) = v(y2−2yak+a2k+[y, ak]) = min{2rk+1, 1+r1} = 1+r1, contradicting the assumption that v is a valuation, as shown in [16]. Of course, even in case r < 1, there may also exist a k such that 2rk > 1 + r1. But the important difference between the two cases is that if r < 1, there is always an ℓ ∈ N such that 1 + rℓ > 2rk for all k ∈ N, which does not hold in case r = 1. Then, since f : R[y; δ] → R[y; δ], f(y) = zℓ, f(a) = a for a ∈ R is a real algebra automorphism of R[y; δ], we can translate the sequence by replacing y with zℓ. We see that since the associated value group is Q, v is a strongly abelian valuation by Corollary 2.4. 254 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 4.2 Case r = 1 The question whether in case r = 1, v can be extended from a sequence (zi)i≥0 to a valuation on R[y; δ] was left open in [16]. In this subsection, we show that it can be done using model theory (for reference, see for example [19]). We also show that the valuation we get in this way is uniquely determined. Suppose we have infinite sequences (zi)i≥0 ⊆ R[y; δ], (ri)i≥1 ⊆ Q and (γi)i≥1 ⊆ R and v : (zi)i≥0 → Q with z0 = y, zi+1 = zi − γi+1x−ri+1 and v(zi) = ri+1 ∈ Q with (ri)i≥1 a strictly increasing sequence with r = limi→∞ ri = 1. Then for each n ≥ 0, there is a valuation vn on R[y; δ] such that vn(zi) = ri+1 for all 0 ≤ i ≤ n− 1 and vn(zn) ̸∈ Q. We now present the first-order theory that the valuation associated to the infinite se- quence we wish to prove exists is a model of. The theory will be a union of the theory of D, the quotient division ring of R[y; δ] and the theory of valuations. We will see that each finite subset of this theory has a model. By compactness, so does the whole theory. The language of our theory will be F ∪ {+,−, ·,−1 , O,<} ∪ {czi | i ≥ −1} where F is the set of all constants ca for each a ∈ D, +, · and < are binary function symbols, − and −1 are unary function symbols, O is an unary relation symbol and czi is a constant for all i ≥ 0. Let A be the theory of the quotient division ring of the ring R[y; δ]. By B we will denote the set of axioms for valuation rings O on division rings: 1. O(0) ∧O(1) 2. ∀a : O(a) ∨O(a−1) 3. ∀a, b : O(a) ∧O(b) → O(a+ b) ∧O(ab) ∧O(ba) We add all sentences C that will give proper meaning to the constants czi for all i ≥ −1: 4. cz0 = cy 5. czi+1 = czi − cγi+1x−ri+1 6. O(cxri+1zi) ∧O(c(xri+1zi)−1) Our theory is then the union of all the above axioms from A to C. Since all finite subsets of the theory have a model, namely, the valuation vn described in the beginning of this subsection, so does, by compactness, the whole theory. Since the theory contains F , the set of all constants ca for each a ∈ D, the models are valued division rings which all contain D. We pick a model of the theory, a pair (D1, v), where D1 is a division ring with valuation v. We now show that the v-value is uniquely determined for every f ∈ R[y; δ]. It will then follow that v is uniquely determined on the whole quotient ring D. Every f ∈ R[y; δ] can be written as f = n∑ i=0 p (0) i (x)y i, with p(0)i (x) ∈ R for each 0 ≤ i ≤ n. L. Vukšić: Valuations and orderings on the real Weyl algebra 255 For the time being, we ignore the terms we get when we change the order of multipli- cation. At the end of this subsection, we will see that they do not influence v(f). For each k ≥ 1, we define ak := y − zk = k∑ i=1 γix −ri and write f = n∑ i=0 p (0) i (x)y i = n∑ i=0 p (0) i (x)(ak + zk) i ∼ n∑ i=0 p (0) i (x) i∑ j=0 ( i j ) ai−jk z j k = n∑ j=0 p (k) j (x)z j k with p (k) j (x) := n∑ i=j ( i j ) p (0) i (x)a i−j k for each 0 ≤ j ≤ n and k ≥ 1. For each 0 ≤ j ≤ n, gj(t) := ∑n i=j ( i j ) p (0) i (x)t i is a polynomial in R[t]. The quotient field of C := { ∑ k≥m akx − kn | ak ∈ C,m ∈ Z, n ∈ N} is the algebraic closure of the quotient field of R, as shown in for example [14]. Since∑∞ i=1 γix −ri is not in the quotient field of C, it is not a root of gj(t) for any 0 ≤ j ≤ n. We conclude that for some K ∈ N, v(p(k)j (x)) = v(p (K) j (x)) for all k ≥ K and all 0 ≤ j ≤ n. We can then write f = n∑ i=0 p (0) i (x)y i = n∑ i=0 p (K) i (x)z i K = n∑ i=0 p (k) i (x)z i k with v(p(k)i (x)) = v(p (K) i (x)) for all k ≥ K. We now show that from some K ′ ≥ K, v(p(k)0 ) < v(p (k) i z i k) for all 1 ≤ i ≤ n and all k > K ′. For all k > K, p (k+1) 0 (x) = n∑ i=0 p (k) i (x)(ak+1 − ak) i = n∑ i=0 p (k) i (x)(γk+1x −rk+1)i with v(p(k)i (x)) = v(p (K) i (x)). Since (ri)i≥1 is an increasing sequence with limi→∞ ri = 1, there exists K ′ ≥ K such that min i=0,...,n {v(p(K ′) i (x)) + irK′+1} = min i=0,...,n {v(p(K)i (x)) + irK′+1} is achieved at exactly one 0 ≤ i ≤ n. We conclude v(p(K ′+1) 0 ) = v(p (K) 0 ). Then for each 1 ≤ i ≤ n, v(p (k) i (x)z i k) = v(p (k) i (x))+irk+1 = v(p (K) i (x))+irk+1 > v(p (K) i (x))+irk ≥ v(p (K) 0 ). 256 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 To show that v(f) is equal to v(p(K)0 (x)) ∈ Q, we must show that the v-value of all the terms we get when we change the order of multiplication must be strictly greater than v(p (K) 0 (x)). For all k ≥ 1, we write p (k) 0 (x) = n∑ i=0 p (0) i (x)a i k = n∑ i=0 p (0) i (x)( k∑ j=1 γjx −rj )i = n∑ i=0 ki∑ j=1 p (0) i (x)αix −qj , with qj ∈ Q for all 1 ≤ j ≤ ki and 1 ≤ i ≤ n. Since for all 0 ≤ i ≤ n, p(0)i (x) ∈ R and (ri)i≥1 is an increasing sequence with limi→∞ ri = 1, there exists some k ≥ 1 such that the term of the sum with v-value min i=1,...,n {v(p(0)i (x)) + (i− 1)r1 + rk} is the only term in the sum with its v-value. We conclude v(p (K) 0 (x)) = v(p (k) 0 (x)) ≤ v(p (0) i (x)) + (i− 1)r1 + rk (4.1) for all 1 ≤ i ≤ n. On the other hand, we can write f = n∑ i=0 p (0) i (x)y i = n∑ i=0 p (0) i (x)(ak + zk) i. For all 0 ≤ i ≤ n, all terms of (ak + zk)i when expaned are of the form aℓ1k z ℓ2 k . . . a ℓi−1 k z ℓi k with ℓ1, . . . , ℓi ≥ 0 and ℓ1 + · · ·+ ℓi = i. Since v(ak) = r1, v(zk) = rk+1 and v[ak, zk] = v[ k∑ j=1 γjx −rj , y − k∑ j=1 γjx −rj ] = v( k∑ j=1 γj [x −rj , y]) = 1 + r1, it follows that the v-value of each term we get when we change the order of multiplication is at least v(p(0)i (x)) + (i − 1)r1 + 1 for each 1 ≤ i ≤ n, which is, as is immediate from (4.1), strictly greater than v(p(K)0 (x)). We thus conclude v(f) = v(p (K) 0 (x)) ∈ Q. It also follows that v(f) = vk(f) for all k ≥ K ′ with vk as defined in the beginning of this section. As every element of D, the quotient ring for R[y; δ], can be written as fg−1 with f, g ∈ R[y; δ], it follows that v is uniquely determined on D. We see that the value group for v is equal to Q. We conclude from Corollary 2.4 that v is strongly abelian. It remains to show that the residue field for v is equal to R. Suppose v(f) = 0 for some f ∈ D. Then vk(f) = 0 for all k ≥ K for some K ∈ N. We can write f = ( m∑ i=0 p (K) i z i K)( n∑ j=0 q (K) j z j K) −1 with p(K)i , q (K) j ∈ R and v( m∑ i=0 p (K) i z i K) = v(p (K) 0 ) = v( n∑ j=0 q (K) j z j K) = v(q (K) 0 ) = q ∈ Q. It follows that v(xq ∑m i=0 p (K) i z i K) = v(x q ∑n j=0 q (K) j z j K) = 0 and α := xq ∑m i=0 p (K) i z i K = xqp (K) 0 ∈ R = R, β := xq ∑n j=0 q (K) j z j K = x qq (K) 0 ∈ R = R. We conclude f = αβ−1 ∈ R. So the residue field for v is indeed equal to R. L. Vukšić: Valuations and orderings on the real Weyl algebra 257 4.3 Extensions of valuations from A1(R) to R[x; δ] In this section, we characterize the valuations on A1(R) with residue field R that have an extension to R[y; δ] with the same residue field. Since x mi ni ∈ R[y; δ], it follows that any valuation v′ that extends v to a valuation on R[y; δ] with residue field R must satisfy v′(x mi ni ωi−1) = 0 and γ̃i := x mi ni ωi−1 ∈ R. In the next proposition, we show the necessary condition for a valuation v on A1(R) to have an extension v′ to R[y; δ] with the same residue field. Proposition 4.1. Let v be a valuation on A1(R) with residue field R associated to a se- quence (ωi)i≥−1 with v(ωi−1) = mini for i ≥ 1 and x miωnii−1 = βi ∈ R. Let αi,j ∈ R be as in Section 3. Let 2hi be the greatest power of two dividing ni for all i ≥ 0. Then v can be extended to a valuation on R[y; δ] with the same residue field only if it fulfils the following conditions: (1) For each i such that ni is even, βi > 0 must hold, and (2) for each i, j, ℓ with hi < hj ≤ hℓ, αi,jαi,ℓ > 0 must hold. Proof. Since γ̃i ni = (x mi ni ωi−1)ni = xmiω ni i−1 = βi due to Proposition 2.2, it is obvious that γ̃i must be equal to an ni-th root of βi. If ni is odd, γ̃i ∈ R is uniquely determined regardless of sgn(βi), while if ni is even, γ̃i ∈ R only if βi > 0. It is thus obvious that βi > 0 must hold for all i where ni is even if v can be extended from a valuation on A1(R) to a valuation on R[y; δ] with the same residue field. This proves the necessity of the first condition. To prove the necessity of the second condition, we first observe that αi,j = ω Ki,j i−1 ω −Kj,i j−1 = γ̃i Ki,j γ̃j −Kj,i , holds for all i, j ≥ 0. If hi < hj , Ki,j is odd while Kj,i is even, so sgn(γi) = sgn(αi,j) must hold. We can therefore see that αi,jαi,ℓ = γ̃i Ki,j+Ki,ℓ γ̃j −Kj,i γ̃ℓ −Kℓ,i for all i, j, ℓ ≥ 0. If hi < hj ≤ hℓ, both Ki,j and Ki,ℓ are odd while Kj,i and Kℓ,i are even. It follows that if v can be extended to a valuation on R[y; δ] with residue field R, αi,jαi,ℓ > 0 must hold for all i, j, ℓ ≥ 0 with hi < hj ≤ hℓ. In this section, we show that the conditions (1) and (2) of Proposition 4.1 are also sufficient for v to have an extension to R[y; δ] with residue field R. Let v be any valuation on A1(R) satisfying the conditions described in Proposition 4.1. We will first determine γ̃i ∈ R for all i ≥ 0. If ni is odd, there is a unique choice of γ̃i ∈ R. Suppose then ni is even and βi > 0 for some i ≥ 0. If hi < hj for some j, then sgn(γ̃i) = sgn(αi,j) = sgn(γ̃i Ki,j γ̃j −Kj,i) since Ki,j is even while Kj,i is odd. We can conclude that if for every power of two 2h there is an i ≥ 0 (or, equivalently, if the v-value group is 2-divisible), γ̃i ∈ R is uniquely determined for all i ≥ 0. If on the other hand, the value group is non-2-divisible, there is an i ≥ 0 such that 2hi is maximal 258 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 for all i ≥ 0. We then have two choices for γ̃i - a positive or a negative one. The sign of γ̃j is then uniquely determined for all j ≥ 0 since αi,j = γ̃i Ki,j γ̃j −Kj,i , where Ki,j is even and Kj,i is odd. We will now take an arbitrary valuation v on A1(R) satisfying the conditions of Proposition 4.1, constructed by a sequence of (ωi)i≥−1 as shown in Section 3. We also pick γ̃i ∈ R for all i. Then we will describe v’s extension to R[y; δ] with a sequence of (zi)i≥0 like in the beginning of Section 4. Suppose the valuation v on A1(R) is given by a sequence (ωi)i≥−1 with ω−1 = x, ω0 = y and ωi = xmiωnii−1 − βi, βi ∈ R and gcd(mi, ni) = 1 for all i. Suppose also that v satisfies conditions (1) and (2) of Proposition 4.1. We will show that there is exactly one extension of v from A1(R) to R[y; δ] for each appropriate choice of (γ̃i)i≥1. Lemma 4.2. For each k ≥ ℓ ≥ 0, ωk can be written in the following form: ωk = (Π k i=ℓ+1x mi ni )ωℓ(Π k i=ℓ+1Bi)− k∑ i=ℓ+1 (Πkj=i+1x mj nj )γ̃i(Π k j=iBj) + k∑ i=ℓ+1 (Πkj=i+1x mj nj )Ai(Π k j=i+1Bi), with Ai = ni−1∑ j=1 (x mi ni ωi−1) ni−j−1x mi ni [x j mi ni , ωi−1]ω j i−1 Bi = nj∑ j=1 (x mi ni ωi−1) ni−j γ̃i j−1 for each 1 ≤ i ≤ k. Proof. We prove the lemma by induction on k ≥ ℓ. For k = ℓ, it is trivially true since we get ωk = ωk. Suppose now the equation holds for some k ≥ ℓ. Then ωk+1 = x mk+1ω nk+1 k − βk+1 = (x mk+1 nk+1 ωk − γ̃k+1)Bk+1 +Ak+1 = x mk+1 nk+1 ((Πki=ℓ+1x mi ni )ωℓ(Π k i=ℓ+1Bi)− k∑ i=ℓ+1 (Πkj=i+1x mj nj )γ̃i(Π k j=iBj) + k∑ i=ℓ+1 (Πkj=i+1x mj nj )Ai(Π k j=i+1Bi))Bk+1 − γ̃k+1Bk+1 +Ak+1 = (Πk+1i=ℓ+1x mi ni )ωℓ(Π k+1 i=ℓ+1Bi)− k+1∑ i=ℓ+1 (Πk+1j=i+1x mj nj )γ̃i(Π k+1 j=i Bj) + k+1∑ i=ℓ+1 (Πk+1j=i+1x mj nj )Ai(Π k+1 j=i+1Bi), L. Vukšić: Valuations and orderings on the real Weyl algebra 259 where we used the induction hypothesis, which is ωk = (Π k i=ℓ+1x mi ni )ωℓ(Π k i=ℓ+1Bi)− k∑ i=ℓ+1 (Πkj=i+1x mj nj )γ̃i(Π k j=iBj) + k∑ i=ℓ+1 (Πkj=i+1x mj nj )Ai(Π k j=i+1Bi) in the second equation. Since v(ωi−1) = mini and x mi ni ωi−1 = γ̃i, v(Bi) = 0 and Bi = niγ̃ini−1 must hold for all i ≥ 1. From [xmi , ωi−1] = ni−1∑ j=0 x mi ni ·j [x mi ni , ωi−1]x mi ni ·(ni−j), we can see that v(Ai) = v[x mi ni , ωi−1] = v[x, ωi−1]+1− mini = −v(xyω1 . . . ωi−1). Since∑k i≥−1 v(ωi) < 0 for every k by Lemma 3.2, v( k∑ i=ℓ+1 (Πj=i+1x mj nj )Ai(Π k j=iBi)) > v(ωk) will hold for every k, which is why we can ignore the terms containing Ai during our evaluations of v(zi) where for each i, zi ∈ R[y; δ] is as in the beginning of this section. Lemma 4.3. Suppose v is a valuation on A1(R) constructed from a sequence (ωi)i≥−1 that extends to a valuation on R[y; δ] and thus D, its quotient division ring. For a given i ≥ 1, define a sequence (Si,j)j≥1 by: (a) Si,1 := (x mi ni ωi−1 − γ̃i)−1(Bi −Bi), (b) Si,j+1 := (x mi ni ωi−1 − γ̃i)−1(Si,j − Si,j) for j ≥ 1. Then for each j ≥ 1: 1 Si,j = ∑ni−j k=1 Nk,j(x mi ni ωi−1) ni−j−kγ̃i k−1 with Nk,1 = k and Nk,j+1 =∑ni−j ℓ=k Nℓ,j , and 2 v(Si,j) = 0, Si,j = N1,j+1γ̃ini−j−1 for all 1 ≤ j ≤ ni − 1. Proof. We prove the first statement of the lemma by induction on j ≥ 1. To show the basis of induction, we evaluate Bi −Bi = (x mi ni ωi−1 − γ̃i)( ni∑ k=1 ((x mi ni ωi−1) ni−k−1+ (x mi ni ωi−1) ni−k−2γ̃i + · · ·+ γ̃ini−k−1)γ̃ik−1) = (x mi ni ωi−1 − γ̃i) ni−1∑ k=1 k(x mi ni ωi−1) ni−k−1γ̃i k−1. 260 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 We can thus see Si,1 = ∑ni−1 k=1 k(x mi ni ωi−1) ni−k−1γ̃i k−1 and since x mi ni ωi−1 = γ̃i, we see that the lemma holds in case j = 1. Now we suppose that the statement is true for some j ≥ 1, i.e., Si,j = ∑ni−j k=1 Nk,j(x mi ni ωi−1) ni−j−kγ̃i k−1 and v(Si,j) = 0, Si,j = N1,j+1γ̃i ni−j−1. We then write Si,j − Si,j = ni−j∑ k=1 Nk,j(x mi ni ωi−1) ni−j−kγ̃i k−1 −N1,j+1γ̃ini−j−1 = ni−j∑ k=1 Nk,j(x mi ni ωi−1) ni−j−kγ̃i k−1 − ni−j∑ k=1 Nk,j γ̃i ni−j−1 = ni−j−1∑ k=1 Nk,j((x mi ni ωi−1) ni−j−k − γ̃ini−j−1)γ̃ik−1 = (x mi ni ωi−1 − γ̃i) ni−j−1∑ k=1 Nk,j((x mi ni ωi−1) ni−j−1−k +· · ·+ γ̃ini−j−1−k)γ̃ik−1 = (x mi ni ωi−1 − γ̃i) ni−j−1∑ k=1 (Nk,j +Nk+1,j + · · ·+Nni−j−1,j) (x mi ni ωi−1) ni−j−kγ̃i k−1 = (x mi ni ωi−1 − γ̃i) ni−j−1∑ k=1 Nk,j+1(x mi ni ωi−1) ni−j−kγ̃i k−1 proving the first statement of the lemma. The second statement immediately follows from the first since x mi ni ωi−1 = γ̃i and thus v(Si,j) = v( ni−j∑ k=1 Nk,j(x mi ni ωi−1) ni−j−kγ̃i k−1) = 0, Si,j = ni−j∑ k=1 Nk,j(x mi ni ωi−1)ni−j−kγ̃i k−1 = N1,j+1γ̃i ni−j−1. Lemma 4.4. Supose v is as in Lemma 4.3. For a given i ≥ 1, define a sequence (Di,j)j≥1 by: (a) Di,1 = B1B2 · · ·Bi −B1B2 · · ·Bi, (b) Di,j+1 = xv(Di,j)Di,j − xv(Di,j)Di,j . Then for each j ≥ 1: 1. Di,j is a R-linear sum of terms which are products of elements from the set {ωi}i ∪ {Bi}i ∪ {B−1i }i ∪ {Si,j}i,j , (4.2) where parts of the product are conjugated by a rational power of x. L. Vukšić: Valuations and orderings on the real Weyl algebra 261 2. v(Di,j) is a sum of v(ωℓ) for finitely many ωℓ. 3. xv(Di,j)Di,j is sum of products of γ̃k for various k. Proof. Since B1B2 · · ·Bi −B1B2 · · ·Bi = i∑ j=1 B1 · · ·Bj−1(Bj −Bj)Bj+1 · · ·Bi = i∑ j=1 B1 · · ·Bj−1ωjB−1j Sj,1Bi+1 · · ·Bi, and x mj+1 nj+1 B1 · · ·Bj−1ωjB−1j Sj,1Bi+1 · · ·Bi = (x mj+1 nj+1 B1 · · ·Bj−1x − mj+1 nj+1 )x mj+1 nj+1 ωjB −1 j Sj,1Bi+1 · · ·Bi, for each 1 ≤ j ≤ k, the first two statements of the lemma follow from: 1. x mi ni ωi−1 − γ̃i = ωiB−1i , 2. Bi −Bi = (x mi ni ωi−1 − γ̃)Si,1 = ωiB−1i Si,1, 3. Si,j − Si,j = (x mi ni ωi−1 − γ̃Si,j+1 = ωiB−1i Si,j+1, 4. B−1i −Bi −1 = −Bi−1(Bi −Bi)Bi −1 = −Bi−1ωiB−1i Si,1Bi −1 , where we ignore the terms we get when we change the order of multiplication. We can do that that since these terms are procucts of Ai as defined in Lemma 4.2 and terms with zero v-value. As we have already mentioned, these terms will not influence the construction of the extension of a valuation on A1(R) to R[y; δ]. Indeed - we can see by induction on j ≥ 1 that each term of the sum is a product of factors equal to, modulo conjugation by a rational power of x, one of the elements of the set (4.2); that is, 1. either equal to ωi, or 2. equal to a power of Bj or Sj,i for some i, j. Since the latter have v-value equal to zero and since both Bi − Bi and Si,j − Si,j are products of ωi, a power of B−1i and Si,j for some i, j, v(Di,j) is a sum of v(ωi) for some i. The last statement of the lemma follows from the fact that for all i, j, Bi and Si,j are of the form Nγ̃i for N ∈ N. If v is a valuation on R[y; δ] with residue field R, then, as we have presented in Sec- tion 4, v can be constructed from a sequence (zi)i≥0 ⊂ R[y; δ]. In the next proposition, we make the first comparison between this construction and the construction of a valuation on A1(R) from a sequence (ωi)i≥−1 described in Section 3. 262 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 Lemma 4.5. Suppose v is as in Lemma 4.3. Suppose that for some k, ℓ ≥ 0, we can write ωk = (Π k i=1x mi ni )zℓ(Π k i=1Bi) + C, where C is a R-linear sum of elements of the form Di,j for some i, j ≥ 0. Then: 1. If v(ωk) > v(Πki=1x mi ni zℓ) ∈ Q, then ωk = (Π k i=1x mi ni )zℓ+1(Π k i=1Bi) + C1. 2. If v(ωk) = v(Πki=1x mi ni zℓ) ∈ Q, then ωk+1 = (Π k+1 i=1 x mi ni )zℓ+1(Π k+1 i=1Bi) + C2. 3. If v(ωk) < v(Πki=1x mi ni zℓ), then, if v(ωk) ∈ Q, ωk+1 = (Π k+1 i=1 x mi ni )zℓ(Π k+1 i=1Bi) + C3. Here, C1, C2 and C3 are other R-linear sums of elements of the form Di,j as in Lemma 4.5 for some i, j ≥ 0. Proof. Suppose first v(ωk) > v(Πki=1x mi ni zℓ) = v(C). Since v(C) = rℓ+1 − ∑k i=1 mi ni ∈ Q, then rℓ+1 := v(zℓ) ∈ Q as well. So, for zℓ+1 = zℓ − x−rℓ+1γℓ+1, we can write ωk = (Π k i=1x mi ni )zℓ+1(Π k i=1Bi) + γℓ+1(Π k i=1x mi ni )x−rℓ+1(Πki=1Bi) + C = (Πki=1x mi ni )zℓ+1(Π k i=1Bi) + x −v(C)(γℓ+1(Π k i=1Bi) + x v(C)C) Since v(zℓ+1) > v(zℓ), we see that v(γℓ+1Πki=1Bi + x v(C)C) > 0. It follows that γi+1 = −Πki=1Bi −1 x−v(C)C, hence C1 := γℓ+1Π k i=1Bi + x −v(C)C = γℓ+1Π k i=1Bi + x −v(C)C − x−v(C)C + x−v(C)C = γℓ+1(Π k i=1Bi −Πki=1Bi) + (x−v(C)C − x−v(C)C) We see that since C is an R-linear sum of Di,j for various i, j, so is, by Lemma 4.4, x−v(C)C − x−v(C)C. Hence, C1 is an R-linear sum of Di,j . Now consider the case v(ωk) = v(Πki=1x mi ni zℓ) ≤ v(C). Since v(ωk) = mk+1nk+1 ∈ Q, we can evaluate ˜γk+1 = x mk+1 nk+1 ωk = (Π k+1 i=1 x mi ni )zℓ(Πki=1Bi) + x mk+1 nk+1 C = γℓ+1Π k i=1Bi + x mk+1 nk+1 C, L. Vukšić: Valuations and orderings on the real Weyl algebra 263 and then deduce ωk+1 = (x mk+1 nk+1 ωk − ˜γk+1)Bk+1 = (Πk+1i=1 x mi ni )zℓ(Π k+1 i=1Bi) + x mk+1 nk+1 CBk+1 − ˜γk+1Bk+1 = (Πk+1i=1 x mi ni )zℓ+1(Π k+1 i=1Bi) + γℓ+1(Π k+1 i=1Bi) + x mk+1 nk+1 CBk+1 − ˜γk+1Bk+1 = (Πk+1i=1 x mi ni )zℓ+1(Π k+1 i=1Bi) + γℓ+1(Π k i=1Bi −Πki=1Bi)Bk+1+ (x mk+1 nk+1 C − x mk+1 nk+1 C)Bk+1 = (Πk+1i=1 x mi ni )zℓ+1(Π k+1 i=1Bi) + C2 where C2 is, again, an R-linear sum of Di,j for some i, j. Lastly, we consider the case v(Πki=1x mi ni zℓ) > v(ωk) = v(C) ∈ Q. In this case, ˜γk+1 = x mk+1 nk+1 ωk = x mk+1 nk+1 C, so we can write ωk+1 = (x mk+1 nk+1 ωk − ˜γk+1)Bk+1 = (Πk+1i=1 x mi ni )zℓ(Π k+1 i=1Bi) + (x mk+1 nk+1 C − x mk+1 nk+1 C)Bk+1 and the statement again follows. Theorem 4.6. Suppose v is a valuation on A1(R), constructed from an either finite or infinite sequence (ωi)i≥−1 with v(ωi) = mi+1 ni+1 . Suppose also that v satisfies the following conditions: 1. For each i such that ni is even, βi > 0 must hold, and 2. for each i, j, ℓ ≥ 0 with hi < hj ≤ hℓ, αi,jαi,ℓ > 0 must hold. Then v has a unique extension to a valuation on R[y; δ] with residue field R for each choice of (γ̃i)i≥1. Proof. As we know from the beginning of Section 4, each valuation on R[y; δ] and D with residue field R can be constructed by either a finite or an infinite sequence (zi)i≥0. For every sequence (ωi)i≥−1, we will use the lemmas proved in this section to find the unique sequence (zi)i≥0 which, as we have shown in the beginning of this section, uniquely determines a valuation on R[y; δ]. Our calculations will then show that the valuation on D defined by the sequence (zi)i≥0 is the extension of the valuation on A1(R) associated to the sequence (ωi)i≥−1. We determine the finite or infinite sequence (zi)i≥0 associated to v’s extension to R[y; δ]. In the first step, we consider v(y). If v(y) ̸∈ Q, then v’s extension to R[y; δ] is clearly uniquely determined, namely the one defined by v( n∑ i=0 pi(x)y i) = min 0≤i≤n {v(pi(x)) + iv(y)} for every ∑n i=0 pi(x)y i ∈ R[y; δ]. 264 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 So suppose v(y) = m1n1 ∈ Q. Then, in the second step of our evaluation, we write ω1 = x m1yn1 − β1 = (x m1 n1 y − γ̃1)B1 = x m1 n1 (y − x− m1 n1 γ̃1)B1 and since v(ω1) > 0, v(y − x− m1 n1 γ̃1) > m1 n1 = v(y). We deduce z1 = y − x− m1 n1 γ1 with γ1 = γ̃1. Obviously, v(z1) = v(ω1) + m1n1 ∈ Q if and only if v(ω1) ∈ Q. If either and hence both values are irrational, we get a unique extension of v to R[y; δ]. Otherwise, if v(ω1) = m2 n2 ∈ Q and hence r2 = v(z2) = m1n1 + m2 n2 , we continue with the third step of our evaluation by writing ω2 = (x m2 n2 ω1 − γ̃2)B2 = (x m2 n2 x m1 n1 z1B1 − γ̃2)B2. Since v(ω2) > 0, we conclude that γ2 = x m2 n2 x m1 n1 z1 must be equal to γ̃2B1 −1 . To evaluate the v-value of z2 = z1 − γ2x−r2 , we write ω2 = (x m2 n2 x m1 n1 z2B1 + γ2B1 − γ̃2)B2 = x m2 n2 x m1 n1 z2B1B2 + (γ2B1 − γ̃2)B2. We note that ω2 is here written as a sum of Π2i=1x mi ni z2Π 2 i=1Bi + C where C is as in Lemma 4.5. To determine v(z2), we compare v(ω2) and v(γ2B1 − γ̃2), the latter being equal to v(ω1) since γ2B1 − γ̃2 = γ2(B1 −B1). There are three possible cases: 1. If v(ω2) < v(ω1), then v(x m2 n2 x m1 n1 z2) must be equal to v(ω2), so r3 = v(z2) is determined by v(z2) = v(ω2) − m1n1 − m2 n2 . If v(ω2) = m3n3 ∈ Q, then v(z2) = r3 ∈ Q and γ3 = γ̃3(B1B2)−1 must hold. It follows that v(z2) ∈ Q if and only if v(ω2) ∈ Q. In this case, the v-value of z3 = z2 − γ3x−r3 will be determined in the subsequent steps, i.e., by considering (v(ωi))i≥3 and (γ̃i)i≥3. By Lemma 4.5, ω3 = Π 3 i=1x mi ni z3Π 3 i=1Bi + C1, C1 being an R-linear sum of Di,j . 2. If v(ω2) = v(ω1), then v(z2) depends on γ̃3: (a) If γ̃3 = x m2 n2 (γ2B1 − γ̃2)B2, then v(x m2 n2 x m1 n1 z2) must be greater than v(ω1) since in this case, ω2 ∼ (γ2B1 − γ̃2)B2 and will be determined in the sub- sequent steps, i.e., by considering (v(ωi))i≥3 and (γ̃i)i≥3. By Lemma 4.5, ω3 = Π 3 i=1x mi ni z2Π 3 i=1Bi + C2, C2 being an R-linear sum of Di,j . (b) Otherwise, v(x m2 n2 x m1 n1 z2) = v(ω2). In this case, we get γ3 = γ̃3 − x m2 n2 (γ2B1 − γ̃2)B2. By Lemma 4.5, ω3 = Π3i=1x mi ni z3Π 3 i=1Bi + C3, with C3 an R-linear sum of Di,j . 3. If v(ω2) > v(ω1), then v(x m2 n2 x m1 n1 z2) = v(ω1) and γ3 = −x m2 n2 (γ2B1 − γ̃2)B−11 . By Lemma 4.5, ω2 = Π2i=1x mi ni z3Π 2 i=1Bi + C4, with C4 an R-linear sum of Di,j . L. Vukšić: Valuations and orderings on the real Weyl algebra 265 The general step of the evaluation is similar to the first three. Suppose that in the previous steps, we have evaluated v(z1) = r2, . . . , v(zℓ−1) = rℓ ∈ Q. In the last step, we have, by considering (ωi)ki=−1 for some k, begun to evaluate v(zℓ) and we are, with ωk = (Π k i=1x mi ni )zℓ+1(Π k i=1Bi) + C, where C is as in Lemma 4.5, in one of the five situations: 1. If v(ωk) < v(C), then v(zℓ) = v(ωk) − ∑k i=1 mi ni and γℓ+1 = ˜γℓ+1Πki=1Bi −1 . In case rℓ+1 = v(zℓ) ∈ Q, our next step is to evaluate the v-value of zℓ+1 = zℓ − rℓ+1γℓ+1 by writing ωk+1 = (Π k+1 i=1 x mi ni )zℓ(Π k+1 i=1Bi) + C1. 2. If v(ωk) > v(C), then v(zℓ) = v(C) and γℓ+1 = CΠki=1Bi −1 . In case rℓ+1 = v(zℓ) ∈ Q, our next step is to evaluate the v-value of zℓ+1 = zℓ − rℓ+1γℓ+1 by writing ωk = (Π k i=1x mi ni )zℓ+1(Π k i=1Bi) + C2. 3. If v(ωk) = v(C) ̸∈ Q, then v(zℓ) = v(ωk − C) − ∑k i=1 mi ni ̸∈ Q since in case v(ωk − C) > v(ωk), C ∼ ωk must hold, but since, given that C ̸= ωk and that C is a sum of Di,j , v(ωk − C) must be in Q. This terminates our evaluation of the sequence (zi)i≥0 associated to v’s extension to R[y; δ]. 4. If v(ωk) = v(C) = mk+1 nk+1 ∈ Q and x mk+1 nk+1 ωk = x mk+1 nk+1 C, v((Πki=1x mi ni )zℓ) > v(ωk). We continue with our evaluation by writing ωk+1 = (Π k+1 i=1 x mi ni )zℓ(Π k+1 i=1Bi) + C3. 5. If v(ωk) = v(C) = mk+1 nk+1 ∈ Q and x mk+1 nk+1 ωk ̸= x mk+1 nk+1 C, then v((Πki=1x mi ni )zℓ) = v(ωk) and γℓ+1 = (x mk+1 nk+1 ωk − x mk+1 nk+1 C)Πki=1Bi −1 . We write ωk+1 = (Π k+1 i=1 x mi ni )zℓ+1(Π k+1 i=1Bi) + C4. For each 1 ≤ i ≤ 4, Ci is, as is C, an R-linear sum of Di,j for various i, j. This is assured by Lemma 4.5. We point out that for each ℓ, v(zℓ) is determined in a finite number of steps. In case v is determined by a finite sequence (ωi)Ni≥−1 for some N ≥ 0, this is immediate. In the infinite case, it follows from Lemma 3.2 that limk→∞ v(ωk) = 0, so, given that by Lemma 4.4, v(C) is a sum of v(ωi), v(ωk) < v(C) must hold for some k. And since, in the step where v(zℓ) is determined, we get v(Πki=1x mi ni zℓ) = v(zℓ) − ∑k i=1 v(ωi−1) ≤ v(ωk) for each ℓ ≥ 0, we get v(zℓ) ≤ ∑k i=0 v(ωi) < 1 by Lemma 3.2. Since for each ℓ ≥ 0, zℓ+1 = zℓ − x−rℓ+1γℓ+1, we did indeed find a unique sequence (zi)i≥0 that uniquely determines a valuation v on R[y; δ]. This valuation is v’s extension from A1(R) to R[y; δ]. The construction introduced in the proof of Theorem 4.6 can be reversed. Given a val- uation v on R[y; δ], we could use the reverse construction to find the sequence (ωi)i≥−1 ⊆ A1(R) associated to v’s restriction to A1(R). 266 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 5 Valuations on R̃[y; δ] The ring R̃[y; δ] is an extension of R[y, δ] where R̃ is defined as R((xQ)), the generalized power ring of sums ∑ q∈Q αqx −q with well-ordered support. We first show that every valuation on R[y; δ] can be easily extended to R̃[y; δ]. Lemma 5.1. Every valuation on R[y; δ] with residue field R can be extended to a valuation on R̃[y; δ] with the same residue field. Proof. Suppose first v is defined on R[y; δ] by a finite sequence (zi)ki=0 with v(zk) ̸∈ Q. Then, as we can write every f ∈ R̃[y; δ] as ∑n i=0 pi(x)z i k with pi(x) ∈ R̃, we define v(f) = min1≤i≤n{v(pi(x)) + iv(zk)}. This gives us a well-defined valuation on R̃[y; δ] which clearly extends the one we defined on R[y; δ] in the previous section. Now suppose v is defined by an infinite sequence (zi)i≥0 with r = limi→∞ v(zi) ≤ 1. Define z := y− ∑∞ i=1 αix −ri ∈ R̃[y; δ]. In the same way as before, write every f ∈ R̃[y; δ] as ∑n i=0 pi(x)z i with pi(x) ∈ R̃ and define v(f) = min1≤i≤n{v(pi(x))y + i(r − µ)} where µ is a positive infinitesimal. Thus we once more get v’s extension to R̃[y; δ]. In case r = 1, this is the only possible extension up to isomorphism of the value group, for v(z) must be 1 − µ. This is because on the one hand, since v(z) > v(zk) = rk+1 for all k ≥ 0, v(z) is greater than any rational number q < 1. On the other hand, since v[y− z, x] = v[y, x] = 0 and the value group is commutative, v(z) ≤ 1. Thus if v(z) ∈ R, v(z) = 1. But if we restricted v to the quotient ring of the R-algebra, generated by x and z, we would get a valuation on a division ring, isomorphic to D1(R), with a rational value group, residue field R and v[z, x] = v(zx), which contradicts Corollary 2.4. Proposition 5.2. Let v be a valuation on R̃[y; δ] with residue field R. Then the value group is not of rational rank one. Proof. The only case in the proof of Lemma 5.1 where it does not immediately follow that the value group is not Q is when v’s restriction to R[y; δ] is constructed by an infinite sequence (zi)i≥0 with r := limi→∞ v(y − zi) < 1. In this case, we can set v(z) = r ∈ R and if r ∈ Q, define y(1) := z and restart the construction of v. We may get another infinite sequence (z(1)i )i≥0 with r (1) := limi→∞ v(z (1) i ) < 1. If r (1) ∈ Q, we start over with y(2) := z(1). Though we may have to repeat the process infinitely many times, the set {z(j)k }j≥1,k≥0 is countable and A := {v(z (j) k )}j≥1,k≥0 is a well-ordered set of rational numbers smaller than one. At one point, v(z) will have to be irrational for some z = ∑ q∈A αqx −q since we would otherwise get z ∈ R̃ such that v(z) = 1 which, as we have shown, contradicts the fact that the value group is rational. Recall that a pseudo-Cauchy sequence in a division ring D with a valuation v is a sequence (aλ)λ∈Λ ⊆ D, where Λ is an ordinal such that there exists λ ∈ Λ for which v(aσ − aρ) < v(aρ − aτ ) for all σ, ρ, τ ∈ Λ with λ ≤ σ < ρ < τ . Let D′ be an extension of D and v′ and extension of v to D′. Then a ∈ D′ is a limit of the pseudo-Cauchy sequence (aλ)λ∈Λ if v′(a − aσ) = v′(aσ+1 − aσ) for all σ ∈ Λ, λ ≤ σ. Recall also the definition of a immediate extension of a valuation. L. Vukšić: Valuations and orderings on the real Weyl algebra 267 Definition 5.3. Let v be a valuation on a division ring D, the division ring D′ an extension of D and v′ a valuation on D′ that extends v. Let Γ and D and Γ′ and D′ be the value groups and residue division rnigs associated to v and v′. Then we say v′ is an immediate extension of v if [Γ′ : Γ] = 1 and [D′ : D]. As a byproduct of our investigations, we show that not every extension of a valued division ring by limits of pseudo-Cauchy sequences is immediate. This differs from the commutative case since, as Kaplansky proved in [8], every extension of a valued field by limits of pseudo-Cauchy sequences is immediate. Corollary 5.4. There exist division rings D ⊆ D′ and a valuation v on D which extends to a valuation v′ on D′ such that D′ is an extension of D by limits of pseudo-Cauchy sequences in D whereas v′ is not an immediate extension of v. Proof. Let v be a valuation on R[y; δ] with residue field R and value group Q as described in Section 4. Then v(x) = −1 and R̃[y; δ] is an extension of R[y; δ] by limits of pseudo- Cauchy sequences. This holds because every ∑ i∈Q aix −i ∈ R̃ is a limit of the pseudo- Cauchy sequence ( ∑k i=1 aix −qi)k≥1 in R[y; δ]. As we have shown in this section, v can be uniquely extended to R̃[y; δ]. By Proposi- tion 5.2, this extension is not immediate. Let D be the quotient division ring of R[y, δ], to which v uniquely extends, and D′ the quotient division ring of R̃[y; δ], to which v′ uniquely extends, since both rings are Ore domains. D′ is not an immediate extension of the ring D with valuation v, even though D′ is an extension of D by limits of pseudo-Cauchy sequences. 6 Compatibility with orderings on A1(R) and R[y; δ] In Section 3, we mentioned that every strongly abelian valuation on a division ring with an ordered residue field is compatible with an ordering on the valued division ring. In this section, we will use a noncommutative version of the Baer-Krull theorem to determine all orderings on A1(R) compatible with one of the valuations v we have described in the previous sections. We will then show which of these orderings on A1(R) can be extended to an ordering on R[y; δ] compatible with a v’s extension to R[y; δ]. Recall that an order P on a division ring D is compatible with a valuation v on D if for every a, b ∈ D∗ such that v(a) = v(b) < v(a− b), ab ∈ P holds. Let v be a strongly abelian valuation on a division ring D with a formally real residue field D. Let Γ be its value group. Let s : Γ → D∗ be a semisection of v, i.e., a map for which 1. s(0) = 1, 2. v(s(g)) = g for all g ∈ Γ, 3. s(g1 + g2) = s(g1)s(g2)u2 for some u ∈ D∗ for all g1, g2 ∈ Γ. Let χ : Γ/2Γ → {−1, 1} be a group homomorphism called a character and let P be an ordering of D. Then, as it was shown in [24], Pχ = {a ∈ D | a · s(v(a))−1χ(v(a) + 2Γ) ∈ P} 268 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 is an order of D compatible with v. Moreover, if X denotes all orders of the residue field, Xv denotes all v-compatible orders on D and (Γ/2Γ)∗ denotes the set of all characters on Γ/2Γ, then by Proposition 3 of [24], the map f : X × (Γ/2Γ)∗ → Xv f((P , χ)) = Pχ is a bijection. The choice of a semisection s on Γ does not matter. Using f , we will now describe all orders on A1(R) that are compatible with a valuation described in Section 3. Suppose v is a valuation on A1(R) associated to an infinite sequence (ωi)i≥−1 with R as a residue field. There is only one possible order of R, so the orders of A1(R) compatible with v will only depend on the characters χ : Γ/2Γ → {−1, 1}. Then there are three different options for the value group Γ ⊆ Q× Z. 1. Γ is a 2-divisible subgroup of Q, 2. Γ is a non-2-divisible subgroup of Q, 3. Γ is a direct sum of a non-2-divisible subgroup of Q with Z. Since in each case the value group Γ is generated by {v(ωi)}i≥−1, the characters and thus the v-compatible orders will be determined by the signs of the ωi. In the first two cases, v is determined by an infinite sequence (ωi)i≥−1 with v(ωi) = mi+1 ni+1 ∈ Q for each i ≥ −1. In the third case, v is determined by a finite sequence (ωi)Ni=−1 with v(ωi) = mi+1 ni+1 ∈ Q for each −1 ≤ i ≤ N − 1 and v(ωN ) ̸∈ Q. If the value group is a 2-divisible subgroup of Q, then for each ωi, there is a j ̸= i such that Ki,j ∈ Z is odd while Kj,i ∈ Z is even. Since ω Ki,j i−1 ω Kj,i j−1 = αi,j , it follows that ωi−1 > 0 if and only if αi,j > 0. Conversely, if the value group is a non-2-divisible subgroup of Q, we choose one i such that ni is divisible by the greatest power of two that divides nj for any j ≥ 0. After choosing either ωi < 0 or ωi > 0, the order on A1(R) is defined. In the last case, where the value group is a direct sum of a non-2-divisible subgroup of Q and Z, the order is determined by choosing either ωi < 0 or ωi > 0 and, independently, either ωN > 0 or ωN < 0 where i is as in the second case and v(ωN ) ̸∈ Q. All four combinations define an ordering on A1(R). We have thus proven the following proposition. Proposition 6.1. Suppose v is a valuation on A1(R) with residue field R and value group Γ. Then: 1. If Γ is a 2-divisible subgroup of Q, there is a unique v-compatible ordering on A1(R). 2. If Γ is a non-2-divisible subgroup of Q, there are two v-compatible orderings on A1(R). 3. If Γ is a direct sum of a non-2-divisible subgroup of Q with Z, there are four possible v-compatible orderings on A1(R). L. Vukšić: Valuations and orderings on the real Weyl algebra 269 6.1 Extensions of orderings on A1(R) to orders on R[y; δ] In this section, we show which orders on A1(R) are extendable to an order on R[y; δ], thereby answering the question posed by Marshall and Zhang in [15]. Every order on A1(R) is compatible with a unique finest valuation v on the same ring with residue field R, as proved in [15]. Suppose v is a valuation on A1(R) associated to an infinite sequence (ωi)i≥−1 with v(ωi−1) = mini and ωi = x miωnii−1 − βi. In Section 4, we showed that provided both conditions of Theorem 4.6 are fulfilled, v can be uniquely extended to a valuation v′ on R[y; δ] with residue field R if the value group is 2-divisible and that it has two extensions to R[y; δ] with the same residue field if the value group is non-2-divisible. Here we show all v-compatible orders on A1(R) we have described in the first part of this section can be extended to a v′-compatible order on R[y; δ] for some extension v′ of v from A1(R) to R[y; δ]. Theorem 6.2. Let P be an ordering on A1(R) and v be the unique finest valuation on A1(R) compatible with P . Then: 1. The order P can be extended to an ordering on R[y; δ] if and only if v can be extended to a valuation on R[y; δ] with residue field R. 2. If the v-value group Γ is a 2-divisible subgroup of Q, then the extension of P is unique. If on the other hand, Γ is a subgroup of Γ1 × Z, where Γ1 is a non-2- divisible subgroup of Q, i.e., when Γ is not a 2-divisible subgroup of Q, there are two extensions of P to R[y; δ]. Each of the two extensions of v to a valuation v′ on R[y; δ] with residue field R uniquely determines one of the two of P ’s extensions to R[y; δ]. Proof. The first statement of the theorem follows from the fact that every ordering on R[y; δ] is compatible with a valuation on the same ring with residue field R. To prove the second statement, suppose that v is the unique P -compatible valuation on A1(R) with residue field R that extends to a valuation on R[y; δ] with the same residue field. If the value group Γ of v on A1(R) is either a 2-divisible or non-2-divisible subgroup of Q, then the value group Γ′ of v’s extension to R[y; δ] is Q. In this case, there is exactly one v′-compatible order of R[y; δ] for each of v’s extensions v′ to R[y; δ]. Suppose Γ is a 2-divisible subgroup of Q. Then there is a unique extension v′ of v to R[y; δ]. It follows that in case Γ is a 2-divisible subgroup of Q, the only v-compatible order on A1(R) extends to an order of R[y; δ] that is compatible with v′. If, on the other hand, Γ is a non-2-divisible subgroup of Q, there are two extensions v′ of v to R[y; δ]. We will now show that for each of the v-compatible orderings on A1(R), there is a unique extension v′ of v to R[y; δ] such that the ordering on A1(R) can be extended to the unique v′-compatible ordering on R[y; δ]. In this case, a v-compatible ordering on A1(R) is, as we have shown in the beginning of this section, uniquely determined by the sign of ωi−1 where i ≥ 1 is such that ni is divisible by the greatest power of two that divides nj for any j ≥ 1. Furthermore, v′, the extension of v to R[y; δ], is uniquely determined by choosing the sign of γ̃i for this i. We first choose an extension v′ of v to R[y; δ]. We observe that x m n > 0 must hold for every mn ∈ Q since all rational powers of x are in R[y; δ]. Since x mi ni ωi−1 = γ̃i for each i ≥ 1, γ̃iωi−1 > 0 must hold for all i ≥ 0 for the order to be extendable to a v-compatible 270 Ars Math. Contemp. 24 (2024) #P2.04 / 231–271 order on R[y; δ]. This holds for exactly one of the two v-compatible orders on A1(R). It is clear from the construction that for each ordering on A1(R), there is exactly one extension v′ of v to R[y; δ] such that this ordering is extendable to the unique v′-compatible ordering on R[y; δ]. In case Γ is a subgroup of Q× Z of rational rank two, Γ′ = Q× Z holds. In this case, there are two v′-compatible orderings on R[y; δ] for every extension v′ of v to R[y; δ]. We will now show that for each of the four v-compatible orders on A1(R), there is a unique extension v′ of v to R[y; δ] and a unique v′-compatible ordering P ′ on R[y; δ] such that P ′ is an extension of P . The ordering P compatible to a valuation v on A1(R) is determined by the signs of ωi−1 and ωN where i ≥ 1 is such that ni is divisible by the greatest power of two that divides nj for any j ≥ 1, and v(ωN ) ̸∈ Q. The extension v′ of v to R[y; δ] and the v′-compatible ordering on R[y; δ] that extends P are the valuation v′ for which ωi−1γ̃i > 0 and the v′-compatible ordering that agrees with the signs of ωi−1 and ωN in P . We have thus proved the second statement of the theorem. References [1] J. Cimprič, Real spectra of quantum groups, J. Algebra 277 (2004), 282–297, doi:10.1016/j. jalgebra.2004.03.011, https://doi.org/10.1016/j.jalgebra.2004.03.011. [2] P. Conrad, On ordered division rings, Proc. Am. Math. Soc. 5 (1954), 323–328, doi:10.2307/ 2032248, https://doi.org/10.2307/2032248. [3] T. C. Craven, Witt rings and orderings of skew fields, J. Algebra 77 (1982), 74–96, doi:10.1016/ 0021-8693(82)90278-2, https://doi.org/10.1016/0021-8693(82)90278-2. [4] N. I. Dubrovin, Noncommutative valuation rings in simple finite-dimensional algebras over a field, Math. USSR, Sb. 51 (1985), 493–505, doi:10.1070/SM1985v051n02ABEH002871, https://doi.org/10.1070/SM1985v051n02ABEH002871. [5] A. Engler and A. Prestel, Valued Fields, Springer Monographs in Mathematics, Springer, Berlin, 2005. [6] Á. Granja, M. C. Martı́nez and C. Rodrı́guez, Real valuations on skew polynomial rings, Algebr. Represent. Theory 17 (2014), 1413–1436, doi:10.1007/s10468-013-9454-7, https://doi. org/10.1007/s10468-013-9454-7. [7] D. N. I., Noncommutative valuation rings, Trans. Moscow Math. Soc. 45 (1984), 273–287. [8] I. Kaplansky, Maximal fields with valuations, Duke Math. J. 9 (1942), 303– 321, doi:10.1215/S0012-7094-42-00922-0, https://doi.org/10.1215/ S0012-7094-42-00922-0. [9] I. Klep and D. Velušček, n-real valuations and the higher level version of the Krull-Baer theo- rem., J. Algebra 279 (2004), 345–361, doi:10.1016/j.jalgebra.2004.05.012, https://doi. org/10.1016/j.jalgebra.2004.05.012. [10] F.-V. Kuhlmann, Book on Valuation Theory, in preparation. [11] F.-V. Kuhlmann, Value groups, residue fields and bad places od rational function fields, Trans. Am. Math. Soc. 356 (2004), 4559–4600. [12] J. Kürschak, Über Limesbildung und allgemeine Körpertheorie., J. für die Reine und Angew. Math. 142 (1913), http://eudml.org/doc/149394. [13] S. MacLane, A construction for absolute values in polynomial rings, Trans. Am. Math. Soc. 40 (1936), 363–395, doi:10.2307/1989629, https://doi.org/10.2307/1989629. L. Vukšić: Valuations and orderings on the real Weyl algebra 271 [14] T. Markwig, A field of generalised Puiseux series for tropical geometry, Rend. Semin. Mat., Univ. Politec. Torino 68 (2010), 79–92. [15] M. Marshall and Y. Zhang, Orderings, real places, and valuations on noncommutative inte- gral domains, J. Algebra 212 (1999), 190–207, doi:10.1006/jabr.1998.7630, https://doi. org/10.1006/jabr.1998.7630. [16] M. Marshall and Y. Zhang, Orderings and valuations on twisted polynomial rings, Commun. Al- gebra 28 (2000), 3763–3776, doi:10.1080/00927870008827055, https://doi.org/10. 1080/00927870008827055. [17] H. Marubayashi, H. Miyamoto and A. Ueda, Non-commutative Valuation Rings and Semi-Hereditary Orders, volume 3 of K-Monogr. Math., Kluwer Academic Publish- ers, Dordrecht, 1997, doi:10.1007/978-94-017-2436-4, https://doi.org/10.1007/ 978-94-017-2436-4. [18] G. Onay, Valued modues over skew-polynomial rings I, J. Symb. Log. 82 (2017), 1519–1540, https://www.jstor.org/stable/26600297. [19] A. Prestel and C. Delzell, Mathematical Logic and Model Theory, Springer, London, 2011, doi: 10.1007/978-1-4471-2176-3, https://doi.org/10.1007/978-1-4471-2176-3. [20] T. Rohwer, Valued difference fields as modules over twisted polynomial rings, Ph.D. thesis, University of Illinois at Urbana-Champaign, 2003, https://www.ideals.illinois. edu/items/88103. [21] O. F. G. Schilling, Noncommutative valuations, Bull. Am. Math. Soc. 51 (1945), 297–304, https://projecteuclid.org/journals/ bulletin-of-the-american-mathematical-society/volume-51/ issue-4/Noncommutative-valuations/bams/1183506882.full. [22] J. I. Stipel’man, Valuations of the quotient field of the ring of quantum mechanics, Funct. Anal. Appl. 7 (1973), 46–52, doi:10.1007/bf01075649, https://doi.org/10.1007/ bf01075649. [23] J.-P. Tignol and A. R. Wadsworth, Value Functions on Simple Algebras, and Associated Graded Rings, Springer Monogr. Math., Springer, Berlin, 2015, doi:10.1007/978-3-319-16360-4, https://doi.org/10.1007/978-3-319-16360-4. [24] A. Tschimmel, Lokal-global Prinzipien für Anordnungen bewerteter Schiefkörper, Arch. Math. 44 (1985), 48–58, doi:10.1007/bf01193780, https://doi.org/10.1007/ bf01193780. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.05 / 273–292 https://doi.org/10.26493/1855-3974.2764.fe5 (Also available at http://amc-journal.eu) Regular dessins with moduli fields of the form Q(ζp, p √ q)* Nicolas Daire Département de Mathématiques et Applications, École Normale Supérieure, 45 rue d’Ulm, 75005 Paris, France Fumiharu Kato , Yoshiaki Uchino Department of Mathematics, Tokyo Institute of Technology, 2-12-1 Ookayama, Meguro, Tokyo 152-8551, Japan Received 10 December 2021, accepted 29 March 2023, published online 27 September 2023 Abstract Gareth Jones asked during the 2014 SIGMAP conference for examples of regular dessins with nonabelian fields of moduli. In this paper, we first construct dessins whose moduli fields are nonabelian Galois extensions of the form Q(ζp, p √ q), where p is an odd prime and ζp is a pth root of unity and q ∈ Q is not a pth power, and we then show that their regular closures have the same moduli fields. Finally, in the special case p = q = 3 we give another example of a regular dessin of degree 219 ·34 and genus 14155777 with moduli field Q(ζ3, 3 √ 3). Keywords: Dessins d’enfants, coverings. Math. Subj. Class. (2020): 14H57, 14H30 1 Introduction Grothendieck first coined the term Dessin d’enfant in Esquisse d’un Programme [4] to denote a connected bicolored graph embedded on a compact connected oriented topolog- ical surface. The study was motivated by the one to one correspondance between dessins d’enfant, the combinatorial data of the associated cartographical group, and the geomet- ric concept of coverings of P1 by compact Riemann surfaces ramified at most over three *The authors are grateful to Professor Jürgen Wolfart for valuable comments. E-mail addresses: nicolas.daire@ens.psl.eu (Nicolas Daire), bungen@math.titech.ac.jp (Fumiharu Kato), uchino.y.ab@m.titech.ac.jp (Yoshiaki Uchino) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 274 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 points. Moreover, by Belyi’s theorem any such covering is given the structure of an alge- braic curve defined over a number field, therefore we obtain a natural action of the absolute Galois group Gal(Q̄/Q) on the set of isomorphism classes of dessins. A lot of the interest for dessins stems from the fact that this action is faithful, providing a way to study the absolute Galois group through its action on the set of dessins. A particularly interesting family of dessins is that of regular dessins, characterized by the fact that their automor- phism groups act transitively on their sets of edges, and the Galois action was proved to remain faithful when restricted to the subset of isomorphism classes of regular dessins [3]. To any dessin we associate a number field called its moduli field, which is defined as the subfield of Q̄ fixed by the subgroup of Gal(Q̄/Q) that fixes the dessin up to isomorphism. Conder, Jones, Streit and Wolfart noted in [1] that the moduli fields of all the examples of regular dessins known at the time were abelian Galois extensions of Q. Herradón con- structed in [6] an explicit equation for a regular dessin whose moduli field Q( 3 √ 2) is not a Galois extension of Q, and Hidalgo later generalized his construction in [7] to produce regular dessins whose moduli fields are of the form Q( p √ 2) where p is an odd prime num- ber. However there is as of yet no known example of regular dessin whose moduli field is a nonabelian Galois extension of Q. This is the starting point of this paper, in which we will exhibit examples of regular dessins with moduli fields that are nonabelian Galois extensions of Q. In the present paper, we begin by recalling the main definitions and results on dessins d’enfant. We will then expose constructions of regular dessins whose moduli fields are nonabelian Galois extensions of Q. We first exhibit dessins whose moduli fields are of the form Q(ζ3, 3 √ q), where ζ3 is a primitive third root of unity and q ∈ Q is not a third power, and show that the regular closures of these dessins possess the same moduli fields. We then generalize this construction to show that there exist regular dessins with moduli fields Q(ζp, p √ q), where ζp is a primitive pth root of unity and q ∈ Q>0 is not a pth power. Finally, we give an example of a regular dessin of degree 219 ·34 and genus 14155777 with moduli field Q(ζ3, 3 √ 3). Notations • SE : the group of self-bijections of the set E, similarly Sn is the group of permu- tations of a set of n elements (we favor a right action, hence we write the product στ := τ ◦ σ) • Gal(E/F ): the Galois group of F -automorphisms of E • ζk: the kth primitive root of unity exp( 2iπk ) • F2: the free group of rank 2 with generators (ξ, η) • Crit: the set of critical values of a function 2 Preliminaries on dessins d’enfant We refer the reader to existing expositions of the theory such as [5, 8, 9] and [2] for proofs of the presented facts and further details. A dessin d’enfant is a connected bipartite graph embedded on a compact connected orientable topological surface, such that the complement of the graph is a disjoint union of N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 275 2-cells. Two such dessins are equivalent if there exists an orientation preserving homeomor- phism between the underlying surfaces that induces an isomorphism between the embedded bipartite graphs. A dessin is determined up to isomorphism by a pair (C, β) where C is a smooth alge- braic curve and β : C → P1 is a meromorphic mapping ramified at most over {0, 1,∞}, and by Belyi’s theorem we can further ask for C and β to both be defined over a number field. We call (C, β) a Belyi pair and β a Belyi function. The corresponding graph embed- ding on the underlying surface is recovered by pulling back the segment [0, 1] along β, we define black and white vertices as the preimages of 0 and 1 respectively, and the edges as the preimages of ]0, 1[. By covering theory a dessin is also determined up to isomorphism by the monodromy action of the fundamental group of the complex projective line π1(P1) on the fiber over the point 12 which is identified to the set of edges of the dessin. The fundamental group π1(P 1) is isomorphic to the free group of rank two F2 = ⟨ξ, η⟩ with generators ξ and η which are two loops with base point 12 and circling counter-clockwise around 0 and 1 respectively. The monodromy action of the generators ξ and η then corresponds to the product of the counter-clockwise cyclic permutation of the edges around black and white vertices respec- tively. We call monodromy map M : F2 → SE the map that associates to each element of F2 the corresponding permutation of the set of edges, and we call cartographic group the image of the monodromy map, which is a transitive subgroup of the group of permutations of the set of edges. When the automorphism group of a dessin D acts transitively on the set of edges, we say that D is a regular dessin. When that is the case the cartographic group G acts transitively and freely on the set of edges, the monodromy action is thus given by the canonical action of G on itself. There is a natural bijection between regular dessins and finite groups generated by two distinguished elements ξ and η up to isomorphism. Two regular dessins determined by G1 = ⟨ξ1, η1⟩ and G2 = ⟨ξ2, η2⟩ respectively are isomorphic if and only if there exists an isomorphism between G1 and G2 that preserves the distinguished generators. Given a dessin D, there exists a unique regular dessin D̃ with a morphism ϕ : D̃ → D such that any morphism from a regular dessin to D factors through ϕ. We call D̃ the regular closure of D. Moreover, there exists an isomorphism Cart(D̃) ∼= Cart(D) that preserves the distinguished generators. There exists a natural action of the absolute Galois group Gal(Q̄/Q) on the set of isomorphism classes of dessins, we denote by Dσ the action of an automorphism σ on a dessin D, and this Galois action commutes with regular closure, i.e. we have (D̃)σ ∼= (̃Dσ). Given a dessin D, we say that a number field k is a field of definition of D if D is isomorphic to a dessin defined over k. However there does not necessarily exist a smallest field of definition. We thus define the moduli field of a dessin D as the subfield of Q̄ fixed by the subgroup of Gal(Q̄/Q) constituted of the elements fixing D up to isomorphism. The moduli field of a dessin is contained in all fields of definition but is not necessarily itself a field of definition, however it is the case in particular for regular dessins. 3 Constructions of regular dessins with nonabelian moduli fields We are now ready to give examples of regular dessins whose moduli fields are nonabelian Galois extensions of Q. To do so, we will first exhibit dessins with such moduli fields, and then prove that their regular closures admit the same moduli fields. 276 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 Before proceeding with the examples, let us first present a classic family of Belyi poly- nomials that we will use in the following constructions. For positive integers m,n ∈ N we define the polynomial Bm,n := (m+ n)m+n mmnn Xm(1−X)n ∈ Q[X]. By computing the derivative B′m,n = (m+n)m+n mmnn X m−1(1 −X)n−1(m − (m + n)X) we verify that Bm,n : P1 → P1 is a Belyi function that ramifies only at 0, 1, ∞ and mm+n with ramification indices m, n, m + n and 2 respectively, and Bm,n(0) = 0, Bm,n(1) = 0, Bm,n(∞) = ∞ and Bm,n( mm+n ) = 1 (see Figure 1). 0 1 m m+n 1 2 3 m − 1 m 1 2 3 n − 1 n Figure 1: Dessin corresponding to the Belyi pair (P1, Bm,n). 3.1 Regular dessins with moduli fields of the form Q(ζ3, 3 √ q) Let q ∈ Q>0 be a positive rational number that is not a third power. Let m,n ∈ N be coprime positive integers such that 2727+q2 = mm+n , and let C : y2 = x(x− (1− ζ3))(x− 3 √ q), β : C → P1, (x, y) 7→ 1 27mq2n (x6 + 27)m(q2 − x6)n. The function β is given by the composition β = β1 ◦ β0 ◦ π of the following maps. 1. π : C → P1 is the projection on the coordinate x, which is ramified over {0, 1 − ζ3, 3 √ q,∞}. 2. β0 := X6 ∈ Q[X], Crit(β0) = {0} so β1 ◦ π ramifies over {0, (1 − ζ3)6 = −27, q2,∞}. 3. β1 := Bm,n(X+27q2+27 ), so β = β1 ◦ β0 ◦ π ramifies over {0, 1,∞}. The pair (C, β) is thus a Belyi pair, and we call D the corresponding dessin. The dessin D is defined over Q(ζ3, 3√q), so its moduli field is a subfield of Q(ζ3, 3√q). By taking the regular closure we then obtain the inclusion of moduli fields M(D̃) ⊆ M(D) ⊆ Q(ζ3, 3 √ q), and moreover D̃ is regular so it is defined over M(D̃). We shall prove that M(D̃) is in fact exactly Q(ζ3, 3√q), which is a nonabelian Galois extension of Q with Galois group Gal(Q(ζ3, 3 √ q)/Q) ∼= S3. To that end we must show that an automorphism σ ∈ Gal(Q̄/Q) fixes D̃ if and only if it fixes ζ3 and 3 √ q, or equivalently that Gal(Q(ζ3, 3 √ q)/Q) acts freely on the orbit of D̃. N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 277 Let σ ∈ Gal(Q̄/Q), the Galois conjugate Dσ is given by the Belyi pair (Cσ, βσ), where Cσ : y2 = x(x− (1− σ(ζ3)))(x− σ( 3 √ q)), and βσ has the same expression as β because all of its coefficients are rational. The orbit of the pair (ζ3, 3 √ q) by Gal(Q̄/Q) is {ζi3, ζj3 3 √ q}1≤i≤2,0≤j≤2. Elliptic curves given by equations of the form y2 = (x− a)(x− b)(x− c) are isomorphic if and only if the cross- ratios of the tuples (a, b, c,∞) coincide. We verify that the cross-ratios are all distinct, so the orbit of D is given by the six dessins Dσ for σ ∈ Gal(Q(ζ3, 3√q)/Q). As a consequence M(D) = Q(ζ3, 3√q). To prove that the regular closures D̃σ constituting the orbit of D̃ are also non isomorphic, we must first draw the dessins Dσ to compute their cartographic groups. Let us first draw the dessin D0 corresponding to the Belyi pair (P1, β1 ◦ β0) (see Fig- ure 3). The dessin D0 is defined over Q, so the dessins Dσ in the orbit are then obtained by lifting D0 to the curves Cσ . To simplify the graphical representations of the dessins, we will use the notation in Figure 2 for consectutive edges incident to a vertex. eξ eξ 2 eξ n−2 eξ n−1 eeξ n = n eeξ n Figure 2: Notation for consecutive edges. (P1, id)0 1 (P1, β1) −27 0 q2 m n (P1, β1 ◦ β0) 0 n m n m n m n m n m n m 3 √ q ζ3 3 √ q ζ23 3 √ q 1− ζ23 1− ζ3 Figure 3: Construction of D0. The dessins D1, . . . ,D6 conjugate to D are embedded on a torus, so in the representa- tions in Figure 4 we will identify the outermost edges on opposite sides. 278 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 ? ∞ 0 0 00 3 √ q 3 √ q 1− ζ3 1− ζ3 n 4 m 3 n 16 m 15 n 6 m 5 n 18 m 17 n 8 m 7 n 20 m 19 n 10 m 9 n 22 m 21 n 12 m 11 n 24 m 23 n n m m 1 13 14 2 131 2 14 (a) D1 := D ? ∞ 0 0 00 ζ3 3 √ q ζ3 3 √ q 1− ζ23 1− ζ23 n 2 m 3 n 14 m 15 n 6 m 5 n 18 m 17 n 8 m 7 n 20 m 19 n 10 m 9 n 22 m 21 n 12 m 11 n 24 m 23 n n m m 1 13 16 4 131 4 16 (b) D2 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , ζ3 3 √ q) ? ∞ 0 0 00 ζ3 3 √ q ζ3 3 √ q 1− ζ3 1− ζ3 n2 m 3 n 14 m 1 5 n 4 m 5 n 16 m 17 n 8 m 7 n 20 m 19 n 10 m 9 n 22 m 21 n 12 m 11 n 24 m 23 n n m m 1 13 18 6 131 6 18 (c) D3 := Dσ, σ : (ζ3, 3√q) 7→ (ζ3, ζ3 3√q) ? ∞ 0 0 00 ζ23 3 √ q ζ23 3 √ q 1− ζ23 1− ζ23 n2 m 3 n 14 m 15 n 4 m 5 n 16 m 17 n 6 m 7 n 18 m 19 n 10 m 9 n 22 m 21 n 12 m 11 n 24 m 23 n n m m 1 13 20 8 131 8 20 (d) D4 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , ζ23 3 √ q) N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 279 ? ∞ 0 0 00 ζ23 3 √ q ζ23 3 √ q 1− ζ3 1− ζ3 n2 m3 n 14 m 15 n 4 m 5 n 1 6 m 17 n 6 m 7 n 18 m 19 n 8 m 9 n 20 m 21 n 12 m 11 n 24 m 23 n n m m 1 13 22 10 131 10 22 (e) D5 := Dσ, σ : (ζ3, 3√q) 7→ (ζ3, ζ23 3 √ q) ? ∞ 0 0 00 3 √ q 3 √ q 1− ζ23 1− ζ23 n2 m3 n 14 m 15 n 4 m 5 n 16 m 1 7 n 6 m 7 n 18 m 19 n 8 m 9 n 20 m 21 n 10 m 11 n 22 m 23 n n m m 1 13 24 12 131 12 24 (f) D6 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , 3 √ q) Figure 4: Dessins D1, . . . ,D6 in the Galois orbit of D. We will now establish that D̃1 is not isomorphic to D̃2, . . . , D̃6. To that end it suffices to show that there is no isomorphism between the cartographic groups fixing the canonical generators. We shall therefore exhibit an element ω ∈ F2 = ⟨ξ, η⟩ such that Mk(ω) commutes with Mk(η2) only when k = 1, where Mk is the monodromy map of Dk. We have defined m and n to be positive coprime integers such that 2727+q2 = m m+n , so we cannot have m = n = 1. We will treat the case where m ̸= 1 does not divide n, the other case being treated similarly. Let ω := ξnη−1ξm−nηξn. We shall show that Mk(ω) commutes with Mk(η2) only when k = 1. Let Ek := {1, 2, . . . , 24} be the set of edges of Dk incident to 0. The action of η fixes the set Ek on which it induces the cyclic permutation (1, 2, . . . , 24), and every white vertex except 0 has degree one so the action of η is trivial on the complement of Ek. We can write Ek = Eoddk ⊔ Eevenk as the disjoint union of the sets of respectively odd and even numbered edges incident to 0, such that η sends one to the other. The black vertices of Eoddk are of degree m except for the two black vertices of the edges 1 and 13 that are of degree 2m. Therefore if m does not divide some integer l then ξl sends every edge of Eoddk to the complement of Ek, and otherwise the action of ξ m on Eoddk corresponds to the sole transposition (1, 13). Similarly if n does not divide l then ξl sends every edge of Eevenk to the complement of Ek, and the action of ξn on Eevenk is the transposition (2k, 2k + 12). In particular, by hypothesis n is not a multiple of m, so m − n is not a multiple of m either, hence both ξn and ξm−n send the edges of Eoddk to the complement of Ek. However η acts trivially on the latter, so ξnη−1ξm−n and ξm−nηξn both fix the set Eoddk on which they induce the same action as ξm, i.e. the transposition (1, 13). Therefore the action of ω = ξnη−1ξm−nηξn is the same as that of ξmηξn on Eoddk and the same as that of ξnη−1ξm on Eevenk . See Figure 5. The action of ω fixes the set Ek on which it induces the permutation Mk(ω)|Ek = (1, 13)(2k, 2k + 12) · (1, 2)(3, 4) · · · (23, 24) · (1, 13)(2k, 2k + 12) 280 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 0 m ξn η−1 ξm−n n ξn η e eω 0 m ξm−n η ξn n ξn η−1 eω e Figure 5: Action of ω on Eoddk \ {1, 2k − 1} and on Eevenk \ {2, 2k}. Therefore for k = 1, M1(ω)|E1 = (1, 13)(2, 14) · (1, 2)(3, 4) · · · (23, 24) · (1, 13)(2, 14) = (1, 2)(3, 4) · · · (23, 24) so ω and η2 commute on E1. Moreover η acts trivially on the complement of E1 so M1(ω)|D1\E1 and M1(η2)|D1\E1 automatically commute. Finally, we obtain that M1(ω) and M1(η2) commute. For k = 2, we observe that 4ωη 2 = 15η 2 = 17 but 4η 2ω = 6ω = 5. Similarly, for 3 ≤ k ≤ 6, we observe that 1ωη2 = 14η2 = 16 but 1η2ω = 3ω = 4. We have thus shown that Mk(ω) and Mk(η2) commute only for k = 1. This concludes the proof that D̃ is a regular dessin with moduli field Q(ζ3, 3√q). 3.2 Regular dessins with moduli fields of the form Q(ζp, p √ q) Let p be an odd prime, and q ∈ Q>0 a positive rational number that is not a pth power. In this example we will need an additional parameter γ ∈ Q \ {0}. Let C : y2 = x(x− (1− ζp))(x− γ p √ q). We construct the Belyi function β : C → P1 as the composition β = β2 ◦ β1 ◦ β0 ◦ π of the following maps. 1. π : C → P1 is the projection on the coordinate x, which ramifies over {0, 1 − ζp, γ p √ q,∞}. 2. β0 := X2p ∈ Q[X], and Crit(β0) = {0,∞} so β0 ◦ π ramifies over {0, (1 − ζp) 2p, γ2pq2,∞}. 3. β1 ∈ Q[X] is chosen independently of γ such that Crit(β1) ∪ {β1((1 − ζp)2p)} = {0, 1,∞}, β1((1−ζp)2p) = 0 < β1(0) < 1 and β′1(0) > 0. The existence of β1 veri- fying those conditions is assured by Proposition 3.2 below. Under those assumptions β1 ◦ β0 ◦ π ramifies over {0, 1, β1(0), β1(γ2pq2),∞}. 4. γ ∈ Q>0 is then chosen small enough so that β′1 > 0 on [0, γ2pq2]. This guarantees us that we have 0 < β1(0) < β1(γ2pq2) < 1. 5. β2 := Br,s ◦ Bm,n, where (m,n) and (r, s) are pairs of coprime positive integers such that β1(γ2pq2) = mm+n and Bm,n(β1(0)) = r r+s . Finally, β = β2 ◦ β1 ◦ β0 ◦ π ramifies over {0, 1,∞}. N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 281 The pair (C, β) is thus a Belyi pair, and we call D the corresponding dessin. With the same arguments as before, the moduli field of D is Q(ζp, p√q), which is a nonabelian Galois extension of Q with Galois group Gal(Q(ζp, p √ q)/Q) ∼= Z/pZ ⋊ (Z/pZ)× generated by σ : ζip p √ q 7→ ζi+1p p √ q and τ : ζip p √ q 7→ ζgip p √ q where g generates (Z/pZ)×. We shall show that there exists γ ∈ Q \ {0} such that the regular closure of the dessin D thus obtained also has moduli field Q(ζp, p √ q). Remark 3.1. In the previous subsection we treated the case p = 3. In that specific case we gave a simpler expression for β, mainly due to the fact that β0 ◦ π already had all of its critical values in Q∪{∞}. However in the general case we must use the intermediate map β1 as well as the parameter γ to conclude the proof. Let us first prove the existence of β1. Proposition 3.2. Let E ⊂ Q̄ ∩ R \ {0} be a finite set. Then there exists P ∈ Q[X] such that P (E) ⊆ {0}, Crit(P ) ⊆ {0, 1}, 0 < P (0) < 1 and P ′(0) > 0. Remark 3.3. In the context of this proposition we only deal with polynomials so for P ∈ Q[X] we define Crit(P ) := {P (z)| z ∈ C, P ′(z) = 0}, which does not include the point at infinity to simplify notations. Proof. To show this we will proceed similarly as in the proof of the only if part of Belyi’s theorem, by applying additional transformations to ensure that 0 < P (0) < 1. Let us first prove that we can reduce to the case where E is a subset of rational numbers. Lemma 3.4. Let E ⊂ Q̄ ∩ R \ {0} be a finite set fixed by Gal(Q̄/Q). Then there exists P ∈ Q[X] such that P (0) = 0 and Crit(P ) ∪ P (E) ⊂ Q \ {0}. Proof. Let {a1, . . . , am} = E∩Q and {b1, . . . , bn} = E\Q. We construct P by induction on the number n of non rational elements of E. For α ∈ Q, define Fα, Gα ∈ Q[X] by Fα := n∏ j=1 (X − (bj − α)2) and Gα := Fα((X − α)2) = n∏ j=1 (X − bj)(X + bj − 2α). Let us first assume that there exists α ∈ Q such that Gα(0) ̸∈ Crit(Gα) ∪ Gα(E). Define P1(X) := Gα(X)−Gα(0) ∈ Q[X], then P1(0) = 0 ̸∈ E′ := Crit(P1)∪P1(E) ⊂ Q̄ ∩ R \ {0}. Note that E′ is stable under the action of Gal(Q̄/Q), and |E′ \ Q| = |Crit(Fα) ∪ Fα(0) \ Q| = |Crit(Fα) \ Q| < degFα = n. By induction, there exists P2 ∈ Q[X] such that P2(0) = 0 and Crit(P2)∪P2(E′) ⊂ Q\{0}. Now P := P2 ◦P1 has the desired properties, since P (0) = 0 and Crit(P )∪P (E) = Crit(P2)∪P2(Crit(P1))∪ P2(P1(E)) = Crit(P2) ∪ P2(E′) ⊂ Q \ {0}. Let us now prove that there exists α ∈ Q such that Gα(0) ̸∈ Crit(Gα) ∪ Gα(E). Let us first treat the case where 0 < b1 < b2, . . . , bn. When α approaches 12 , Gα(0) =∏n j=1 −bj(bj−2α) approaches 0 but the critical values of Gα do not. Indeed, Crit(Gα) = CritFα∪Fα(Crit((X−α)2)) = Crit(Fα)∪{Fα(0)}; Fα(0) approaches F b1 2 (0) ̸= 0, and since F b1 2 does not have multiple roots, the critical values of Fα approach the critical values 282 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 of F b1 2 which are all non zero. Therefore for α ̸= b12 in the neighborhood of b12 we have Gα(0) ̸∈ Crit(Gα). Moreover Gα(0), Gα(a1), . . . , Gα(am) are all distinct polynomials in the indeterminate α, so they coincide at only finitely many points. In particular for α ̸= b12 in the neighborhood of b12 we have Gα(0) ̸∈ {Gα(a1), . . . , Gα(am)}. Since α ∈ Q we also have Gα(0) ̸= 0 = Gα(b1) = · · · = Gα(bn) hence Gα(0) ̸∈ Gα(E), proving the existence of α as desired. Let us now treat the general case where b1, . . . , bn are not assumed to be positive by reducing it to the previous case. For α′ ∈ Q, define Hα′ ∈ Q[X] by Hα′ := (X − α′)2 − α′2 ∈ Q[X]. Note that Crit(Hα′) = {−α′}. For α′ > 0 sufficiently small we have −α′2 < Hα′(0) = 0 < Hα′(a1), . . . ,Hα′(am), Hα′(b1), . . . ,Hα′(bn). Let E′′ := Crit(Hα′) ∪ Hα′(E). The set E′′ is a finite subset of Q̄ ∩ R \ {0} fixed by Gal(Q̄/Q), and E′′ has at most n non rational elements, which are all positive. By the above, there exists P3 ∈ Q[X] such that Crit(P3) ∪ P3(E′′) ⊂ Q \ {0} and P3(0) = 0. Then P := P3 ◦ Hα′ has the desired properties, since P (0) = 0 and Crit(P ) ∪ P (E) = Crit(P3) ∪ P3(Crit(Hα′)) ∪ P3(Hα′(E)) = Crit(P3) ∪ P3(E′′) ⊂ Q \ {0}. Let us denote by P1 the polynomial obtained using this lemma, which verifies P1(0) = 0 and E′ := Crit(P1) ∪ P1(E) ⊂ Q \ {0}. We can further assume that P ′1(0) > 0 by taking (−P1) if necessary. We now send the points E′ to {0, 1}. Lemma 3.5. Let E ⊂ Q \ {0} a finite set. Then there exists P ∈ Q[X] such that P (E) ⊆ {0}, Crit(P ) ⊆ {0, 1}, 0 < P (0) < 1 and P ′(0) > 0. Proof. For α ∈ Q, let Fα := (X − α)2 ∈ Q[X], and note that Crit(Fα) = {0}. There exists α < 0 sufficiently small such that 0 < Fα(0) < Fα(a) for all a ∈ E. We take F := Fα maxa∈E Fα(a) . Let {a1, . . . , al} = F (E) such that 0 < F (0) < a1 < · · · < al = 1. We also add a rational point a0 ∈ Q such that F (0) < a0 < a1. Let m and n be the coprime positive integers such that al−1 = mm+n . We recall that Bm,n verifies Crit(Bm,n) = {0, 1}, Bm,n(0) = Bm,n(1) = 0, Bm,n( mm+n ) = 1, and Bm,n is strictly increasing between 0 and mm+n . Let P1 := Bm,n, then Crit(P1) = {0, 1} and 0 < P1 ◦F (0) < P1(a0) < · · · < P1(al−1) = 1. There is one point fewer than before, so we can iteratively construct P2, . . . , Pl in the same way, so that P := Pl ◦ · · · ◦ P1 verifies Crit(P ) ⊆ {0, 1}, P (a1) = · · · = P (al) = 0 < P (F (0)) < 1 = P (a0) and P ′(F (0)) > 0. Therefore P ◦ F has the desired properties. Let us denote by P2 the polynomial obtained using this lemma with the finite set E′ obtained previously. Then the polynomial P := P2 ◦ P1 verifies P (E) ⊆ {0}, Crit(P ) ⊆ {0, 1}, 0 < P (0) < 1 and P ′(0) > 0, thus concluding the proof of Proposition 3.2. We can now use Proposition 3.2 with the finite set E := {(1− ζkp )2p}1≤k≤ p−12 N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 283 to obtain the map β1 as desired. For 1 ≤ k ≤ p−12 we have (1 − ζkp )2p = (|1 − ζkp |ζ2k−12p )2p = |1 − ζkp |2p ∈ R so E ⊂ Q̄ ∩ R, and the set E is the Galois orbit of (1− ζp)2p so it is fixed by Gal(Q̄/Q), hence E verifies the conditions of Proposition 3.2. Let us denote by D(β1) the dessin corresponding to the Belyi pair (P1, β1). The Belyi pair (P1, β1) is fixed by the action of the complex conjugation, so the embedding of D(β1) on P1 admits a symmetry along the real line. Moreover the Belyi function β1 is a polyno- mial, so D(β1) ∩ R is a (graph theoretic) path. Let vl < · · · < v1 be the negative vertices on the path, and let ek denote the edge (vk−1, vk). By hypothesis β′1(0) > 0 so v1 is a black vertex, and for k < l, the vertex vk is of even degree 2dk. We then have e ξdk k = ek+1 and eξ dk k+1 = ek if k is odd, or e ηdk k = ek+1 and e ηdk k+1 = ek if k is even. See Figure 6. As remarked earlier, the Galois orbit of (1− ζp)2p is {(1− ζkp )2p}1≤k≤ p−12 ⊂ R−, and (1− ζ p−1 2 p )2p < · · · < (1− ζp)2p < 0. By construction β1((1− ζp)2p) = 0, so (1− ζp)2p and all its Galois conjugates are black vertices of D(β1) lying on the path (v1, · · · , vl). Let t > 0 be the index such that vt = (1− ζp)2p, and vt is a black vertex so t is odd. Then µ0 := ξ d1ηd2 · · · ηdt−1ξ2dtηdt−1 · · · ηd2ξd1 fixes the edge e1 (Figure 6). vt v2 v1 ×0 e3 e2 e1 µ0 ξd1 ξd1 ηd2 ηd2 ξdt ξdt Figure 6: Dessin D(β1) corresponding to (P1, β1). Let γ > 0 small enough so that β′1 > 0 on [0, γ 2pq2]. Let us next draw the dessin D(β2) corresponding to the Belyi pair (P1, β2 = Br,s ◦Bm,n). See Figure 7. (P1, id)0 1 (P1, Br,s) 0 r r+s 1 r s (P1, β2 = Br,s ◦Bm,n)0 m m+nβ1(0) 1 s s mr nr Figure 7: Dessin D(β2) corresponding to (P1, β2). By lifting the dessin D(β2) along β1 we obtain the dessin D(β2 ◦ β1) corresponding to the Belyi pair (P1, β2 ◦ β1). This amounts to replacing each edge of D(β1) by a copy of D(β2). Note that the degrees of the black and white vertices are thus multiplied by mr and 284 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 nr, respectively. Analogously to µ0 we define µ :=(ξmrd1ηξsη)(ξnrd2ηξsη) · · · (ξmrdt−2ηξsη)(ξnrdt−1ηξsη) · (ξ2mrdtηξsη)(ξnrdt−1ηξsη)(ξmrdt−2ηξsη) · · · (ξnrd2ηξsη)ξmrd1 and we verify again that µ fixes the edge (0, v1). Note also that ξ2s fixes the edge (0, γ p √ q). See Figure 8. vt v2 v1 γ2pq20 s s s s mr mr mr mr mr mr mr mr µ nr nr nr nr η η η η η η ξs ξs ξs ξs ξmrd1 ξmrd1 ξnrd2 ξnrd2 ξmrdt ξmrdt Figure 8: Dessin D(β2 ◦ β1) corresponding to (P1, β2 ◦ β1). Let D0 be the dessin corresponding to the Belyi pair (P1, β2 ◦ β1 ◦ β0). To simplify the representations of the dessins we only show the vertices 0, ζk2p(1 − ζp), 1 − ζkp , and ζk2pγ p √ q. We decorate the vertices ζk2p(1 − ζp) (which map to (1 − ζp)2p ∈ R− by β0) and ζk2pγ p √ q (which map to γ2pq2 ∈ R+ by β0) respectively with the symbols ⊖ and ⊕ to distinguish them. See Figure 9. 0 ⊕ γ p √ q 1− ζp 1− ζ̄p ⊕ ⊕ 1− ζ 2p 1− ζ − 2 p ⊕ ζ p γ p√ q ⊕ ζ − 1 p γ p √ q 1 − ζ 3 p 1− ζ − 3 p D0 = D(β2 ◦ β1 ◦ β0) D(β2 ◦ β1) 0 γ2pq2 ⊕ (1− ζp)2p (1− ζ2p)2p (1− ζ3p)2p Figure 9: Dessin D0 corresponding to (P1, β2 ◦ β1 ◦ β0). We may now draw the Galois conjugates Dσ for σ ∈ Gal(Q̄/Q) by lifting the dessin D0 along the projection π, by treating separately the cases σ(ζp) ∈ {ζp, ζ̄p} and σ(ζp) ∈ N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 285 {ζ2p , . . . , ζp−2p }. We call the dessins respectively Dk and Djk, see Figure 10. We identify the outermost edges on opposite sides in the representations. ? ∞ A A C C B B D D 0 0 00 ζ•p p √ q ζ•p p √ q ⊕ ⊕ 1− ζ±1p 1− ζ±1p ⊕ ×(2p− k) ⊕ ×(2p− k) ⊕ ×(k − 1) ⊕ ×(k − 1) µ µ ξ2s ξ2s (a) Dk := Dσ with σ(ζp) = ζ±1p . ? ∞ A A C C B B D D 0 0 00 ζ•p p √ q ζ•p p √ q ⊕ ⊕ 1− ζjp 1− ζjp ⊕ ×(2p− k) ⊕ ×(2p− k) ⊕ ×(k − 1) ⊕ ×(k − 1) µ µ ξ2s ξ2s (b) Djk := Dσ with σ(ζp) = ζjp where j ∈ {2, . . . , p− 2}. Figure 10: Dessins (a) Dk and (b) Djk in the Galois orbit of D. For all k, we have in fact D2k−1 = Dσ where σ : (ζp, p√q) 7→ (ζp, ζk−1p p √ q), and D2k = Dσ where σ : (ζp, p√q) 7→ (ζ̄p, ζkp p √ q). We have similar expressions for the dessins Djk. Let k be fixed, and let us consider the dessin Dk. Let A denote one of the two edges incident to 0 and on the path to the ramification point 1 − ζ±1p . We also call B := Aη 2k−1 , C := A4p, D := Cη 2k−1 (See Figure 10a). Let E denote the set of edges inci- dent to 0. The action of η induces the cyclic permutation of the edges of E = {Aηi}0≤i<8p. Furthermore by construction every white vertex aside from 0 has degree 1 or 2, so η2 fixes every edge in the complement of E. We can write E = E⊖ ⊔ E⊕ as the disjoint union of E⊖ := {A2i}0≤i<4p and E⊕ := {B2i}0≤i<4p, such that η sends one to the other. The action of µ on E⊖ is the transposition (A,C), and similarly the action of ξ2s on E⊕ is the transposition (B,D). We do the same for the dessins of the form Djk, with the only difference that this time the action of µ on E⊖ is trivial, including on the edges A and C (see Figure 10b). We are almost in the same configuration as in the first example. We define analogously ω := µηµ−1ξ2sη−1µ, and we shall prove that for some choices of γ, the actions of ω and of η2 commute only for D1. To reproduce the proof in the first example we need only show that for some choice of γ the actions of µηµ−1ξ2s and µ−1ξ2sη−1µ on the set E⊕ is the same as that of ξ2s. Note that for any edge e ∈ E⊕, the edge eξi is fixed by η if i is not a multiple of s. To that end we shall show that for some choice of γ the action of µ on E⊕ is the same as that of ξδ , where δ is the number of occurences of ξ in the word µ, and then that δ is not a multiple of s. 286 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 We define the words ρ1, ρ′1, ρ2, ρ ′ 2, . . . , ρ2t−2, ρ ′ 2t−2 ∈ F2 to be the increasing subse- quence of the prefixes ending in η of the word µ defined above, such that ρ1 := ξmrd1η, ρ′1 := ρ1ξ sη, ρ2 := ρ′1ξ nrd2η, ρ′2 := ρ2ξ sη, etc., and µ = ρ′2t−2ξ mrd1 . We shall show by induction that for some choice of γ the action of ρi (resp. ρ′i) is the same as the action of ξδi (resp. ξδ ′ i ), where δi (resp. δ′i) is the number of occurences of ξ in the word ρi (resp. ρ′i). By induction it suffices to show that δi, δ ′ i are not multiples of s. Modulo s we have δi ≡ δ′i equal to the non empty partial sum of mrd1+nrd2+ · · ·+mrdt−2+nrdt−1+2mrdt+nrdt−1+mrdt−2+ · · ·+nrd2+mrd1 consisting of the first i terms. To proceed we shall use the following result, but let us first introduce some notations. Let P = ∑d i=0 ciX i ∈ Z[X] and c ∈ Z>0 such that β1 = Pc . Note that P and c do not depend on the choice of γ, and 0 < β1(0) = P (0) c < 1 so 0 < c0, c− c0. We define α := v2(γ 2pq2), ν := v2(c0) + v2(c− c0), where v2 denotes the 2-valuation. Lemma 3.6. If α > ν, then there exists e ∈ Z such that em ≡ c0 mod 2α and en ≡ c−c0 mod 2α, and v2(s) ≥ α− ν. Proof. Let a, b ∈ Z coprime such that γ2pq2 = ab 2α. Firstly, m m+ n = β1( a b 2α) = P (ab 2 α) c = ∑d i=0 cia i2αibd−i bdc , so there exists f ∈ Z such that fm = ∑di=0 ciai2αibd−i and f(m+ n) = bdc, so em ≡ c0 mod 2α, en ≡ c− c0 mod 2α for e ∈ Z such that ebd ≡ f mod 2α. Secondly, r r + s = Bm,n(β1(0)) = β1(0) m(1− β1(0))n β1( a b 2 α)m(1− β1(ab 2α))n = bd(m+n)cm0 (c− c0)n (bdP (ab 2 α))m(bdc− bdP (ab 2α))n , so there exists g ∈ Z such that gr = bd(m+n)cm0 (c−c0)n and g(r+s) = (bdP (ab 2α))m(bdc− bdP (ab 2 α))n. In the expansion of (bdP (ab 2 α))m, aside from the constant term bdmcm0 , ev- ery other term is a multiple of an integer of the form ci02 αj with i ≤ m− 1 and j ≥ m− i. By hypothesis α > ν ≥ v2(c0), so those other terms are all multiples of 2α+(m−1)v2(c0), hence there exists A ∈ Z such that (bdP (ab 2α))m = bdmcm0 +A2α+(m−1)v2(c0). Similarly there exists B ∈ Z such that (bdc − bdP (ab 2α))n = bdn(c − c0)n + B2α+(n−1)v2(c−c0). Then g(r+ s) = bd(m+n)cm0 (c− c0)n +C2α+(m−1)v2(c0)+(n−1)v2(c−c0) for some C ∈ Z, so gr = bd(m+n)cm0 (c − c0)n and gs = C2α+(m−1)v2(c0)+(n−1)v2(c−c0). The integers r and s are coprime, so after dividing gr and gs by their greatest common dividor we obtain that v2(s) ≥ α− ν > 0. N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 287 Using this lemma, we know that if α > ν, then there exists e ∈ Z such that em ≡ c0 mod 2α and en ≡ c − c0 mod 2α, v2(s) ≥ α − ν where ν does not depend on γ, and r is coprime to s so is not a multiple of 2. Therefore there exists e′ ∈ Z such that e′mr ≡ c0 mod 2α and e′nr ≡ c − c0 mod 2α. Moreover 2α−ν is a common divisor of 2α and s, so by the above modulo 2α−ν we have e′δi ≡ e′δ′i equal to the non empty partial sum δ̃i consisting of the first i terms of the sum c0d1 + (c− c0)d2 + · · ·+ c0dt−2 + (c− c0)dt−1 + 2c0dt +(c− c0)dt−1 + c0dt−2 + · · ·+ (c− c0)d2 + c0d1. Similarly e′δ is equal modulo 2α−ν to the whole sum δ̃ := 2(c0d1 + (c− c0)d2 + · · ·+ c0dt−2 + (c− c0)dt−1 + c0dt). By construction c0, c − c0, di are positive and do not depend on the choice of γ, so 0 < c0d1 ≤ δ̃i ≤ δ̃, thus for any choice of γ such that α > ν and δ̃ < 2α−ν (for instance γ = 2 u 2v+1 with 1 ≪ u ≪ v), we obtain δ̃i, δ̃ ̸≡ 0 mod 2α−ν , and in consequence δi, δ′i and δ are not multiples of s. Therefore we can now conclude by induction that the actions of ρi and ρ′i are the same as that of ξ δi and ξδ ′ i , respectively. Indeed, δ1 is not a multiple of s so ρ1 = ξδ1η and ξδ1 have the same action on E⊕. If ρi has the same action as ξδi on E⊕, then ρ′i = ρiξ sη has the same action as ξδiξsη = ξδ ′ iη on E⊕, and also the same action as ξδ ′ i because δ′i is not a multiple of s. Similarly, if ρ ′ i has the same action as ξ δ′i on E⊕, then ρi+1 has the same action as ξδi+1η on E⊕, and also the same action as ξδi+1 because δi+1 is not a multiple of s. We have thus proved that µ has the same action as ξδ on E⊕, and by symmetry µ−1 has the same action as ξ−δ on E⊕. And δ and 2s− δ are not multiples of s, so µηµ−1ξ2s and µ−1ξ2sη−1µ have the same action as ξ2s on E⊕, as announced. We shall now observe the action of ω = µηµ−1ξ2sη−1µ on E. Let Mk and M j k denote the monodromy maps of the dessins Dk and Djk. For the dessins Dk for 1 ≤ k ≤ 2p, the action of µ on E⊖ is the transposition (A,C), and the action of ξ2s on E⊕ is the transposition (B,D), therefore the action of ω fixes the set E on which it induces the permutation Mk(ω)|E = (A,C)(B,D) · 4p−1∏ i=0 (Aη 2i , Aη 2i+1 ) · (A,C)(B,D). Hence for k = 1, M1(ω)|E = (A,Aη 4p )(Aη, Aη 4p+1 ) · 4p−1∏ i=0 (Aη 2i , Aη 2i+1 ) · (A,Aη4p)(Aη, Aη4p+1) = 4p−1∏ i=0 (Aη 2i , Aη 2i+1 ) so ω and η2 commute on E. Moreover η2 acts trivially on the complement of E, so finally M1(ω) and M1(η2) commute. 288 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 For k = 2, we observe that Bωη 2 = Dη −1η2 = Dη but Bη 2ω = Bη 2η−1 = Bη . Similarly, for 3 ≤ k ≤ 2p, we observe that Aωη2 = Cηη2 = Cη3 but Aη2ω = Aη2η = Aη3 . Therefore Mk(ω) and Mk(η2) do not commute for 2 ≤ k ≤ 2p. For the dessins Djk for 1 ≤ k ≤ 2p and 2 ≤ j ≤ p−12 , ξ2s on E⊕ is the transposition (B,D), and µ acts trivially on E⊖, therefore the action of ω fixes the set E on which it induces the permutation M jk(ω)|E = (B,D) · 4p−1∏ i=0 (Aη 2i , Aη 2i+1 ) · (B,D). Hence we observe that Bωη 2 = Dη −1η2 = Dη but Bη 2ω = Bη 2η−1 = Bη , so M jk(ω) and M jk(η 2) do not commute. We have thus shown that the actions of ω and η2 commute only for D1, this concludes the proof that D̃ is a regular dessin with moduli field Q(ζp, p√q). 3.3 Regular dessin with moduli field Q(ζ3, 3 √ 3) Finally, let us exhibit a regular dessin with moduli field Q(ζ3, 3 √ 3) of smaller degree by choosing a Belyi map that is a rational function instead of a polynomial as was done in the previous subsections. Let C : y2 = x(x− (1− ζ3))(x− 3 √ 3), β : C → P1, (x, y) 7→ (x+ 3 3)3 35(x− 32)2 . The function β is given by the composition of the following maps β = β1 ◦ β0 ◦ π. 1. π : C → P1 is the projection on the coordinate x, which ramifies over {0, 1 − ζ3, 3 √ 3,∞}. 2. β0 := X6 ∈ Q[X], Crit(β0) = {0} so β0 ◦ π ramifies over {0, (1 − ζ3)6 = −33, 32,∞}. 3. β1 := (X+33)3 35(X−32)2 , Crit(β1) = {0, 1} so β = β1 ◦ β0 ◦ π ramifies over {0, 1,∞}. The pair (C, β) is thus a Belyi pair, and we call D the dessin corresponding to (C, β). Similarly as in 3.1, D has moduli field Q(ζ3, 3 √ 3). We will proceed analogously to show that the regular closure D̃ has the same field of moduli. Let us first draw the dessin D0 corresponding to the Belyi pair (P1, β1 ◦ β0) (see Figure 11), and lift it to the conjugate curves Cσ to obtain the conjugate dessins Dσ for σ ∈ Gal(Q(ζ3, 3 √ 3)/Q) (see Figure 12). As usual we identify the outermost edges on opposite sides. We can now compute the cartographic groups of the dessins. Let Mk denote the mon- odromy map of Dk. Then Mk(ξ) = (1,13,14,7,25,26)(2,15,16)(3,17,18)(4,19,20)(5,21,22) (6,23,24)(8,27,28)(9,29,30)(10,31,32)(11,33,34)(12,35,36) for all 1 ≤ k ≤ 6, and N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 289 (P1, id)0 1 (P1, β1)? −33 0 1 32 (P1, β1 ◦ β0) 0 ? ?? ? ? ? 1− ζ23 1− ζ3 3 √ 3 ζ3 3 √ 3 ζ23 3 √ 3 Figure 11: Construction of D0. ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 25 26 0 0 00 ∞3√3 3 √ 3 1− ζ3 1− ζ3 1 7 1 7 (a) D1 := D ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 25 26 0 0 00 ∞ ζ3 3 √ 3 ζ3 3 √ 3 1− ζ23 1− ζ23 1 7 1 7 (b) D2 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , ζ3 3 √ q) 290 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 25 26 0 0 00 ∞ ζ3 3 √ 3 ζ3 3 √ 3 1− ζ3 1− ζ3 1 7 1 7 (c) D3 := Dσ, σ : (ζ3, 3√q) 7→ (ζ3, ζ3 3√q) ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 2526 0 0 00 ∞ ζ23 3 √ 3 ζ23 3 √ 3 1− ζ23 1− ζ23 1 7 1 7 (d) D4 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , ζ23 3 √ q) ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 25 26 0 0 00 ∞ ζ23 3 √ 3 ζ23 3 √ 3 1− ζ3 1− ζ3 1 7 1 7 (e) D5 := Dσ, σ : (ζ3, 3√q) 7→ (ζ3, ζ23 3 √ q) ? ?? 2 8 3 9 4 10 5 11 6 12 15 27 16 28 17 29 18 30 19 31 20 32 21 33 22 34 23 35 24 36 13 14 25 26 0 0 00 ∞3√3 3 √ 3 1− ζ23 1− ζ23 1 7 1 7 (f) D6 := Dσ, σ : (ζ3, 3√q) 7→ (ζ23 , 3 √ q) Figure 12: Dessins D1, . . . ,D6 in the Galois orbit of D. N. Daire et al.: Regular dessins with moduli fields of the form Q(ζp, p √ q) 291 • M1(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,36)(14,15)(16,17)(18,19) (20,21)(22,23)(24,25)(26,27)(28,29)(30,31)(32,33)(34,35) , • M2(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,36)(14,27)(15,26)(16,29) (17,28)(18,19)(20,21)(22,23)(24,25)(30,31)(32,33)(34,35) , • M3(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,36)(14,27)(15,26)(16,17) (18,31)(19,30)(20,21)(22,23)(24,25)(28,29)(32,33)(34,35) , • M4(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,36)(14,27)(15,26)(16,17) (18,19)(20,33)(21,32)(22,23)(24,25)(28,29)(30,31)(34,35) , • M5(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,36)(14,27)(15,26)(16,17) (18,19)(20,21)(22,35)(23,34)(24,25)(28,29)(30,31)(32,33) , • M6(η) = (1,2,3,4,5,6,7,8,9,10,11,12)(13,24)(14,27)(15,26)(16,17) (18,19)(20,21)(22,23)(25,36)(28,29)(30,31)(32,33)(34,35) . Using the computer algebra system SageMath [10], we determined that |⟨M1(ξ),M1(η)⟩| = 42467328 = 219 · 34. Moreover, M1(ξ), M1(η) and M1(ξη) respectively have orders 6, 12 and 12, so the Euler characteristic of the underlying surface of D̃1 is χ = |⟨M1(ξ),M1(η)⟩| · ( 1 ordM1(ξ) + 1 ordM1(η) + 1 ordM1(ξη) − 1) = −28311552 = −220 · 33, and its genus is g = 1− χ2 = 14155777. We will now show that D̃1 is not isomorphic to D̃2, . . . , D̃6. We claim that ω := [ξ−1η2ξ, ξη] ∈ kerM1 \ ⋃ 2≤k≤6 kerMk, thus concluding the proof. Indeed, we obtain: • M1(ω) = id; • M2(ω) = (13, 25)(15, 27)(21, 33)(23, 35); • M3(ω) = (17, 29)(21, 33); • M4(ω) = (13, 25)(15, 27)(19, 31)(21, 33); • M5(ω) = (13, 25)(17, 29); • M6(ω) = (13, 25)(19, 31)(21, 33)(23, 35). We have thus constructed a regular dessin D̃ of degree 219 · 34 and genus 14155777 with moduli field Q(ζ3, 3 √ 3). ORCID iDs Fumiharu Kato https://orcid.org/0009-0002-4800-0029 292 Ars Math. Contemp. 24 (2024) #P2.05 / 273–292 References [1] M. D. E. Conder, G. A. Jones, M. Streit and J. Wolfart, Galois actions on regular dessins of small genera, Rev. Mat. Iberoam. 29 (2013), 163–181, doi:10.4171/rmi/717, https://doi. org/10.4171/rmi/717. [2] E. Girondo and G. González-Diez, Introduction to compact Riemann surfaces and dessins d’enfants, volume 79 of London Mathematical Society Student Texts, Cambridge University Press, Cambridge, 2012. [3] G. González-Diez and A. Jaikin-Zapirain, The absolute Galois group acts faithfully on regular dessins and on Beauville surfaces, Proc. Lond. Math. Soc. (3) 111 (2015), 775–796, doi:10. 1112/plms/pdv041, https://doi.org/10.1112/plms/pdv041. [4] A. Grothendieck, Esquisse d’un programme, in: Geometric Galois actions, 1, Cambridge Univ. Press, Cambridge, volume 242 of London Math. Soc. Lecture Note Ser., pp. 5–48, 1997, with an English translation on pp. 243–283. [5] P. Guillot, An elementary approach to dessins d’enfants and the Grothendieck-Teichmüller group, Enseign. Math. 60 (2014), 293–375, doi:10.4171/lem/60-3/4-5, https://doi.org/ 10.4171/lem/60-3/4-5. [6] M. Herradón Cueto, An explicit quasiplatonic curve with non-abelian moduli field, Rev. Mat. Complut. 29 (2016), 725–739, doi:10.1007/s13163-016-0196-z, https://doi.org/10. 1007/s13163-016-0196-z. [7] R. A. Hidalgo and S. Quispe, Regular dessins d’enfants with field of moduli Q( p √ 2), Ars Math. Contemp. 13 (2017), 323–330, doi:10.26493/1855-3974.1202.9c1, https://doi. org/10.26493/1855-3974.1202.9c1. [8] G. A. Jones and J. Wolfart, Dessins d’enfants on Riemann surfaces, Springer Monographs in Mathematics, Springer, Cham, 2016, doi:10.1007/978-3-319-24711-3, https://doi. org/10.1007/978-3-319-24711-3. [9] S. K. Lando and A. K. Zvonkin, Graphs on surfaces and their applications, volume 141 of Encyclopaedia of Mathematical Sciences, Springer-Verlag, Berlin, 2004, doi:10.1007/ 978-3-540-38361-1, with an appendix by Don B. Zagier, Low-Dimensional Topology, II, https://doi.org/10.1007/978-3-540-38361-1. [10] The Sage Developers, SageMath, the Sage Mathematics Software System (Version 9.0), 2020- 01-01, https://www.sagemath.org. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.06 / 293–315 https://doi.org/10.26493/1855-3974.2619.06c (Also available at http://amc-journal.eu) Generalized X-join of graphs and their automorphisms* Javad Bagherian †, Hanieh Memarzadeh Department of Pure Mathematics, Faculty of Mathematics and Statistics, University of Isfahan, P.O. Box: 81746-73441, Isfahan, Iran Received 4 May 2021, accepted 6 April 2023, published online 27 September 2023 Abstract In this paper, we first introduce a new product of finite graphs as a generalization of the X-join of graphs. We then give necessary and sufficient conditions for a graph to be isomorphic to a generalized X-join. As a main result, we give necessary and sufficient conditions under which the full automorphism group of a generalized X-join is equal to the generalized wreath product of the automorphism groups of its factors. Keywords: Automorphism, generalized wreath product, graph, lexicographic product, permutation group, X-join. Math. Subj. Class. (2020): 05C25, 20B25, 20E22 1 Introduction One of the main problems in the theory of graphs, known as the König problem, asks for a concrete characterization of all automorphism groups of graphs. In particular, the problem of computing a generating set of the automorphism group is equivalent to the graph isomor- phism problem [9]. The automorphism groups of many graphs can be expressed in terms of the automorphism groups of their subgraphs. For instance, in most cases the automorphism groups of the graphs which are the lexicographic product of graphs are expressed in terms of the automorphism groups of their factors. The lexicographic product of graphs is one of the important products of graphs, defined by Harary in [7]. Sabidussi in [11] showed that under some conditions the automorphism group of the lexicographic product of two graphs *The authors are grateful to the anonymous referees, whose comprehensive reports helped to improve the quality of this paper. †Corresponding author. E-mail addresses: bagherian@sci.ui.ac.ir (Javad Bagherian), h.memarzadeh.762@sci.ui.ac.ir (Hanieh Memarzadeh) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 294 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 Γ and Γ′ can be expressed as the wreath product of the automorphism groups of Γ and Γ′. An important generalization of the lexicographic product is the X-join. It was introduced by Sabidussi as the graph formed from a given graph Γ = (V,R) by replacing every vertex v of Γ by a graph Bv and joining the vertices of Bv with those of Bu whenever uv ∈ R [11]. Note that the graphs Bv , v ∈ V , need not be mutually isomorphic. Hemminger in [8] gave necessary and sufficient conditions for the automorphism group of the X-join of graphs {Bv}v∈V to be the natural ones, i.e., those that are obtained by first permuting the graphs Bv , v ∈ V , according to a permutation of subscripts by an automorphism of Γ and then performing an arbitrary automorphism of each Bv . Note that Hemminger did not determine the structure of the automorphism group of the X-join of {Bv}v∈V in terms of automorphism groups of Bv , v ∈ V . It should be mentioned that the above results have been generalized to directed color graphs in [3]. If for a color digraph C = (V,R) and a collection of color digraphs {Dc | c ∈ V }, each vertex c of C is replaced by a copy of Dc and all possible arcs of color k from Dc to Dc′ are included, if and only if there is an arc of color k from c to c′ in C, we get the C-join of these color digraphs. The wreath product of two color digraphs C and D is the C-join of {Dc | c ∈ V } where Dc ∼= D for every c ∈ V . In [3], all automorphism groups of digraphs that can be written as a wreath product have been determined. In this paper we first give a generalization of the X-join of graphs (see Definition 2.1). This generalization, as a new operation on finite graphs, is a natural generalization of the X-join of graphs (a more algebraic way was considered by Weisfeiler [12, page 45] as the wreath product of a family of stable graphs with another stable graph). Also this new graph product generalizes the generalized wreath product of circulant digraph which de- fined in [2] (see Remark 2.9). It is also closely related with the wedge product of association schemes introduced and studied in [10] (see Remark 2.8). In Section 2 we give necessary and sufficient conditions under which a graph is isomorphic to a generalized X-join (see Theorem 2.4). But the main result of this paper deals with the connections between the au- tomorphism group of a generalized X-join and the automorphism groups of its factors. For computing the automorphism group of the generalized X-join of graphs, we need a general- ization of the wreath product of permutation groups. Recently, such a generalization, called the generalized wreath product, has been given in [1, 5]. We first show that under some conditions the automorphism group of the generalized X-join of graphs contains the gen- eralized wreath product of the automorphism groups of their factors (Theorem 4.1). As a main result, we then give necessary and sufficient conditions under which the full automor- phism group of the generalized X-join of graphs is equal to the generalized wreath product of the automorphism groups of their factors (Theorem 4.2). In particular, we determine the structure of the natural automorphism group of the X-join of graphs (Corollary 4.7). Terminology and notation: Throughout this paper, by a graph Γ = (V,R) we mean a finite undirected graph without multiple edges with the vertex set V = V (Γ) and the edge set R = E(Γ). We denote the complement of Γ by Γ. If all pairs of vertices of a subgraph Γ′ of Γ that are adjacent in Γ are also adjacent in Γ′, then Γ′ is an induced subgraph. For X ⊆ V we write Γ[X] for the subgraph of Γ induced by X and we also denote by Γ(X) the graph with vertices X and edge set E(Γ[X]) ∪ {(x, x) | x ∈ X}. For two graphs Γ = (V,R) and Γ′ = (V ′, R′), by a graph homomorphism f : Γ → Γ′ we mean a mapping f : V → V ′ such that (f(u), f(v)) ∈ R′ whenever (u, v) ∈ R. In the case when f : V → V ′ is surjective, f : Γ → Γ′ is called a graph epimorphism. More- J. Bagherian: Generalized X-join of graphs and their automorphisms 295 over, if f : V → V ′ is a bijection and f−1 : Γ′ → Γ is also a graph homomorphism, then f : Γ → Γ′ is called a graph isomorphism. Two graphs Γ and Γ′ are called isomorphic if there exists a graph isomorphism between Γ and Γ′. In this case we write Γ ≃ Γ′. When Γ = Γ′ every graph isomorphism f : Γ → Γ is called a graph automorphism of Γ. The set of all graph automorphisms of Γ is denoted by Aut(Γ) and is called the automorphism group of Γ. If Π is a partition of the vertices of a graph Γ, then the quotient graph Γ/Π is a graph with vertex set Π, for which distinct classes X,X ′ ∈ Π are adjacent if some vertex in X is adjacent to a vertex of X ′. Let Γ = (V,R) be a graph. The X-join of a set of graphs {Bx = (Yx, Ex) | x ∈ V } with Γ, denoted by Γ[Bx]x∈V , is a graph W = (Y,E) where Y = ⋃̇ x∈V Yx and E = {(yx, y′x′) ∈ Yx × Yx′ | (x, x′) ∈ R, or else x = x′ and (yx, y′x) ∈ Ex}. If B = (Y ′, E′) and Bx = B for every x ∈ V , we can identify ⋃̇ x∈V Yx with Y ′ × V and then the X-join of {Bx = (Yx, Ex) | x ∈ V } is the lexicographic product of Γ and B and is denoted by Γ ◦B. We denote by Kn a complete graph with n vertices. For the graph theoretical terminol- ogy and notation that are not defined here, we refer the reader to [6]. For a finite set V , we denote by Sym(V ) the group of all permutations of V . Every subgroup of Sym(V ) is called a permutation group on V . For F ≤ Sym(V ) and ∆ ⊆ V , the setwise stabilizer of ∆ in F is F{∆} = {f ∈ F | ∆f = ∆} and the pointwise stabilizer of ∆ in F is F(∆) = {f ∈ F | xf = x, ∀x ∈ ∆}. We say that two permutation groups F ≤ Sym(V ) and F ′ ≤ Sym(V ′) are permutation isomorphic if there exist a bijection λ : V → V ′ and a group isomorphism η : F → F ′ such that for every f ∈ F and v ∈ V we have λ(vf ) = λ(v)η(f). By a system of blocks Π for a permutation group F ≤ Sym(Ω) we mean (1) Π is a partition of Ω; (2) for every ∆ ∈ Π and every f ∈ F , ∆f ∩∆ = ∅ or ∆f = ∆. If Π is a system of blocks for F and ∆ ∈ Π, by F∆ we mean the group induced by the action of F{∆} on ∆. Then F∆/F(∆) ≤ Sym(∆) is a permutation group. 2 A generalization of the X-join of graphs In this section we first introduce a new product of graphs, called the generalized X-join of graphs. Then we give necessary and sufficient conditions under which a graph is isomor- phic to a generalized X-join. Definition 2.1. Let Γ = (V,R) be a graph and Π be a partition of V . Suppose that for every X ∈ Π we are given a graph BX = (YX , EX) and a graph epimorphism πX : YX → X from BX onto Γ(X). Put Y = ⋃̇ X∈ΠYX and π = ⋃̇ X∈ΠπX where for every y ∈ YX , π(y) := πX(y). We define a graph W with vertex set Y and edge set E such that (y, y′) ∈ E if and only if (1) either (y, y′) ∈ EX , for some X ∈ Π; (2) or (y, y′) ∈ π−1X (x)× π −1 X′ (x ′) where X ̸= X ′ and (x, x′) ∈ R. 296 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 We call the graph W = (Y,E) the generalized X-join of Γ and {BX}X∈Π with respect to π, and we denote it by Γ ◦π {BX}X∈Π. (See Figure 1.) Figure 1: The generalized X-join of Γ and {BX}X∈Π. In the following we show that the X-join of graphs is a special case of the generalized X-join of graphs. Example 2.2. Let Γ = (V,R) be a graph and Π be a partition of V such that for every X ∈ Π, X = {x} for some x ∈ V . Suppose that {Bx = (Yx, Ex) | x ∈ V } is a set of graphs. Define a graph epimorphism πx : Yx → X from Bx onto Γ(X) such that πx(yx) = x for every yx ∈ Yx. Then the generalized X-join of Γ and {Bx}x∈Π with respect to π = ⋃̇ x∈V πx is a graph with vertices Y = ⋃̇ x∈V Yx and the edge set E such that (yx, y′x′) ∈ E if and only if (1) either x = x′ and (yx, y′x) ∈ Ex; (2) or x ̸= x′ and (x, x′) ∈ R. One can see that in this case Γ ◦π {Bx}x∈V = Γ[Bx]x∈V , the X-join of graphs {Bx}x∈V . Example 2.3. Let Γ = (V,R) be the graph in Figure 2. Consider the partition Π = {X,X ′, X ′′} of V where X = {1, 2}, X ′ = {3, 4}, and X ′′ = {5, 6}. Suppose that BX = (YX , EX), BX′ = (YX′ , EX′), and BX′′ = (YX′′ , EX′′) are the graphs in Figure 2 with vertices YX = {a, b, c}, YX′ = {d, e, f}, and YX′′ = {g, h, i, k}, respectively. Now define the graph epimorphisms πX : BX → Γ(X), πX′ : BX′ → Γ(X ′), and πX′′ : BX′′ → Γ(X ′′) as follows:{ πX(a) = πX(b) = 1 πX(c) = 2{ πX′(e) = πX′(d) = 3 πX′(f) = 4 J. Bagherian: Generalized X-join of graphs and their automorphisms 297 2 1 3 4 6 5 Γ c b a BX e d f BX′ hg i k BX′′ Figure 2: Graph Γ and set of graphs {BX}X∈Π. { πX′′(g) = πX′′(h) = 5 πX′′(i) = πX′′(k) = 6. Then the generalized X-join of Γ and {BX , BX′ , BX′′} with respect to π is the graph in Figure 3. c b a e d f h g i k Figure 3: Graph W = Γ ◦π {BX , BX′ , BX′′}. Let Γ = (V,R) be a graph and let A,B ⊆ V . We say that A is externally related with respect to B, if every vertex v ∈ B that is adjacent to at least one element in A is adjacent to all vertices of A. Moreover, if B is also externally related with respect to A, we say that A and B are externally related to each other. Suppose that W = (Y,E) is the generalized X-join of Γ = (V,R) and {BX = (YX , EX) | X ∈ Π} with respect to π. Then we can define two equivalence relations E0 and E1 on Y as follows: (u, v) ∈ E0 ⇔ u, v ∈ π−1X (x), for some X ∈ Π and x ∈ X; (2.1) (u, v) ∈ E1 ⇔ u, v ∈ YX , for some X ∈ Π. (2.2) Clearly, E0 ⊆ E1. In the following we give a characterization of the generalized X-join of graphs in terms of the equivalence relations E0 and E1. Theorem 2.4. A graph W = (Y,E) is a generalized X-join of graphs if and only if there exist two equivalence relations E0 and E1 on Y such that 298 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 (i) E0 ⊆ E1; (ii) for every equivalence class P of E0 which is contained in a equivalence class Q of E1, P is externally related with respect to every equivalence class of E0 which is not in Q. Proof. Suppose that W = (Y,E) is the generalized X-join of Γ and {BX}X∈Π with re- spect to π. Then as we saw above, there are two equivalence relations E0 and E1 on Y such that E0 ⊆ E1. Since for every x ∈ X and x′ ∈ X ′ where X ̸= X ′, π−1X (x) and π−1X′ (x ′) are externally related to each other, it follows that condition (ii) holds. Now suppose that there exist two equivalence relations E0 and E1 on Y such that conditions (i) and (ii) hold. Let Y/E0 and Y/E1 be the sets of the equivalence classes of E0 and E1 on Y , respectively. Let Γ be the quotient graph W/E0. Moreover, for every U ∈ Y/E1, let U0 be the equivalence classes of E0 which are contained in U and BU0 be the subgraph of W induced by U . Since E0 ⊆ E1, {U0 | U ∈ Y/E1} gives a partition Π on Y/E0. Then for every U ∈ Y/E1 we can define a graph epimorphism πU0 from the graph BU0 onto Γ(U0). Suppose that W ′ is the generalized X-join of Γ and {BU0}U0∈Π with respect to π. Then V (W ′) = Y and it follows from condition (ii) that the set of edges ofW andW ′ are the same. ThusW =W ′ and soW is a generalized X-join of graphs. Remark 2.5. The following example shows that unlike the X-join of graphs, a graph can be represented as a generalized X-join of graphs, but not a unique way. This means that if W and W ′ are two isomorphic generalized X-join of graphs then it is not necessarily true that the factors of W and W ′ are isomorphic. Example 2.6. Consider the graph W = (Y,E) in Figure 4. If we consider two equiva- lence relations E0 ⊆ E1 such that Y/E0 = {{a}, {b, c}, {d, e, f, g}, {h}} and Y/E1 = {{a, b, c}, {d, e, f, g, h}}, then one can see that conditions (i) and (ii) of Theorem 2.4 hold. So it follows from Theorem 2.4 that W = Γ ◦π {BX , BX′} where the graphs Γ, BX and BX′ are shown in Figure 5. On the other hand, consider the graphs Γ′, B′X and B ′ X′ that are shown in Figure 6. Set Π′ = {X = {1, 2, 3}, X ′ = {4, 5}} and define the graph epimorphisms π′X : B ′ X → Γ′(X) by  a −→ 1 b, c −→ 2 h −→ 3 and π′X′ : B ′ X′ → Γ′(X ′) by { d, e −→ 4 f, g −→ 5 . Then one can see that W is the generalized X-join Γ′ ◦π′ {B′X , B′X′} with respect to π′. The following lemma that gives a sufficient condition under which two generalized X-join are isomorphic, is straightforward and therefore left to the reader. Lemma 2.7. Let W = Γ ◦π {BX}X∈Π and W ′ = Γ′ ◦π′ {BX′}X′∈Π′ . Suppose that the following conditions hold. J. Bagherian: Generalized X-join of graphs and their automorphisms 299 a b h c d f g e Figure 4: Graph W . 2 1 3 4 Γ a b c BX d f g e h BX′ Figure 5: Graph Γ and set of graphs {BX , BX′}. (1) There exists a graph isomorphism α : Γ → Γ′ which maps every partition class X ∈ Π onto a partition class X ′ ∈ Π′; (2) For every X ∈ Π, there exist graph isomorphisms βXX′ : BX → BX′ with X ′ = Xα such that the following diagram is commutative. YX βXX′−−−−→ YX′ πX y yπX′ X α−−−−→ X ′ Then ψ : ⋃̇ X∈ΠYX → ⋃̇ X′∈Π′YX′ defined by yX → βXX′(yX) is a graph isomorphism between W and W ′. Remark 2.8. The generalized X-join is closely related with the wedge product of asso- ciation schemes. The wedge product of association schemes which provides a way to construct new association schemes from old ones has been given in [10]. In the following we give the relationship between the relations of a wedge product of symmetric association schemes and the generalized X-join. 300 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 1 2 3 5 4 Γ′ a b h c B′X d f g e B′X′ Figure 6: Graph Γ′ and set of graphs {B′X , B′X′}. Suppose (V,G) is an association scheme and E is an equivalence relation on V such that it is a union of some relations R0, R1, . . . , Rt of G. Put D = {R0, R1, . . . , Rt} and suppose Σ is the set of equivalence classes of E. For every X ∈ Σ, let DX = {gX | g ∈ D} where gX = g ∩X ×X . Moreover, assume that (1) there is a set of association schemes {(YX , BX) | X ∈ Σ} such that all YX are pairwise disjoint and for every X ∈ Σ there exists a scheme normal epimorphism πX : YX ∪BX → X ∪DX . (2) for every X,X ′ ∈ Σ, there exists an algebraic isomorphism φXX′ : BX → BX′ such that the diagram BX φXX′−−−−→ BX′ πX y yπX′ DX εXX′−−−−→ DX′ is commutative, where εXX′(gX) = gX′ . Put Y := ⋃̇ X∈ΣYX , π := ⋃̇ X∈ΣπX and for every b ∈ BX , b̃ = ⋃ X′∈Σ φXX′(bX). Moreover, for every g ∈ G put g = ⋃ (x,x′)∈g∩X×X′, X,X′∈Σ,X ̸=X′ ψ−1X (x)× ψ −1 X′ (x ′). Fix Z ∈ Σ. Put B̃Z = {b̃ | b ∈ BZ}. Then it follows from [10, Theorem 2.2] that the pair (Y, B̃Z ∪ (G \ D)) is an association scheme, is called the wedge product of (YX , BX), X ∈ Σ, and (V,G). Now let g ∈ G\D and bX ∈ BX such that πX(bX) = gX . Then one can see that the graph with vertices Y and the edge set g ∪ b̃ is the generalized X-join of g and {φXX′(bX)}X′∈Σ with respect to π. Remark 2.9. The generalized wreath product of Cayley digraphs on abelain groups was first introduced in [2] and an entire section of the recent book [4, Section 5] is devoted to their study. A Cayley digraph Cay(G,S) of G with connection set S is a generalized wreath product if there are subgroups 1 < K ≤ L < G such that S \L is a union of cosets of K. In the following we show that the generalized X-join generalizes the generalized wreath product. To see this, let Cay(G,S) be a generalized wreath product on abelian J. Bagherian: Generalized X-join of graphs and their automorphisms 301 group G such that 1 ̸∈ S and S = S−1. Let V = {g1, . . . , gt} be a set of left coset representatives of K in G and Γ be the subgraph of G induced on V . Suppose that {a0 = 1, a1, . . . , am} is a set of left coset representatives of L in G. For every 0 ≤ i ≤ m, let Xi = {gj ∈ V | gj ∈ aiL}. Then Π = {X0, X1, . . . , Xm} is a partition of V . Put B0 = Cay(L,L ∩ S) and for every 1 ≤ i ≤ m, let Bi = ϕi(B0) where ϕi : B0 → Bi is a graph isomorphism defined by ϕi(l) = ail for every l ∈ L. Then there is the graph epimorphism πi : Bi → Γ(Xi) such that πi(gjK) = gj . Now let W = (G,E) be the generalized X-join of Γ and {B0, B1, . . . , Bm} with respect to π where π = ⋃̇m i=0πi. We show that for every x, y ∈ G, xy ∈ E if and only if xy−1 ∈ S. Clearly, if x, y ∈ aiL, then xy ∈ E if and only if xy−1 ∈ S ∩ L. If x ∈ grK ⊆ aiL and y ∈ gsK ⊆ ajL where i ̸= j, then xy−1 ∈ grg−1s K. Then xy ∈ E if and only if grg−1s ∈ S \ L if and only if grg −1 s K ⊆ S \ L, since S \ L is a union of cosets of K. So in this case xy ∈ E if and only if xy−1 ∈ S \L. Thus we conclude that W = (G,E) = Cay(G,S). This means that Cay(G,S) is a generalized X-join. 3 Generalized wreath product, definition and construction The generalized wreath product of permutation groups has been defined in [1, 5]. Since in the next section we need to construct the generalized wreath product of the automorphism group of graphs, here we have a look at the definition of this product which has been given in [1]. Let Γ = (V,R) be a graph and F = Aut(Γ). Suppose that Π is a system of blocks for F . Moreover, suppose that we are given a set of graphs {BX = (YX , EX) | X ∈ Π} such that the following conditions hold. (G1) If for some f ∈ F , Xf = X ′, then BX ≃ BX′ , (G2) If ∆ is an orbit of F on Π, then for some X ∈ ∆, there exists a graph epimorphism πX : YX → X fromBX onto Γ(X) and there exists an epimorphism ηX : Aut(BX) → FX/F(X) such that πX(y l) = (πX(y)) ηX(l), ∀y ∈ YX , l ∈ Aut(BX). By condition (G1), if there exists fXX′ ∈ F such that XfXX′ = X ′, we have a graph isomorphism ϕXX′ : YX → YX′ from graph BX onto BX′ . Then ψXX′ : Aut(BX) → Aut(BX′) defined by ψXX′(α) = ϕXX′αϕ −1 XX′ , ∀ α ∈ Aut(BX), is an isomorphism from Aut(BX) onto Aut(BX′). Moreover, by condition (G2), ΛX = {π−1X (x) | x ∈ X} is a system of blocks for Aut(BX), ηX : Aut(BX)/KX → FX/F(X) is an isomorphism, and Aut(BX)/KX ≤ Sym(ΛX) and FX/F(X) ≤ Sym(X) are per- mutation isomorphic where KX = ker(ηX). Lemma 3.1. Let X ′ ∈ ∆ with X ′ ̸= X . Then 302 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 (1) there exists a graph epimorphism πX′ from BX′ onto Γ(X ′) such that the following diagram is commutative, where ξXX′ : X → X ′, given by ξXX′(x) = xfXX′ for every x ∈ X. YX ϕXX′−−−−→ YX′ πX y yπX′ X ξXX′−−−−→ X ′ (2) there exists an epimorphism ηX′ : Aut(BX′) → FX ′ /F(X′) such that the follow- ing diagram is commutative, where ρXX′ : FX/F(X) → FX ′ /F(X′), defined by ρXX′(hF(X)) = fXX′hf −1 XX′F(X′) for every hF(X) ∈ FX/F(X). Aut(BX) ψXX′−−−−→ Aut(BX′) ηX y yηX′ FX/F(X) ρXX′−−−−→ FX′/F(X′) Proof. (1) If we define πX′ = ξXX′πXϕ−1XX′ , then πX′ : YX′ → X ′ is a graph epimor- phism fromBX′ onto Γ(X ′) such that the diagram mentioned above is commutative. (2) Define ηX′ = ρXX′ηXψ−1XX′ . Then ηX′ : Aut(BX′) → FX ′ /F(X′) is an epimor- phism such that the above diagram is commutative. Now suppose that a graph Γ = (V,R) and a set of graphs {BX = (YX , EX) | X ∈ Π} satisfy conditions (G1) and (G2). Set Y = ⋃̇ X∈ΠYX . Since for every X ∈ Π, KX ≤ Aut(BX) it follows that the action of ∏ X∈ΠKX on Y defined by yk := ykX , y ∈ YX , k = ∏ X∈Π kX ∈ K is faithful. Set K = ∏ X∈ΠKX . Then K ≤ Sym(Y ). Moreover, every element of F can be also considered as an element of Sym(Y ). In fact, for every g ∈ F we can associate g ∈ Sym(Y ). To do this, let yX ∈ YX and TX be a set of left coset representatives for KX in Aut(BX) such that idYX ∈ TX . Let g ∈ F . We associate to g an element g ∈ Sym(Y ) as follows: (i) if Xg = X , then (yX)g = (yX)t where ηX(tKX) = gF(X) for some t ∈ TX ; (ii) if Xg = X ′, then (yX)g = ϕXX′((yX)t) where ηX(tKX) = f−1XX′gF(X) for some t ∈ TX . Set F = {g | g ∈ F}. Clearly, F ⊆ Sym(Y ) and ⟨K,F ⟩ ≤ Sym(Y ). According to [1, Definition 2.1], the permutation group ⟨K,F ⟩, is the generalized wreath product of {Aut(BX)}X∈Π and F . We denote it by F ◦ {Aut(BX)}X∈Π. Remark 3.2. It should be mentioned that the generalized wreath product of {Aut(BX)}X∈Π and F is independent of the choice of representatives TX for every X ∈ Π. To see this, let yX ∈ YX and T ′X be a set of left coset representatives for KX in Aut(BX) such that J. Bagherian: Generalized X-join of graphs and their automorphisms 303 idYX ∈ T ′X and T ′X ̸= TX . Let g ∈ F and ĝ be an element of Sym(Y ) associated with g by the above argument. If Xg = X , then (yX)ĝ = (yX)t ′ where ηX(t′KX) = gF(X) for some t′ ∈ T ′X . Since t′ = tkX , for some t ∈ TX and kX ∈ KX , we have (yX) ĝ = (yX) t′ = (yX) tkX = (yX) gkX . Similarly, if Xg = X ′, then (yX)ĝ = ϕXX′((yX)t ′ ) where ηX(t′KX) = f−1XX′gF(X) for some t′ ∈ T ′X . If t′ = tkX for some t ∈ TX and kX ∈ KX , then (yX) ĝ = ϕXX′((yX) t′) = ϕXX′((yX) tkX ) = ϕXX′((yX) gkX ). Then we conclude that < F̂ ,K >=< F,K >, where F̂ = {f̂ | f ∈ F}. This shows that F ◦ {Aut(BX)}X∈Π =< F̂ ,K >. Example 3.3. Let Γ and {BX , BX′ , BX′′} be the graphs in Figure 7. Then Aut(Γ) = {idV , (14)(25)(36), (23), (56), (23)(56), (2635)(14), (2536)(14), (26)(35)(14)} and Π = {X = {2, 3}, X ′ = {5, 6}, X ′′ = {1, 4}} is a system of blocks for F = Aut(Γ). Put fXX′ = (14)(25)(36). Since XfXX′ = X ′ we have the following graph isomorphism from BX onto BX′ . ϕXX′ : YX → YX′ a −→ a′ b −→ b′ c −→ c′ d −→ d′ So condition (G1) holds, because, {X,X ′} and {X ′′} are the orbits of F on Π. More- over, there exist the graph epimorphisms πX : BX → Γ(X), πX′ : BX′ → Γ(X ′), and πX′′ : BX′′ → Γ(X ′′) such that {π−1X (x) | x ∈ X} = {{a, c}, {b, d}}, {π −1 X′ (x) | x ∈ X ′} = {{a′, c′}, {b′, d′}}, and {π−1X′′(x) | x ∈ X ′′} = {{b′′, c′′}, {a′′, d′′}}. If we de- fine the epimorphisms ηX : Aut(BX) → FX/F(X), ηX′ : Aut(BX′) → FX ′ /F(X′), and ηX′′ : Aut(BX′′) → FX ′′ /F(X′′) by{ ηX(idYX ) = F(X) ηX((ab)(cd)) = (23)F(X){ ηX′(idYX′ ) = F(X′) ηX′((a ′b′)(c′d′)) = (56)F(X′) and { ηX′′(idYX′′ ) = F(X′′) ηX′′((a ′′b′′)(c′′d′′)) = (14)(25)(36)F(X′′) 304 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 1 2 3 4 6 5 Γ ca b d BX c ′a′ b′ d′ BX′ c ′′a′′ b′′ d′′ BX′′ Figure 7: Graph Γ and set of graphs {BX}X∈Π. then it is easy to verify that condition (G2) holds. Put Y = {a, b, c, d, a′, b′, c′, d′, a′′, b′′, c′′, d′′}. Consider element g = (2536)(14) ∈ F . Since X = {2, 3}, X ′ = {5, 6}, and X ′′ = {1, 4} we have Xg = X ′, X ′g = X and X ′′ g = X ′′. Now we associate to g, an element g such that (i) (yX)g = ϕXX′(yX), since ηX(KX) = f−1XX′gF(X) = (56)F(X) = F(X); (ii) (yX′)g = ϕX′X((yX′)(a ′b′)(c′d′)), since ηX′((a′b′)(c′d′)KX′) = f−1X′XgF(X′) = (56)F(X′); (iii) (yX′′)g = (yX′′)(a ′′b′′)(c′′d′′), since ηX′′((a′′b′′)(c′′d′′)KX′′) = gF(X′′) = (14)(25)(36)F(X′′). Then g = (aa′bb′)(cc′dd′)(a′′b′′)(c′′d′′). Similarly, (1) if g = (23) then g = (ab)(cd); (2) if g = (56) then g = (a′b′)(c′d′); (3) if g = (23)(56) then g = (ab)(cd)(a′b′)(c′d′); (4) if g = (2635)(14) then g = (ab′ba′)(cd′dc′)(a′′b′′)(c′′d′′); (5) if g = (14)(25)(36) then g = (aa′)(bb′)(cc′)(dd′)(a′′b′′)(c′′d′′); (6) if g = (14)(26)(35) then g = (ab′)(ba′)(cd′)(dc′)(a′′b′′)(c′′d′′). Since KX , KX′ and KX′′ are trivial groups, it follows that Aut(Γ) ◦ {Aut(BX)}X∈Π = ⟨idY , (ab)(cd), (a′b′)(c′d′), (aa′bb′)(cc′dd′)(a′′b′′)(c′′d′′), (ab′ba′)(cd′dc′)(a′′b′′)(c′′d′′), (aa′)(bb′)(cc′)(dd′)(a′′b′′)(c′′d′′), (ab′)(ba′)(cd′)(dc′)(a′′b′′)(c′′d′′)⟩. 4 Automorphism group of the generalized X-join of graphs In this section we show that the automorphism group of some graphs which are isomorphic to a generalized X-join can be expressed in terms of the generalized wreath product of automorphism groups of its factors. J. Bagherian: Generalized X-join of graphs and their automorphisms 305 Theorem 4.1. With the notation above, suppose that a graph Γ = (V,R) and a set of graphs {BX = (YX , EX) | X ∈ Π} satisfy the conditions (G1) and (G2). Then Aut(Γ) ◦ {Aut(BX)}X∈Π ≤ Aut(Γ ◦π {BX}X∈Π). Proof. Let W = (Y,E) be the generalized X-join of Γ and {BX}X∈Π with respect to π, and H = ⟨K,F ⟩ be the generalized wreath product of {Aut(BX)}X∈Π and Aut(Γ). We show that for every h ∈ H and u, v ∈ Y , if (u, v) ∈ E, then (uh, vh) ∈ E. To do this, we assume that (u, v) ∈ E and we consider the following cases. Case 1. Suppose that h = ∏ X∈Π kX ∈ K. (i) If u, v ∈ YX for someX ∈ Π, then since for everyX ∈ Π, kX ∈ Aut(BX) we have (u, v)h = (uh, vh) = (ukX , vkX ) ∈ EX . (ii) If u ∈ YX and v ∈ YX′ for some X and X ′ in Π where X ̸= X ′, then (x, x′) = (πX(u), πX′(v)) ∈ R and since (ukX , vkX′ ) ∈ π−1X (x)×π −1 X′ (x ′) we have (u, v)h = (uh, vh) = (ukX , vkX′ ) ∈ E. Case 2. Suppose that h = g for some g ∈ F and u, v ∈ YX for some X ∈ Π. (i) If Xg = X , then since (u, v)g = (ug, vg) = (ut, vt) where ηX(tKX) = gF(X) for some t ∈ Aut(BX), we have (u, v)h = (uh, vh) = (ut, vt) ∈ E. (ii) If Xg = X ′ for some X ′ ∈ Π, then since (u, v)g = (ug, vg) = (ϕXX′(ut), ϕXX′(v t)) where ηX(tKX) = f−1XX′gF(X) for some t ∈ Aut(BX) we have (u, v)h = (uh, vh) = (ϕXX′(u t), ϕXX′(v t)) ∈ E. Case 3. Let h = g for some g ∈ F , u ∈ YX and v ∈ YX′ for some X,X ′ ∈ Π where X ̸= X ′. In this case since (x, x′) = (πX(u), πX′(v)) ∈ R and g ∈ Aut(Γ) we have (xg, x′ g ) ∈ R. Then the following cases arise. (i) If Xg = X and X ′g = X ′, then (u, v)g = (ug, vg) = (ut, vt ′ ) where ηX(tKX) = gF(X) and ηX′(t′KX′) = gF(X′). Since πX(ut) = πX(u)ηX(t) = xg and πX′(v t′) = πX′(v) ηX′ (t ′) = x′ g we have (ut, vt ′ ) ∈ π−1X (x g)× π−1X′ (x ′g). Then (uh, vh) = (ut, vt ′ ) ∈ E. (ii) If Xg = X and X ′g = X ′′, then (u, v)g = (ug, vg) = (ut, ϕX′X′′(vt ′ )) where ηX(tKX) = gF(X) and ηX′(t′KX′) = f −1 X′X′′gF(X′). Then πX(u t) = πX(u) ηX(t) = xg and by condition (G2) we have πX′′(ϕX′X′′(v t′)) = ξX′X′′(πX′(v t′)) = ξX′X′′(πX′(v) ηX′ (t ′)) = fX′X′′ηX′(t ′KX′)(πX′(v)) = gF(X′)(πX′(v)) = (πX′(v)) g = x′ g . 306 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 So (ut, ϕX′X′′(vt ′ )) ∈ π−1X (xg)×π −1 X′′(x ′g) and then (uh, vh) = (ut, ϕX′X′′(vt ′ )) ∈ E. (iii) If Xg = X ′ and X ′g = X ′′, then (u, v)g = (ug, vg) = (ϕXX′(ut), ϕX′X′′(vt ′ )) where ηX(tKX) = f−1XX′gF(X) and ηX′(t ′KX′) = f −1 X′X′′gF(X′). From condition (G2) we have πX′(ϕXX′(u t)) = ξXX′(πX(u t)) = ξXX′(πX(u) ηX(t)) = fXX′ηX(tKX)(πX(u)) = gF(X)(πX(u)) = (πX(u)) g = xg. Similarly, πX′′(ϕX′X′′(vt ′ )) = x′ g . Then (ϕXX′(u t), ϕX′X′′(v t′)) ∈ π−1X′ (x g)× π−1X′′(x ′g) and so (uh, vh) = (ϕXX′(ut), ϕX′X′′(vt ′ )) ∈ E. (iv) If Xg = X ′′ and X ′g = X ′′′, then (u, v)g = (ug, vg) = (ϕXX′′(ut), ϕX′X′′′(vt ′ )) where ηX(tKX) = f−1XX′′gF(X) and ηX′(t ′KX′) = f −1 X′X′′′gF(X′). Then an argu- ment similar to that given in (iii) shows that πX′′(ϕXX′′(ut)) = xg and πX′′′(ϕX′X′′′(v t′)) = x′ g . So (ϕXX′′(ut), ϕX′X′′′(vt ′ )) ∈ π−1X′′(xg) × π −1 X′′′(x ′g) and hence (uh, vh) = (ϕXX′′(ut), ϕX′X′′′(vt ′ )) ∈ E. Then we conclude that ⟨K,F ⟩ ⊆ Aut(Γ ◦π {BX}X∈Π). Thus Aut(Γ) ◦ {Aut(BX)}X∈Π ≤ Aut(Γ ◦π {BX}X∈Π). The inclusion Aut(Γ)◦{Aut(BX)}X∈Π ≤ Aut(Γ◦π{BX}X∈Π) in the above theorem may be proper. For example D8 = Aut(K2) ◦Aut(K2) < Aut(K2 ◦K2) = S4, where S4 is the symmetric group on 4 elements V = {1, 2, 3, 4} and D8 = {idV , (12), (34), (12)(34), (13)(24), (14)(23), (1324), (1423)} is the dihedral group of order 8; see [6, Chapter 10]. In the following we give necessary and sufficient conditions under which the above inclusion is proper. Theorem 4.2. With the notation above, suppose that the graph Γ = (V,R) and the set of graphs {BX = (YX , EX) | X ∈ Π} satisfy the conditions (G1) and (G2). Let W = (Y,E) be the generalized X-join of Γ and {BX}X∈Π with respect to π and let E0 ⊆ E1 be the equivalence relations defined in (1) and (2). Then the inclusion Aut(Γ) ◦ {Aut(BX)}X∈Π ≤ Aut(W ) is proper if and only if there exist equivalence relations E′0 ⊆ E′1 on Y such that the following conditions hold. J. Bagherian: Generalized X-join of graphs and their automorphisms 307 (i) E′0 ⊊ E0 and E′1 ⊊ E1, and W = Γ′ ◦π′ {AZ}Z∈Π′ , where the graph Γ′ is the quotient graph W/E′0, Π ′ is a partition of V (Γ′), {AZ}Z∈Π′ are the subgraphs of W induced by the equivalence classes of E′1, π ′ = ⋃̇ Z∈Π′π ′ Z , and {π′ −1 Z (x) | x ∈ V (Γ′), Z ∈ Π′} is the set of equivalence classes of E′0. (ii) There exist Z ̸= Z ′ ∈ Π′−Π and a graph isomorphism ϕ : YZ → YZ′ fromAZ onto AZ′ , such that ϕ preserves equivalence classes of E′0 contained in YZ . (iii) For every equivalence class S′ of E′0 contained in YZ , if W [S ′] is a union of con- nected components of W [S] for some S ∈ W/E0, where S′ ⊊ S then π′Z(S′) and π′Z′(ϕ(S ′)) are nonadjacent, otherwise π′Z(S ′) and π′Z′(ϕ(S ′)) are adjacent. In both cases π′Z(S ′) and π′Z′(ϕ(S ′)) have the same neighbors in V (Γ′) \ Z ∪ Z ′. (iv) For each two distinct equivalence classes S′1 , S ′ 2 of E ′ 0 that are contained in YZ , S ′ 1 and ϕ(S′2) are adjacent if and only if ϕ(S ′ 1) and S ′ 2 are adjacent. Proof. Suppose that there exists ϕ ∈ Aut(W ) \ Aut(Γ) ◦ {Aut(BX)}X∈Π. Let U be the set of all elements of Y that are moved by ϕ. Since Y/E1 = {YX | X ∈ Π} and Y/E0 = ⋃ X∈Π ΛX , where ΛX = {π −1 X (x) | x ∈ X}, are two system of blocks for Aut(Γ) ◦ {Aut(BX)}X∈Π, then there exist (1) X ∈ Π such that UX = U ∩ YX ̸= ∅; (2) X ′ ∈ Π such that X ̸= X ′ and ϕ(UX) ⊆ YX′ . Note that if X = X ′, then the restriction of ϕ to YX is an automorphism of BX and ϕ(UX) ⊆ YX . But since ϕ /∈ Aut(Γ) ◦ {Aut(BX)}X∈Π we must have at least two equivalence classes S1, S2 ∈ E0 such that a part of S1 is moved by the automorphism ϕ to a part of S2. This contradicts the fact that ΛX = {π−1X (x) | x ∈ X} is a system of blocks for Aut(BX). (3) at least one equivalence class S of E0 such that S ∩ UX ⊊ S. Indeed, if UX =⋃t i=1 Si ⊊ YX where every Si is an equivalence class of E0, then by (2), ϕ(UX) ⊆ YX′ for some X ′ ̸= X and ϕ(UX) is a union of some equivalence classes of E0 which are contained in YX′ . This means that the vertices πX(S1), . . . , πX(St) of X can be moved to the vertices πX′(ϕ(S1)), . . . , πX′(ϕ(St)) of X ′. This contradicts the fact that Π is a system of blocks for Aut(Γ). Put VX′ = ϕ(UX). Let S1, S2, ..., St be the equivalence classes of E0 contained in YX such that for every 1 ≤ i ≤ t, SXi = Si ∩ UX ̸= ∅. Then for at least one i, SXi ⊊ Si. Moreover, we have the following. (a) The restriction of ϕ to UX gives an isomorphism between W [UX ] and W [VX′ ], the subgraphs of W induced by UX and VX′ . (b) For each i the vertices in SXi ∪ ϕ(SXi ) have the same neighbors in Y \ (UX ∪ VX′). Indeed, suppose that u ∈ SXi and w is a neighbor of u. Suppose that T1, . . . , Tt are equivalence classes of E0 such that ϕ(SXi ) ∩ Ti ̸= ∅ and vi ∈ Ti \ ϕ(SXi ). If w ∈ Y \ (YX ∪ YX′), then ϕ(w) is adjacent to all vertices of Ti, specially vi. So w and ϕ−1(vi) = vi are adjacent. Thus w is adjacent to all vertices of ϕ(SXi ). 308 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 Moreover, if w ∈ YX \ UX , then since w is adjacent to u, ϕ(w) = w is adjacent to ϕ(u). Then w is adjacent to all vertices of ϕ(SXi ). Similarly, if w ∈ YX′ \ VX′ , then w is adjacent to all vertices of ϕ(SXi ). Hence we conclude that S X i ∪ ϕ(SXi ) have the same neighbors in Y \ (UX ∪ VX′). (c) If W [SXi ] is a union of connected components of W [Si], then S X i and ϕ(S X i ) are nonadjacent; otherwise by the definition of W , Si \ SXi and ϕ(SXi ) are adjacent and since ϕ ∈ Aut(W ), SXi and Si \SXi are adjacent and this contradicts the hypothesis that W [SXi ] is a union of connected components of W [Si]. Also if W [S X i ] and W [Si \ SXi ] are adjacent, then since ϕ ∈ Aut(W ), Si \ SXi and ϕ(SXi ) must be adjacent and the definition of W implies that ϕ(SXi ) and Si are externally related to each other. Moreover, Si \ SXi and SXi are also externally related to each other. (d) For two different equivalence classes S1 and S2 of E0 with SX1 , S X 2 ̸= ∅, if S1 and ϕ(SX2 ) are adjacent then from the definition of W it follows that S1 and ϕ(S X 2 ) are externally related to each other. Moreover, since ϕ ∈ Aut(W ) we must have ϕ(SX1 ) and S2 are also externally related to each other. Similarly, if S2 and ϕ(SX1 ) are adjacent then ϕ(SX2 ) and S1 are externally related to each other. Now we consider two equivalence relations E′0 ⊆ E′1 on Y such that Y/E′1 = {UX , VX′ , YX \ UX , YX′ \ VX′ , Y/E1 \ {YX , YX′}}, and the equivalence classes of E′0 are equal SXi , Si \ SXi , ϕ(SXi ), Ti \ ϕ(SXi ), 1 ≤ i ≤ t, and Y/E0 \ {Si, Ti | 1 ≤ i ≤ t}, where for every i, Ti is an equivalence class of E0 such that ϕ(SXi ) ⊆ Ti. From statements (b) and (c) we conclude that the condition (ii) of Theorem 2.4 holds and so W = Γ′ ◦π′ {AX}X∈Π′ , where Γ′ is the quotient graph W/E′0, Π′ is a partition of V (Γ′) induced by Y/E′1, and {AX}X∈Π′ are the subgraphs of W induced by the equivalence classes of E′1. So (i) holds. If we denote by AX and AX′ the subgraphs of W induced by UX and VX′ , respectively, then the restriction of ϕ to UX gives a graph isomorphism betweenAX andAX′ . Clearly, ϕ preserves the equivalence classes of E′0 contained in UX . Thus condition (ii) of theorem holds. Moreover, (b), (c) and the definition of π′ imply that condition (iii) holds. Finally, condition (iv) follows from statement (d). Conversely, suppose that there exist equivalence relations E′0 ⊆ E′1 on Y such that conditions (i) – (iv) hold. Let UZ and UZ′ be the vertex sets of AZ and AZ′ , respectively. Assume that ϕ : UZ −→ UZ′ is the graph isomorphism from AZ onto AZ′ . We define a bijection ψ : Y −→ Y as follows: ψ(v) =  ϕ(v) if v ∈ UZ , ϕ−1(v) if v ∈ UZ′ , v if v /∈ {UZ , UZ′}. We claim that ψ ∈ Aut(W ). To do this, we suppose that (u, v) ∈ E and we consider the following cases. J. Bagherian: Generalized X-join of graphs and their automorphisms 309 (1) If u, v ∈ Y \ UZ ∪ UZ′ , then clearly (ψ(u), ψ(v)) = (u, v) ∈ E. (2) If u, v ∈ UZ , then since ϕ is a graph isomorphism it follows that (ψ(u), ψ(v)) = (ϕ(u), ϕ(v)) ∈ E. Similarly, if u, v ∈ UZ′ we have (ψ(u), ψ(v)) = (ϕ−1(u), ϕ−1(v)) ∈ E. (3) If u ∈ UZ and v ∈ Y \ UZ ∪ UZ′ , then since v is a neighbor of u it follows from (iii) that v is also a neighbor of ϕ(u). Hence (ψ(u), ψ(v)) = (ϕ(u), v) ∈ E. (4) If u ∈ UZ and v ∈ UZ′ such that for some equivalence class S′ of E′0, u ∈ S′ and v ∈ ϕ(S′), then the definition of W implies that all vertices in S′ are adjacent to all vertices in ϕ(S′) and so (ψ(u), ψ(v)) = (ϕ(u), ϕ−1(v)) ∈ E. (5) If u ∈ UZ and v ∈ UZ′ such that for two equivalence classes S′1 and S′2 of E′0, u ∈ S′1 and v ∈ ϕ(S′2), then by the definition of W , all vertices in S′1 are adjacent to all vertices in ϕ(S′2). On the other hand, it follows from (iv) that ϕ(S ′ 1) and S ′ 2 are adjacent. So all vertices of ϕ(S′1) are adjacent to all vertices of S ′ 2 and thus (ψ(u), ψ(v)) = (ϕ(u), ϕ−1(v)) ∈ E. Hence ψ ∈ Aut(W ). SinceE′0 ⊊ E0 and Z,Z ′ ∈ Π′−Π, and ψ preserves the equivalence classes of E′0 we conclude that ψ ∈ Aut(W ) \Aut(Γ) ◦ {Aut(BX)}X∈Π. Example 4.3. Suppose that Γ is the graph in Figure 8 with vertices V = {1, 2, 3, 4, 5, 6}. Then one can see that F = Aut(Γ) = {idV , (12)(56)(34), (13)(24), (23)(56)(14)} and Π = {X = {1, 2}, X ′ = {3, 4}, X ′′ = {5, 6}}, is a system of blocks for F . Moreover, FX = FX ′ = {idV , (12)(56)(34)}, FX ′′ = F , F(X) = F(X′) = {idV }, and F(X′′) = {idV , (13)(24)}. Suppose that BX , BX′ , and BX′′ are the graphs in Figure 8 with vertices YX = {a, b, c, d}, YX′ = {a′, b′, c′, d′}, YX′′ = {a′′, b′′, c′′, d′′}, respectively. Now consider the graph epimorphisms πX : BX → Γ(X), πX′ : BX′ → Γ(X ′), and πX′′ : BX′′ → Γ(X ′′) as the following: 1 3 4 2 5 6 Γ a bc d BX d′ c′ b′ a′ BX′ a ′′ b′′ c′′ d′′ BX′′ Figure 8: Graph Γ and set of graphs {BX}X∈Π. { πX(a) = πX(b) = 1 πX(c) = πX(d) = 2 310 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 { πX′(a ′) = πX′(b ′) = 3 πX′(c ′) = πX′(d ′) = 4 and { πX′′(a ′′) = πX′′(b ′′) = 6 πX′′(c ′′) = πX′′(d ′′) = 5 . Since Aut(BX) = {idYX , (bc)(ad)}, Aut(BX′) = {idYX′ , (b ′c′)(a′d′)}, and Aut(BX′′) = {idYX′′ , (a ′′b′′), (c′′d′′), (a′′b′′)(c′′d′′), (a′′c′′b′′d′′), (a′′d′′b′′c′′) , (a′′c′′)(b′′d′′), (a′′d′′)(b′′c′′)}, we can define epimorphisms ηX : Aut(BX) → FX/F(X), ηX′ : Aut(BX′) → FX ′ /F(X′), and ηX′′ : Aut(BX′′) → FX ′′ /F(X′′) by{ ηX(idYX ) = idV ηX((bc)(ad)) = (12)(56)(34)F(X){ ηX′(idYX′ ) = idV ηX′((b ′c′)(a′d′)) = (12)(56)(34)F(X′) and ηX′′(idYX′′ ) = ηX′′((a ′′b′′)) = ηX′′((c ′′d′′)) = ηX′′((a ′′b′′)(c′′d′′)) = idV ηX′′((a ′′c′′b′′d′′)) = ηX′′((a ′′d′′b′′c′′)) = ηX′′((a ′′c′′)(b′′d′′))= ηX′′((a ′′d′′)(b′′c′′))= (12)(56)(34)F(X′′) Then KX and KX′ are trivial groups and KX′′ = {idYX′′ , (a ′′b′′), (c′′d′′), (a′′b′′)(c′′d′′)}. Let TX = {idYX , (bc)(ad)}, TX′ = {idYX′ , (b ′c′)(a′d′)}, and TX′′ = {idYX′′ , (a′′c′′)(b′′d′′)}. Put fXX′ = (13)(24). Then the elements (12)(56)(34), (13)(24) and (23)(56)(14) in Aut(Γ) are associated to (bc)(ad)(b′c′)(a′d′)(a′′c′′)(b′′d′′), (aa′)(bb′)(cc′)(dd′) and (bc′)(ad′)(cb′)(da′)(a′′c′′)(b′′d′′), respectively. Then we have Aut(Γ)◦{Aut(BX)}X∈Π=⟨idY , (aa′)(bb′)(cc′)(dd′), (bc′)(ad′)(cb′)(da′)(a′′c′′)(b′′d′′), (a′′b′′), (c′′d′′), (bc)(ad)(b′c′)(a′d′)(a′′c′′)(b′′d′′)⟩. Now letW = (Y,E) be the generalized X-join of Γ and {BX}X∈Π with respect to π. (See Figure 9.) Consider the equivalence relations E′0 and E ′ 1 on Y with the following classes, Y/E′0 = {{a}, {b}, {c}, {d}, {a′}, {b′}, {c′}, {d′}, {a′′, b′′}, {c′′, d′′}} Y/E′1 = {{a, d}, {b, c}, {a′, d′}, {b′, c′}, {a′′, b′′, c′′, d′′}}. Then one can see that J. Bagherian: Generalized X-join of graphs and their automorphisms 311 a b c d a′′ b′′ d′ c′ b′ a′ c′′ d′′ Figure 9: Graph W = Γ ◦π {BX}X∈Π. (1) Γ ◦π {BX}X∈Π = Γ′ ◦π′ {AZ}Z∈Π′ , where Γ′ is the quotient graph W/E′0 with vertices V (Γ′) = {1, 2, . . . , 10}, and {AZ}Z∈Π′ are the subgraphs of W induced by the equivalence classes of E′1, and π ′ = ⋃̇ Z∈Π′π ′ Z maps {a}, {b}, {c}, {d}, {a′}, {b′}, {c′}, {d′}, {c′′, d′′}, {a′′, b′′} onto 1, 2, . . . , 10, respectively. (See Figure 10.) (2) Put YZ = {b, c} and YZ′ = {b′, c′} and let AZ =W [YZ ] and AZ′ =W [YZ′ ]. Then ϕ : YZ → YZ′ such that ϕ(b) = b′ and ϕ(c) = c′ is a graph isomorphism from AZ onto AZ′ . Clearly, ϕ preserves the equivalence classes of E′0 contained in YZ . (3) YZ contains two equivalence classes S′1 = {b} and S′2 = {c} of E′0 such that S′1 ⊆ S1 and S′2 ⊆ S2 where S1 = {a, b} ∈ Y/E0 and S2 = {c, d} ∈ Y/E0. Moreover, W [S1] is connected and π′Z(b) and π ′ Z′(ϕ(b)) are adjacent and have the same neigh- bors in V (Γ′) \ {Z ∪Z ′}, where Z = {π′Z(b), π′Z(c)} and Z ′ = {π′Z′(b′), π′Z′(c′)}. Similarly, W [S2] is connected and π′Z(c) and π ′ Z′(ϕ(c)) are adjacent and have the same neighbors in V (Γ′) \ {Z ∪ Z ′}. (4) The vertex b is nonadjacent to ϕ(c) and vertex c is nonadjacent to ϕ(b). Then the conditions of Theorem 4.2 hold. So Aut(Γ) ◦ {Aut(BX)}X∈Π ⪇ Aut(Γ ◦π {BX}X∈Π). In the following as a main result, we give necessary and sufficient conditions under which the full automorphism group of the generalized X-join of graphs is equal to the generalized wreath product of the automorphism groups of their factors. 312 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 9 1 2 3 4 10 8 7 6 5 b c a d b′ c′ a′ d′ a′′ b′′ c′′ d′′ Figure 10: Graph Γ′ and set of graphs {AZ}Z∈Π′ . Corollary 4.4. Suppose that W = Γ ◦π {BX}X∈Π is such that the graph Γ = (V,R) and the set of graphs {BX = (YX , EX) | X ∈ Π} satisfy the conditions (G1) and (G2). Then Aut(Γ◦π {BX}X∈Π) = Aut(Γ)◦{Aut(BX)}X∈Π if and only if there are no equivalence relations E′0 ⊆ E′1 on Y satisfying the conditions (i), (ii), (iii), and (iv) of Theorem 4.2. Proof. This follows immediately from Theorem 4.2. Example 4.5. Let W = (Y,E) be the graph in Figure 11. It is easy to see that W is the graph Γ ◦π {BX , BX′ , BX′′} where Γ and {BX , BX′ , BX′′}, and π = πX ∪ πX′ ∪ πX′′ are given in Example 3.3. Moreover, Y/E0 = {{a, c}, {b, d}, {a′, c′}, {b′, d′}, {b′′, c′′}, {a′′, d′′}} and Y/E1 = {{a, c, b, d}, {a′, c′, b′, d′}, {b′′, c′′, a′′, d′′}}. Since there are no equivalence relations E′0 ⊆ E′1 on Y that satisfy the conditions (i), (ii), (iii), and (iv) of Theorem 4.2, it follows that Aut(W ) = Aut(Γ)◦{Aut(BX)}X∈Π = ⟨idY , (ab)(cd), (aa′bb′)(cc′dd′)(a′′b′′)(c′′d′′), (a′b′)(c′d′), (ab′ba′)(cd′dc′)(a′′b′′)(c′′d′′), (aa′)(bb′)(cc′)(dd′)(a′′b′′)(c′′d′′), (ab′)(ba′)(cd′)(dc′)(a′′b′′)(c′′d′′)⟩. The next corollary follows directly from Theorem 4.2. Corollary 4.6. Suppose that the graph Γ = (V,R) and the set of graphs {BX = (YX , EX) | X ∈ Π} satisfy the conditions (G1) and (G2). Let W = (Y,E) be the generalized X-join of Γ and {BX}X∈Π with respect to π and letE0 ⊆ E1 be the equivalence relations defined in (1) and (2). Then Aut(Γ) ◦ {Aut(BX)}X∈Π = Aut(Γ ◦π {BX}X∈Π) if W is uniquely determined by E0 and E1. Corollary 4.7 (See [8, Theorem 2.10]). Let Γ = (V,R) be a graph and {Bx | x ∈ V } be a set of graphs such that Bx ≃ Bx′ whenever xf = x′ for some f ∈ Aut(Γ). Then Aut(Γ[Bx]x∈V ) = Aut(Γ) ◦ {Aut(Bx)}x∈V if and only if J. Bagherian: Generalized X-join of graphs and their automorphisms 313 d′′ d b a c a′′ b′′ c′′ d′ b′ a′ c′ Figure 11: Graph W . (1) Bx is connected if there exists at least one vertex w ∈ V such that x and w are nonadjacent and have the same neighbors in V , (2) Bx is connected if there exists at least one vertex w ∈ V such that x and w are adjacent and have the same neighbors in V \ {x,w}. Proof. By Example 2.2, W = Γ[Bx]x∈V = Γ ◦π {Bx}x∈V where Π = {{x} | x ∈ V } and πx : Yx → X is a graph epimorphism from Bx onto Γ(X) such that πx(yx) = x for every yx ∈ Yx. In this case E0 = E1 and FX = F(X). If we define ηX := Aut(BX) → FX/F(X) by ηX(α) = 1FX/F(X) for every α ∈ Aut(BX), then ηX is an epimorphism and condition (G2) holds. Then it follows from Corollary 4.4 that Aut(Γ[Bx]x∈V ) = Aut(Γ) ◦ {Aut(Bx)}x∈V if and only if there is no equivalence relation E′0 on Y satisfying the conditions (i), (ii), (iii), and (iv) of Theorem 4.2. Now suppose that Aut(Γ[Bx]x∈V ) = Aut(Γ) ◦ {Aut(Bx)}x∈V and there exist x,w ∈ V such that x and w are nonadjacent and have the same neighbors in V . If Bx is discon- nected then Bw is also disconnected and we can define an equivalence relation E′0 on Y such that the equivalence classes of E′0 are Yz, z ̸∈ {x,w}, together with the connected components of Bx and Bw. Since x and w are nonadjacent and have the same neighbors in V , one can see that the conditions (i), (ii), (iii), and (iv) of Theorem 4.2 hold, a contradic- tion. So Bx is connected. Moreover, suppose that there exist x,w ∈ V such that x and w are adjacent and have the same neighbors in V \ {x,w}. If Bx is disconnected, then there exist at least two subsets S1, S2 ⊂ Yx such that all vertices of S1 are adjacent to all vertices of S2. Similarly, there exist subsets S′1, S ′ 2 ⊂ Yw with the property that all vertices of S′1 are adjacent to all vertices of S′2. Then we can define an equivalence relation E ′ 0 on Y such that S1, S2, S′1, and S ′ 2 together with Yz, z ̸∈ {x,w} are its equivalence classes. One can see that in this case the conditions (i), (ii), (iii), and (iv) of Theorem 4.2 hold and thus again we have a contradiction. 314 Ars Math. Contemp. 24 (2024) #P2.06 / 293–315 Conversely, suppose that conditions (1) and (2) hold and suppose on the contrary that φ ∈ Aut(Γ[Bx]x∈V ) \Aut(Γ) ◦ {Aut(Bx)}x∈V . Then there is an equivalence relation E′0 on Y satisfying the conditions (i), (ii), (iii), and (iv) of Theorem 4.2. It follows from condition (i) that W = Γ′ ◦π′ {Az}z∈V (Γ′), where the graph Γ′ is the quotient graphW/E′0 and {Az}z∈V (Γ′) are the subgraphs ofW induced by the equivalence classes of E′0. It follows from (ii) that there exist x,w ∈ V such that the equivalence classes of E′0 contain Yz ⊊ Yx and Yz′ = φ(Yz) ⊊ Yw. By (iii) if Bx[Yz] is a union of connected components of Bx, then z and z′ are nonadjacent and all of their neighbors are exactly the same in V (Γ′), otherwise z and z′ are adjacent and have the same neighbors in V (Γ′) \ {z, z′}. This implies that if Bx is disconnected then z and z′ are nonadjacent and all vertices in Yz and all vertices in Yz′ have the same neighbors in Y \ (Yz ∪ Yz′). Since Yz ⊊ Yx and Yz′ ⊊ Yw it follows that x and w must be nonadjacent and have the same neighbors in V , which contradicts (1). Moreover, if Bx is connected, since z and z′ are adjacent and have the same neighbors in V (Γ′) \ {z, z′} it follows that all vertices in Yz are adjacent to all vertices of Yz′ and all vertices in Yz and all vertices in Yz′ have the same neighbors in Y \ (Yz ∪ Yz′). Then all vertices in Yx are adjacent to all vertices of Yw. So x and w must be adjacent and have the same neighbors in V \ {x,w}. Furthermore, since all vertices in Yz are adjacent to all vertices of Yx \ Yz it follows that Bx is disconnected, which contradicts (2). Thus we have Aut(Γ[Bx]x∈V ) = Aut(Γ) ◦ {Aut(Bx)}x∈V . 5 Conclusion A generalization of the X-join of graphs has been introduced and necessary and sufficient conditions under which a graph is isomorphic to a generalized X-join has been given. A generating set for the automorphism groups of a class of graphs which are isomorphic to a generalized X-join has been computed. Since the generalized X-join of graphs is a natural generalization of the X-join of graphs, the results on the X-join or lexicographic product of graphs can be also studied for the generalized X-join of graphs. References [1] J. Bagherian, Schurity of the wedge product of association schemes and generalized wreath product of permutation groups, Discrete Math. 343 (2020), 10, doi:10.1016/j.disc.2020. 112084, id/No 112084, https://doi.org/10.1016/j.disc.2020.112084. [2] S. Bhoumik, T. Dobson and J. Morris, On the automorphism groups of almost all circulant graphs and digraphs, Ars Math. Contemp. 7 (2014), 499–518, doi:10.26493/1855-3974.315. 868, https://doi.org/10.26493/1855-3974.315.868. [3] E. Dobson and J. Morris, Automorphism groups of wreath product digraphs, Electron. J. Comb. 16 (2009), research paper r17, 30, doi:10.37236/106, https://doi.org/10.37236/ 106. [4] T. Dobson, A. Malnič and D. Marušič, Symmetry in Graphs, Cambridge Studies in Advanced Mathematics, Cambridge University Press, Cambridge, 2022, doi:10.1017/9781108553995, https://doi.org/10.1017/9781108553995. J. Bagherian: Generalized X-join of graphs and their automorphisms 315 [5] S. A. Evdokimov and I. N. Ponomarenko, Schurity of S-rings over a cyclic group and generalized wreath product of permutation groups, St. Petersbg. Math. J. 24 (2013), 431–460, doi:10.1090/S1061-0022-2013-01246-5, https://doi.org/10. 1090/S1061-0022-2013-01246-5. [6] R. Hammack, W. Imrich and S. Klavžar, Handbook of Product of Graphs, Discrete Mathemat- ics and its Applications, CRC Press, Boca Raton, 2nd edition, 2011. [7] F. Harary, On the group of the composition of two graphs, Duke Math. J. 26 (1959), 29–36, doi:10.1215/S0012-7094-59-02603-1, https://doi.org/10.1215/ S0012-7094-59-02603-1. [8] R. L. Hemminger, The group of an X-join of graphs, J. Comb. Theory 5 (1968), 408–418, doi: 10.1016/S0021-9800(68)80017-1, https://doi.org/10.1016/S0021-9800(68) 80017-1. [9] R. Mathon, A note on the graph isomorphism counting problem, Inf. Process. Lett. 8 (1979), 131–132, doi:10.1016/0020-0190(79)90004-8, https://doi.org/10.1016/ 0020-0190(79)90004-8. [10] M. Muzychuk, A wedge product of association schemes, Eur. J. Comb. 30 (2009), 705–715, doi:10.1016/j.ejc.2008.07.008, https://doi.org/10.1016/j.ejc.2008.07.008. [11] G. Sabidussi, The lexicographic product of graphs, Duke Math J. 26 (1961), 573–578, doi:10.1215/S0012-7094-61-02857-5, https://doi.org/10.1215/ S0012-7094-61-02857-5. [12] B. Weisfeiler, On Construction and Identifcation of Graphs, volume 558 of Lecture Notes in Mathematics, Springer-Verlag, Berlin-Heidelberg-New York, 1976. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.07 / 317–325 https://doi.org/10.26493/1855-3974.2126.5b3 (Also available at http://amc-journal.eu) The automorphism group of the zero-divisor digraph of matrices over an antiring David Dolžan * Faculty of Mathematics and Physics, University of Ljubljana, Jadranska 21, 1000 Ljubljana, Slovenia and IMFM, Jadranska 19, 1000 Ljubljana, Slovenia Gabriel Verret Department of Mathematics, University of Auckland, Private Bag 92019, Auckland 1142, New Zealand Received 26 September 2019, accepted 29 May 2023, published online 3 October 2023 Abstract We determine the automorphism group of the zero-divisor digraph of the semiring of matrices over an antinegative commutative semiring with a finite number of zero-divisors. Keywords: Automorphism group of a graph, zero-divisor graph, semiring. Math. Subj. Class. (2020): 05C60, 16Y60, 05C25 1 Introduction In recent years, the zero-divisor graphs of various algebraic structures have received a lot of attention, since they are a useful tool for revealing the algebraic properties through their graph-theoretical properties. In 1988, Beck [3] first introduced the concept of the zero- divisor graph of a commutative ring. In 1999, Anderson and Livingston [1] made a slightly different definition of the zero-divisor graph in order to be able to investigate the zero- divisor structure of commutative rings. In 2002, Redmond [15] extended this definition to also include non-commutative rings. Different authors then further extended this concept to semigroups [6], nearrings [4] and semirings [8]. Automorphisms of graphs play an important role both in graph theory and in algebra, and finding the automorphism group of certain graphs is often very difficult. Recently, a *Corresponding author. The author acknowledges the financial support from the Slovenian Research Agency (research core funding no. P1-0222). E-mail addresses: david.dolzan@fmf.uni-lj.si (David Dolžan), g.verret@auckland.ac.nz (Gabriel Verret) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 318 Ars Math. Contemp. 24 (2024) #P2.07 / 317–325 lot of effort has been made to determine the automorphism group of various zero-divisor graphs. In [1], Anderson and Livingston proved that Aut(Γ(Zn)) is a direct product of symmetric groups for n ≥ 4 a non-prime integer. In the non-commutative case, the case of matrix rings and semirings is especially interesting. Thus, it was shown in [10] that, when p is a prime, Aut(Γ(M2(Zp))) is isomorphic to Sym(p+ 1), the symmetric group of degree p+1. More generally, it was proved in [13], that Aut(Γ(M2(Fq))) ∼= Sym(q+1). In [18], the authors determined the automorphism group of the zero-divisor graph of all rank one upper triangular matrices over a finite field, and in [16] they determined the automorphism group of the zero-divisor graph of the matrix ring of all upper triangular matrices over a finite field. Recently, the automorphism group of the zero-divisor graph of the complete matrix ring of matrices over a finite field has been found independently in [17] and [20]. In this paper, we study the zero-divisor graph of matrices over commutative semirings. The theory of semirings has many applications in optimization theory, automatic control, models of discrete event networks and graph theory (see e.g. [2, 5, 12, 19]) and the zero- divisor graphs of semirings were recently studied in [9, 7, 14]. For an extensive theory of semirings, we refer the reader to [11]. There are many natural examples of commutative semirings, for example, the set of nonnegative integers (or reals) with the usual operations of addition and multiplication. Other examples include distributive lattices, tropical semir- ings, dioı̈ds, fuzzy algebras, inclines and bottleneck algebras. The theory of matrices over semirings differs quite substantially from the one over rings, so the methods we use are necessarily distinct from those used in the ring setting. The main result of this paper is the determination of the automorphism group of the zero- divisor digraph of a semiring of matrices over an antinegative commutative semiring with a finite number of zero-divisors (see Theorem 3.12). 2 Definitions and preliminaries 2.1 Digraphs A digraph Γ consists of a set V(Γ) of vertices, together with a binary relation → on V(Γ). An automorphism σ of Γ is a permutation of V(Γ) such that u → v ⇐⇒ σ(u) → σ(v). The automorphisms of Γ form its automorphism group Aut(Γ). Let Γ be a digraph and let v ∈ V(Γ). We write N−(v) = {u ∈ V(Γ): u → v} and N+(v) = {u ∈ V(Γ): v → u}. If, for u, v ∈ V(Γ), we have N−(u) = N−(v) and N+(u) = N+(v), then we say u and v are twin vertices. The relation ∼ on V(Γ), defined by u ∼ v if and only if u and v are twin vertices, is clearly an equivalence relation preserved by Aut(Γ). For v ∈ V(Γ), we shall denote by v the ∼-equivalence class of v. Let Γ be the graph with these equivalence classes as vertices and u →Γ v if and only if u →Γ v. For σ ∈ Aut(Γ) we denote by σ the induced automorphism of Γ. An automorphism σ ∈ Aut(Γ) is called regular if σ is the identity map. 2.2 Semirings A semiring is a set S equipped with binary operations + and · such that (S,+) is a commu- tative monoid with identity element 0, and (S, ·) is a semigroup. Moreover, the operations + and · are connected by distributivity and 0 annihilates S. A semiring S is commutative if ab = ba for all a, b ∈ S, and antinegative if, for all a, b ∈ S, a + b = 0 implies that a = 0 or b = 0. Antinegative semirings are also called D. Dolžan et al.: The automorphism group of the zero-divisor digraph of matrices . . . 319 zerosum-free semirings or antirings. The smallest nontrivial example of an antiring is the Boolean antiring B = {0, 1} with addition and multiplication defined so that 1 + 1 = 1 · 1 = 1. Let S be a semiring. For x ∈ S, we define the left and right annihilators in S by AnnL(x) = {y ∈ S : yx = 0} and AnnR(x) = {y ∈ S : xy = 0}. If S is commutative, we simply write Ann(x) for AnnL(x) = AnnR(x). We denote by Z(S) the set of zero- divisors of S, that is Z(S) = {x ∈ S : ∃y ∈ S \ {0} such that xy = 0 or yx = 0}. The zero-divisor digraph Γ(S) of S is the digraph with vertex-set S and u → v if and only if uv = 0. It is easy to see that if n ≥ 1 and S is a semiring, then the set Mn(S) of n×n matrices forms a semiring with respect to matrix addition and multiplication. If S is antinegative, then so is Mn(S). If S has an identity 1, let Eij ∈ Mn(S) with entry 1 in position (i, j), and 0 elsewhere. For s ∈ S, define sEij ∈ Mn(S) as the matrix with entry s in position (i, j), and 0 elsewhere. 3 The automorphisms of the zero-divisor digraph The following fact will be used repeatedly. Lemma 3.1. Let S be a semiring. If A,B ∈ S and σ ∈ Aut(Γ(S)), then σ(AnnL(A)) = AnnL(σ(A)) and σ(AnnR(A)) = AnnR(σ(A)). Proof. We have X ∈ σ(AnnL(A)) ⇐⇒ σ−1(X) ∈ AnnL(A) ⇐⇒ σ−1(X)A = 0 ⇐⇒ Xσ(A) = 0 ⇐⇒ X ∈ AnnL(σ(A)). The proof of the second part is analogous. Lemma 3.2. Let S be an antiring and let Γ = Γ(S). If A,B ∈ S and σ ∈ Aut(Γ), then σ(A + B) and σ(A) + σ(B) are twin vertices and, in particular, σ(A+B) = σ(A) + σ(B). Proof. Using antinegativity, we have X ∈ AnnL(σ(A+B)) ⇐⇒ Xσ(A+B) = 0 ⇐⇒ σ−1(X)(A+B) = 0 ⇐⇒ σ−1(X)A = σ−1(X)B = 0 ⇐⇒ Xσ(A) = Xσ(B) = 0 ⇐⇒ X(σ(A) + σ(B)) = 0 ⇐⇒ X ∈ AnnL(σ(A) + σ(B)). We have proved that AnnL(σ(A+B)) = AnnL(σ(A)+σ(B)). An analogous proof yields AnnR(σ(A+B)) = AnnR(σ(A)+σ(B)). This implies that σ(A+B) and σ(A)+σ(B) are twin vertices. 320 Ars Math. Contemp. 24 (2024) #P2.07 / 317–325 Definition 3.3. Let S be a commutative semiring, let n ∈ N and let A ∈ Mn(S) with (i, j) entry aij . For every i, j ∈ {1, . . . , n}, we define Ci(A) = ⋂n k=1 Ann(aki) and Rj(A) = ⋂n k=1 Ann(ajk). Let AR(A) := (C1(A), . . . , Cn(A)) ∈ P(S)n and AL(A) := (R1(A), . . . , Rn(A)) ∈ P(S)n, where P(S) denotes the power set of S. The next theorem characterizes the twin vertices of Γ(Mn(S)). Theorem 3.4. Let S be a commutative antiring, let n ∈ N and let A,B ∈ Mn(S). Then A and B are twin vertices of Γ(Mn(S)) if and only if AL(A) = AL(B) and AR(A) = AR(B). Proof. Let aij and bij be the (i, j) entry of A and B, respectively. Suppose first that A and B are twin vertices of Γ(Mn(S)) and assume that AR(A) ̸= AR(B). This implies that, for some i ∈ {1, . . . , n}, we have Ci(A) ̸= Ci(B). Swapping the role of A and B if necessary, there exists s ∈ S such that s ∈ Ci(A) and s /∈ Ci(B). Therefore, there exists k ∈ {1, . . . , n} such that s /∈ Ann(bki). Now, let C = sEik ∈ Mn(S) and observe that AC = 0 but BC ̸= 0, so N+(A) ̸= N+(B), which is a contradiction with the fact that A and B are twin vertices. We have thus proved that AR(A) = AR(B). A similar argument yields that AL(A) = AL(B). Conversely, assume now that AL(A) = AL(B) and AR(A) = AR(B). Suppose there exists X ∈ Mn(S) such that AX = 0. Therefore, for all i, j ∈ {1, . . . , n} we have∑n k=1 aikxkj = 0. Since S is an antiring, this further implies that aikxkj = 0 for all i, j, k ∈ {1, . . . , n}. So, xkj ∈ Ann(aik) and therefore xkj ∈ Ck(A) = Ck(B) for all k ∈ {1, . . . , n}. Thus, for all i, j, k ∈ {1, . . . , n}, we have xkj ∈ Ann(bik). This yields bikxkj = 0 for all i, j, k ∈ {1, . . . , n}, so BX = 0. Thus, we have proved that N+(A) ⊆ N+(B). By swapping the roles of A in B we also get N+(B) ⊆ N+(A), so N+(A) = N+(B). A similar argument yields that N−(A) = N−(B), thus A and B are twin vertices. Definition 3.5. Let S be a commutative semiring and let α ∈ S \ Z(S). We say that α = e1 + e2 + · · · + es such that ei ̸= 0 for all i and eiej = 0 for all i ̸= j is a decomposition of α of length s. The length ℓ(α) of α is the supremum of the length of a decomposition of α (note that ℓ(α) can be infinite). We say that α is of maximal length if ℓ(α) ≥ ℓ(β) for all β ∈ S \ Z(S). A semiring S is decomposable if S \ Z(S) contains an element of length at least 2, otherwise it is indecomposable. Lemma 3.6. Let S be a commutative antiring and let α ∈ S \ Z(S) be of finite maximal length s with decomposition α = e1 + e2 + · · · + es. Then, for every i ∈ {1, . . . , s}, the subsemiring eiS is indecomposable. Proof. Suppose that eiS is decomposable for some i ∈ {1, . . . , s}, say i = 1 without loss of generality. By definition, there exists e1w ∈ e1S \ Z(e1S) such that e1w = f1 + f2, where f1, f2 ∈ e1S \ {0} and f1f2 = 0. For all j ̸= 1, we have eje1w = 0 and thus ejf1 = ejf2 = 0 by antinegativity. Let β = e1w + e2 + · · ·+ es. Suppose that βx = 0 for some x ∈ S. By antinegativity, we have (e1w)(e1x) = 0 and e2x = · · · = esx = 0. Since e1w is not a zero-divisor in e1S this implies that e1x = 0 and therefore also αx = 0. However, α is not a zero-divisor, so we can conclude that x = 0. This shows that β = f1 + f2 + e2 + · · · + es is not a zero-divisor in S, which is a contradiction with the maximal length of α. D. Dolžan et al.: The automorphism group of the zero-divisor digraph of matrices . . . 321 We shall investigate commutative antirings with identity where 1 is an element of finite maximal length. The next lemma shows that in this case, we can study the automorphisms of the zero-divisor digraph of the matrix ring componentwise. Lemma 3.7. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1 + e2 + · · ·+ es. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))). Then there exists ω ∈ Sym(s) such that, for every r ∈ {1, . . . , s}, we have σ(erMn(S)) = eω(r)Mn(S). Proof. Let r ∈ {1, . . . , s}, let i, j ∈ {1, . . . , n} and let B = σ(erEij). So, σ(erEij) =∑s k=1 ekB. By Lemma 3.2, we have erEij = ∑s k=1 σ −1(ekB). Since S is antinegative, for every k ∈ {1, . . . , s}, there exists fk ∈ S such that σ−1(ekB) = fkEij . Since fk = fk(e1 + e2 + · · · + es) = fker ∈ erS, we have ∑s k=1 fk = erz for z = ∑s k=1 fk. Observe that erEij = zEij yields z ∈ S \ Z(S). Let k, k′ ∈ {1, . . . , s}. We have AnnL(fkEij),AnnL(fk′Eij) ⊆ AnnL(fkfk′Eij) hence AnnL(ekB),AnnL(ek′B) ⊆ AnnL(σ(fkfk′Eij)). If k ̸= k′, then Mn(S) ⊆ AnnL(ekB) + AnnL(ek′B), which is possible only if fkfk′ = 0. We have shown that fkfk′ = 0 for every k, k′ ∈ {1, . . . , s} with k ̸= k′. It follows that z = (e1 + e2 + · · ·+ es)z = ∑ i ̸=r eiz + s∑ k=1 fk is a decomposition of z. Since z /∈ Z(S), eiz ̸= 0 and, since ℓ(z) ≤ ℓ(1) = s, it follows that all but exactly one of the fk’s are 0. This implies that all but one of the ekB’s are 0 and there exists k ∈ {1, . . . , s} such that σ(erEij) = ekB. This shows the existence of a permutation ω ∈ Sym(s) such that σ(erEij) = eω(r)B. Let t ∈ {1, . . . , n}. We have e2r = er ̸= 0, so erEijerEjt ̸= 0 and (eω(r)B)σ(erEjt) ̸= 0. This implies σ(erEjt) ∈ eω(r)Mn(S). As this holds for all j, t ∈ {1, . . . , n} and for any A ∈ Mn(S), we have A = (e1+e2+ · · ·+es)A, we have σ(erMn(S)) ⊆ eω(r)Mn(S). By the same token, we can conclude that a twin vertex to a vertex from erMn(S) is itself in erMn(S), therefore also σ(erMn(S)) ⊆ eω(r)Mn(S). Since σ is a bijection, σ(erMn(S)) = eω(r)Mn(S). We next focus on the automorphisms restricted to the matrices over indecomposable subsemirings. Proposition 3.8. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1 + e2 + · · · + es. Let u, v ∈ {1, . . . , s}, S1 = euS and S2 = evS. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))) such that σ(Mn(S1)) = Mn(S2). If i, j ∈ {1, . . . , n}, then there exist y ∈ S2 \ Z(S2) and k, ℓ ∈ {1, . . . , n} such that σ(euEij) = yEkℓ. Proof. Write σ(euEij) = ∑ k,ℓ βkℓEkℓ. Let Akℓ = σ −1(βkℓEkℓ). By Lemma 3.2, σ(euEij) and σ (∑ k,ℓ Akℓ ) are twin vertices, therefore euEij and ∑ k,ℓ Akℓ are twin ver- tices as well. Now, twin vertices of euEij must be of the form zEij , so ∑ k,ℓ Akℓ = zEij for some z ∈ S. Since z = z(e1 + e2 + · · · + es), we conclude that z ∈ S1. Note also 322 Ars Math. Contemp. 24 (2024) #P2.07 / 317–325 that z is not a zero-divisor in S1, since euEij and zEij are twin vertices. Since S is an- tinegative, we can conclude that, for all k, ℓ ∈ {1, . . . , n}, there exist αkℓ ∈ S1 such that Akℓ = αkℓEij and ∑ k,ℓ αkℓ = z. Let k, k′, ℓ, ℓ′ ∈ {1, . . . , n} with (k, ℓ) ̸= (k′, ℓ′). Now, we either have k ̸= k′ or ℓ ̸= ℓ′. Suppose first that k ̸= k′. Since S is commutative, we have AnnL(Akℓ) = AnnL(αkℓEij) ⊆ AnnL(αkℓαk′ℓ′Eij). By Lemma 3.1, this implies AnnL(βkℓEkℓ) ⊆ AnnL(σ(αkℓαk′ℓ′Eij)). Similarly, we have AnnL(βk′ℓ′Ek′ℓ′) ⊆ AnnL(σ(αkℓαk′ℓ′Eij)). Since k ̸= k′, Mn(S) = AnnL(βkℓEkℓ) + AnnL(βk′ℓ′Ek′ℓ′) ⊆ AnnL(σ(αkℓαk′ℓ′Eij)), which implies σ(αkℓαk′ℓ′Eij) = 0 and thus αkℓαk′ℓ′ = 0. If ℓ ̸= ℓ′, we arrive at the same conclusion by using right annihilators, namely that distinct αkℓ’s annihilate each other. Since S1 is indecomposable by Lemma 3.6, the sum ∑ k,ℓ αkℓ = z has at most one non- zero summand. It follows that there is at most one non-zero Akℓ and at most one non-zero βkℓEkℓ. This concludes the proof of the first part, with y = βkℓ. It remains to show that y /∈ Z(S2). Suppose, on the contrary, that y ∈ Z(S2). By the first part of the result, there exist y′ ∈ S1 and i′, j′ ∈ {1, . . . , n} such that σ−1(evEkℓ) = y′Ei′j′ . Since y ∈ Z(S2), we have AnnL(evEkℓ) ⊊ AnnL(yEkℓ) and AnnR(evEkℓ) ⊊ AnnR(yEkℓ). By Lemma 3.1, it follows that AnnL(y′Ei′j′) ⊊ AnnL(euEij) and of course also AnnR(y′Ei′j′) ⊊ AnnR(euEij). This is only possible if i = i′ and j = j′ which implies AnnL(y′Eij) ⊊ AnnL(euEij), a contradiction. Lemma 3.9. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1 + e2 + · · · + es. Let u, v ∈ {1, . . . , s}, S1 = euS and S2 = evS. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))) such that σ(Mn(S1)) = Mn(S2). If x ∈ Z(S1) and i, j ∈ {1, . . . , n}, then there exist z ∈ Z(S2) and k, ℓ ∈ {1, . . . , n} such that σ(xEij) = zEkℓ. Proof. By Proposition 3.8, we know that σ(euEij) = yEkℓ for some y /∈ Z(S2) and k, ℓ ∈ {1, . . . , n}. Since x ∈ Z(S1), we have AnnL(euEij) ⊊ AnnL(xEij) and AnnR(euEij) ⊊ AnnR(xEij). By Lemma 3.1, it follows that AnnL(yEkℓ) ⊊ AnnL(σ(xEij)) and also AnnR(yEkℓ) ⊊ AnnR(σ(xEij)). This implies that all entries of σ(xEij) are zeros except entry (k, ℓ), so σ(xEij) = zEkℓ for some z ∈ S2. Because AnnL(yEkℓ) ̸= AnnL(σ(xEij)) = AnnL(zEkℓ) and AnnR(yEkℓ) ̸= AnnR(σ(xEij)) = AnnR(zEkℓ), we must have z ∈ Z(S2). Lemma 3.10. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1 + e2 + · · ·+ es. Let u, v ∈ {1, . . . , s}, S1 = euS and S2 = evS. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))) such that σ(Mn(S1)) = Mn(S2). Then there exists π ∈ Sym(n) such that σ(euEij) = evEπ(i)π(j) for all i, j ∈ {1, . . . , n}. Proof. Let i, j, j′ ∈ {1, . . . , n} with j ̸= j′. By Proposition 3.8, there exist k, k′, ℓ, ℓ′ ∈ {1, . . . , n} such that σ(euEij) = evEkℓ and σ(euEij′) = evEk′ℓ′ . For all r, s ∈ {1, . . . , n} with s ̸= i, we have euErs(euEij + euEij′) = 0. By Lemma 3.2, this implies that σ(euErs)(euEkℓ + euEk′ℓ′) = 0 and thus (∑ r,s ̸=i σ(euErs) ) (euEkℓ + euEk′ℓ′) = 0. By Proposition 3.8, σ(euErs) = evEr′s′ for some r′, s′ ∈ {1, . . . , n}. Since σ is a per- mutation, ∑ r,s ̸=i σ(euErs) is a matrix with exactly n entries equal to 0. It follows that k = k′. By the paragraph above, there exists π ∈ Sym(n) such that σ(euEab) = evEπ(a)c, for some c. A similar argument yields that there exists a permutation such that σ(euEab) = D. Dolžan et al.: The automorphism group of the zero-divisor digraph of matrices . . . 323 evEcπ′(b), for some c. However, for every j, k ∈ {1, . . . , n} with j ̸= k, we have EjjEkk = 0 and thus Eπ(j)π′(j)Eπ(k)π′(k) = 0. This implies that π(k) ̸= π′(j) for every k ̸= j, so π(j) = π′(j). Therefore π′ = π. For π ∈ Sym(n) and A ∈ Mn(S), let θπ(A) be the matrix obtained from A by applying the permutation π to its rows and columns. Note that θπ induces a permutation of Mn(S). Corollary 3.11. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1+e2+ · · ·+es. Let u, v ∈ {1, 2, . . . , s}, S1 = euS and S2 = evS. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))) such that σ(Mn(S1)) = Mn(S2). Then there exist π ∈ Sym(n) and τ an isomorphism from Γ(S1) to Γ(S2) such that, if we extend τ entry-wise to a mapping Mn(S1) → Mn(S2) and restrict σ to Mn(S1), then σ = θπ ◦ τ . Proof. By Lemma 3.10, there exists π ∈ Sym(n) such that σ(euEij) = θπ(evEij) for all i, j ∈ {1, . . . , n}. Let ρ = θ−1π ◦ σ and note that ρ ∈ Aut(Γ(Mn(S))) and we have ρ(euEij) = evEij for all i, j ∈ {1, . . . , n}. Let x ∈ Z(S1) and i, j, j′ ∈ {1, . . . , n}. Clearly, ρ(Mn(S1)) = Mn(S2) so, by Lemma 3.9, there exist z, z′ ∈ Z(S2) such that ρ(xEij) = zEij and ρ(xEij′) = z′Eij′ . Let X = {s ∈ S2; sz = 0}. We show that X ⊆ Ann(z′). Let a ∈ S2 such that az = 0. Note that (aEii)(zEij) = 0. Since a ∈ Z(S2), Lemma 3.9 implies that there exists b ∈ Z(S1) such that aEii = ρ(bEii), hence ρ(bEii)ρ(xEij) = 0 and there- fore also (bEii)(xEij) = 0 which implies bx = 0. It follows that (bEii)(xEij′) = 0, ρ(bEii)ρ(xEij′) = 0 and (aEii)(z′Eij′) = 0 which yields az′ = 0. We have shown that X ⊆ Ann(z′). A symmetrical argument yields Ann(z′) ⊆ X hence X = Ann(z′) which implies that ρ(xEij′) = z′Eij′ = zEij′ (where for A ∈ Mn(S2), by a slight abuse of notation, A now refers to the image in Γ(Mn(S2)) and not in Γ(Mn(S))). A similar argument shows that ρ(xEi′j) = zEi′j for all i′ ∈ {1, . . . , n}. This implies that ρ(xEkℓ) = zEkℓ for all k, ℓ ∈ {1, . . . , n}. Let τ denote the mapping S1 → S2 that satisfies ρ(xE11) = τ(x)E11. Since ρ is a bijection from Mn(S1) to Mn(S2), τ is a bijection from S1 to S2. If x, y ∈ S1, then xy = 0 if and only if (xE11)(yE11) = 0 if and only if τ(x)τ(y) = 0, therefore τ is an isomorphism from Γ(S1) to Γ(S2). Now, extend τ to an entry-wise mapping Mn(S1) → Mn(S2). Let A = [aij ], B = [bij ] ∈ Mn(S1). Note that AB = 0 if and only if ∑n k=1 aikbkj = 0 for every 1 ≤ i, j ≤ n if and only if aikbkj = 0 for every 1 ≤ i, j, k ≤ n, so τ induces an isomorphism from Γ(Mn(S1)) to Γ(Mn(S2)). Observe that, restricted to V(Γ(Mn(S1))), we have ρ = τ . As σ = θπ ◦ ρ, this concludes the proof. We can now join these findings into the following theorem. Theorem 3.12. Let S be a commutative antiring and suppose 1 ∈ S is of finite maximal length s with decomposition 1 = e1 + e2 + · · ·+ es. Let n ∈ N and σ ∈ Aut(Γ(Mn(S))). Then there exist ω ∈ Sym(s) and, for every i ∈ {1, . . . , s}, there exist πi ∈ Sym(n) and an isomorphism τi : Γ(eiS) → Γ(eω(i)S) such that, if we extend τi entry-wise to a mapping Mn(eiS) → Mn(eω(i)S), then σ(A) = ( s∑ i=1 (θπi ◦ τi)(eiA) ) for all A ∈ Mn(S). 324 Ars Math. Contemp. 24 (2024) #P2.07 / 317–325 Conversely, if ω ∈ Sym(s) has the property that, for every i ∈ {1, . . . , s}, we have Γ(eiS) ∼= Γ(eω(i)S), τi is an isomorphism from Γ(eiS) to Γ(eω(i)S) and πi ∈ Sym(n), then σ defined with σ(A) = ∑s i=1 (θπi ◦ τi)(eiA) is an automorphism of Γ(Mn(S)). Proof. By Lemma 3.7, there exists ω ∈ Sym(s) such that, for every i ∈ {1, . . . , s}, we have σ(eiMn(S)) = eω(i)Mn(S). By Corollary 3.11, there exist πi ∈ Sym(n) and τi an isomorphism from Γ(eiS) to Γ(eω(i)S) such that, if we extend τi entry-wise to a mapping Mn(eiS) → Mn(eω(i)S) and restrict σ to Mn(eiS), then σ = θπi ◦ τi. Now, let A ∈ Mn(S). We have A = e1A+ e2A+ · · ·+ esA and the result follows by Lemma 3.2. Remark 3.13. Throughout the paper, we restricted ourselves to studying semirings with the property that no non-zero-divisor element can be written as a sum of infinitely many mutually orthogonal zero-divisors. Obviously, any semiring with a finite set of zero- divisors satisfies this condition. ORCID iDs David Dolžan https://orcid.org/0000-0001-6548-3945 Gabriel Verret https://orcid.org/0000-0003-1766-4834 References [1] D. F. Anderson and P. S. Livingston, The zero-divisor graph of a commutative ring, J. Algebra 217 (1999), 434–447, doi:10.1006/jabr.1998.7840, https://doi.org/10.1006/jabr. 1998.7840. [2] F. Baccelli and J. Mairesse, Ergodic theorems for stochastic operators and discrete event net- works, in: Idempotency, Cambridge University Press, Cambridge, pp. 171–208, 1998. [3] I. Beck, Coloring of commutative rings, J. Algebra 116 (1988), 208–226, doi:10.1016/ 0021-8693(88)90202-5, https://doi.org/10.1016/0021-8693(88)90202-5. [4] G. A. Cannon, K. M. Neuerburg and S. P. Redmond, Zero-divisor graphs of nearrings and semigroups, in: H. Kiechle, A. Kreuzer and M. Thomsen (eds.), Nearrings and Nearfields, Springer, Dordrecht, pp. 189–200, 2005. [5] R. Cuninghame-Green, Minimax Algebra, volume 166, Springer Science & Business Media, 2012. [6] F. R. DeMeyer, T. McKenzie and K. Schneider, The zero-divisor graph of a commutative semigroup, Semigroup Forum 65 (2002), 206–214, doi:10.1007/s002330010128, https: //doi.org/10.1007/s002330010128. [7] D. Dolžan and P. Oblak, The zero-divisor graphs of rings and semirings, Int. J. Algebra Com- put. 22 (2012), 1250033, 20, doi:10.1142/s0218196712500336, https://doi.org/10. 1142/s0218196712500336. [8] S. Ebrahimi Atani, The zero-divisor graph with respect to ideals of a commutative semiring., Glas. Mat., III. Ser. 43 (2008), 309–320, doi:10.3336/gm.43.2.06, https://doi.org/10. 3336/gm.43.2.06. [9] S. Ebrahimi Atani, An ideal based zero-divisor graph of a commutative semiring., Glas. Mat., III. Ser. 44 (2009), 141–153, doi:10.3336/gm.44.1.07, https://doi.org/10.3336/gm. 44.1.07. D. Dolžan et al.: The automorphism group of the zero-divisor digraph of matrices . . . 325 [10] J. Han, The zero-divisor graph under group actions in a noncommutative ring, J. Korean Math. Soc. 45 (2008), 1647–1659, doi:10.4134/jkms.2008.45.6.1647, https://doi.org/10. 4134/jkms.2008.45.6.1647. [11] U. Hebisch and H. J. Weinert, Semirings: Algebraic theory and Applications in Computer Science, volume 5 of Ser. Algebra, World Scientific Publishing, Singapore, 1998. [12] P. Li, A heuristic method to compute the approximate postinverses of a fuzzy matrix, IEEE Trans. Fuzzy Syst. 22 (2014), 1347–1351, doi:10.1109/tfuzz.2013.2282231, https://doi. org/10.1109/tfuzz.2013.2282231. [13] S. Park and J. Han, The group of graph automorphisms over a matrix ring, J. Korean Math. Soc. 48 (2011), 301–309, doi:10.4134/jkms.2011.48.2.301, https://doi.org/10. 4134/jkms.2011.48.2.301. [14] A. Patil, B. N. Waphare and V. Joshi, Perfect zero-divisor graphs, Discrete Math. 340 (2017), 740–745, doi:10.1016/j.disc.2016.11.027, https://doi.org/10.1016/j. disc.2016.11.027. [15] S. P. Redmond, The zero-divisor graph of a non-commutative ring, Int. J. Commut. Rings 1 (2002), 203–211. [16] L. Wang, A note on automorphisms of the zero-divisor graph of upper triangular matrices., Linear Algebra Appl. 465 (2015), 214–220, doi:10.1016/j.laa.2014.09.035, https://doi. org/10.1016/j.laa.2014.09.035. [17] L. Wang, Automorphisms of the zero-divisor graph of the ring of all n × n matrices over a finite field, Discrete Math. 339 (2016), 2036–2041, doi:10.1016/j.disc.2016.02.021, https: //doi.org/10.1016/j.disc.2016.02.021. [18] D. Wong, X. Ma and J. Zhou, The group of automorphisms of a zero-divisor graph based on rank one upper triangular matrices, Linear Algebra Appl. 460 (2014), 242–258, doi:10.1016/j. laa.2014.07.041, https://doi.org/10.1016/j.laa.2014.07.041. [19] S. Zhao and X.-p. Wang, Invertible matrices and semilinear spaces over commutative semirings, Inf. Sci. 180 (2010), 5115–5124, doi:10.1016/j.ins.2010.08.033, https://doi.org/10. 1016/j.ins.2010.08.033. [20] J. Zhou, D. Wong and X. Ma, Automorphisms of the zero-divisor graph of the full matrix ring, Linear Multilinear Algebra 65 (2017), 991–1002, doi:10.1080/03081087.2016.1219302, https://doi.org/10.1080/03081087.2016.1219302. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.08 / 327–346 https://doi.org/10.26493/1855-3974.2947.cd6 (Also available at http://amc-journal.eu) Quotients of skew morphisms of cyclic groups Martin Bachratý * Faculty of Civil Engineering, Slovak University of Technology, Bratislava, Slovakia Received 25 August 2022, accepted 19 April 2023, published online 4 October 2023 Abstract A skew morphism of a finite groupB is a permutation φ ofB that preserves the identity element of B and has the property that for every a ∈ B there exists a positive integer ia such that φ(ab) = φ(a)φia(b) for all b ∈ B. The problem of classifying skew morphisms for all finite cyclic groups is notoriously hard, with no such classification available up to date. Each skew morphism φ of Zn is closely related to a specific skew morphism of Z|⟨φ⟩|, called the quotient of φ. In this paper, we use this relationship and other observations to prove new theorems about skew morphisms of finite cyclic groups. In particular, we classify skew morphisms for all cyclic groups of order 2em with e ∈ {0, 1, 2, 3, 4} and m odd and square-free. We also develop an algorithm for finding skew morphisms of cyclic groups, and implement this algorithm in MAGMA to obtain a census of all skew morphisms for cyclic groups of order up to 161. During the preparation of this paper we noticed a few flaws in Section 5 of the paper Cyclic complements and skew morphisms of groups from 2016. We propose and prove weaker versions of the problematic original assertions (namely Lemma 5.3(b), Theorem 5.6 and Corollary 5.7), and show that our modifications can be used to fix all consequent proofs (in the aforementioned paper) that use at least one of those problematic assertions. Keywords: Skew morphism, cyclic group, coset-preserving, quotient, square-free. Math. Subj. Class. (2020): 20B25, 05C25, 05E18 1 Introduction A skew morphism of a finite group B is a permutation φ of B that preserves the identity element of B and has the property that for every a ∈ B there exists a positive integer ia such that φ(ab) = φ(a)φia(b) for all b ∈ B. The order of a skew morphism, denoted by *The author acknowledges the use of the MAGMA system [7] to find examples of skew morphisms relevant to this paper. The author also acknowledges support from the APVV Research Grants 17-0428 and 19-0308, and the VEGA Research Grants 1/0206/20 and 1/0567/22. E-mail address: martin.bachraty@stuba.sk (Martin Bachratý) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 328 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 ord(φ), is defined as the order of the cyclic group ⟨φ⟩. Note that for each a ∈ B there is a unique choice for ia such that ia ∈ {1, 2, . . . , ord(φ) − 1} (unless φ is the identity permutation). The function π that maps each element a ∈ B to this integer ia is called the power function of φ, and it satisfies φ(ab) = φ(a)φπ(a)(b) for all a, b ∈ B. In the case when φ is the identity permutation of B, we define π(a) = 1 for all a ∈ B. Skew morphisms were first introduced by Jajcay and Širáň in [18], with primary inter- est in their connection to the regular Cayley maps. Skew morphisms are also intriguing from a purely group-theoretical point of view, mainly due to their close relationship with group automorphisms, with which they share a number of important features. The prob- lem of classifying all skew morphisms for given families of finite groups has gained much attention in the last two decades; see [8, 10, 23, 24] for example. Recently, in [4], skew morphisms were classified for all finite simple groups, and we understand that a classifica- tion for dihedral groups is imminent; see [19]. On the other hand, the problem of finding all skew morphisms for finite cyclic groups remains open, despite recent positive progress, which we discuss next. Automorphisms of a finite group B are special cases of skew morphisms (with π(a) = 1 for all a ∈ B), and as such can be viewed as an important family of skew morphisms. There are also other intriguing families of skew morphisms, for example, coset-preserving (sometimes also called smooth) skew morphisms, which are defined as skew morphisms satisfying π(a) = π(φ(a)) for all a ∈ B. Coset-preserving skew morphisms have been fully classified for all finite cyclic groups in [6]. Another interesting family of skew mor- phisms that is fully understood for finite cyclic groups consists of all skew morphisms φ such that φ2 is an automorphism of the same group; see [16]. While there is no classification of skew morphisms of finite cyclic groups available to date, skew morphisms have been fully classified for some specific (infinite) families of finite cyclic groups. Most notably, this was done for cases where the order of a cyclic group is a prime [18] (in this case, all skew morphisms are automorphisms), a product of two distinct primes [20], and any power of an odd prime [21]. Some partial progress for cyclic 2-groups can be found in [14]. Another approach for studying skew morphisms of finite cyclic groups is to find a connection between skew morphisms of a given cyclic group B and skew morphisms of cyclic groups of smaller orders. Presumably the strongest finding to date made in this direction is the observation of Kovács and Nedela proved in [20] which states that if gcd(m,n) = gcd(m,ϕ(n)) = gcd(ϕ(m), n) = 1, then the skew morphisms of Zmn are exactly the direct products of skew morphisms of Zm and Zn. There is also a useful connection between general skew morphisms and coset-preserving skew morphisms for fi- nite cyclic groups. Namely, for each skew morphism φ of a finite cyclic group B there exists an exponent e such that φe is a non-trivial coset-preserving skew morphism of B; see [5]. In this paper, we combine a number of known facts about skew morphisms (which we summarise in Sections 2 and 3) with new observations presented in Section 4, to prove a number of theorems about skew morphisms of cyclic groups. Namely, in Section 5 we develop a new method for finding skew morphisms of cyclic groups, and implement it to obtain a census of all skew morphisms of cyclic groups of order up to 161. (Up to the time of writing this paper, and apart from some specific orders, skew morphisms of cyclic groups were known only up to order 60; see [9].) Further, in Section 6 we show that all skew morphisms of Zn are coset-preserving if and only if n = 2em for some e ∈ {0, 1, 2, 3, 4} M. Bachratý: Quotients of skew morphisms of cyclic groups 329 and m odd and square-free. As a consequence, we obtain a complete classification of skew morphisms for all cyclic groups of order expressible in this form, significantly expanding the list of finite cyclic groups for which such classification is available. During the review process, it was communicated to us that Kan Hu, István Kovács and Young Soo Kwon has recently submitted a paper devoted to similar ideas to those we investigated in Section 6 of this paper. 2 Preliminaries In this section, we recall some definitions from group theory and provide some background from the theory of skew morphisms. All groups considered in this paper are assumed to be finite. For the cyclic group of order n we use the additive notation Zn, and so the elements of Zn may be viewed as integers in the interval [0, n− 1]. We also let Sym(G) denote the symmetric group on (the underlying set of) a group G. The core of a subgroup H in a group G is the largest normal subgroup of H contained in G. We say that H is core-free in G if the core of H in G is trivial. A complement for H in G is a subgroup K of G such that G = HK and H ∩K = {1}. The following theorem proved by Lucchini in [22] will be helpful. Theorem 2.1 ([22]). Let C be a cyclic proper subgroup of a group G. If C is core-free in G, then |C| < |G : C|. Next, let φ be a skew morphism of a group B, and identify B with the subgroup of Sym(B) which acts by left multiplication. Then it can be easily checked that B⟨φ⟩ is a subgroup of Sym(B) (see [20] for example). Moreover, B⟨φ⟩ is a complementary factori- sation and ⟨φ⟩ is core-free in B⟨φ⟩ (see [12, Lemma 4.1]). A group G containing B which has a cyclic core-free complement C for B is called a skew product group for a group B, and we say that C is a skew complement (for B in G). The skew product group B⟨φ⟩ (for B) with skew complement ⟨φ⟩ described in this paragraph is said to be induced by φ. Conversely, let G be a skew product group for a group B, and let c be a generator of a skew complement for B in G. Note that every element g ∈ G is uniquely expressible in a form g = ac′ with a ∈ B and c′ ∈ C. Then for every a ∈ B there exists a unique a′ ∈ B and a unique exponent j ∈ {1, 2, . . . , |C| − 1} such that ca = a′cj , and this induces a bijection φ : B → B and a function π : B → N, defined by φ(a) = a′ and π(a) = j. It can be easily checked that φ is a skew morphism of B with power function π. We say that φ is induced by the pair (B, c). Recall that if φ is a skew morphism of B with power function π, then we have φ(ab) = φ(a)φπ(a)(b) for all a, b ∈ B. (Hence, if B = Zn, then φ(a+ b) = φ(a)+φπ(a)(b) for all a, b ∈ Zn.) Also recall that an automorphism of B is a skew morphism with π(a) = 1 for all a ∈ B. In what follows, it will be often convenient to distinguish between general skew morphisms and skew morphisms that are not automorphisms, and so we will refer to the latter as proper skew morphisms. We say that φ is trivial if it is the identity permutation of B. The kernel of φ, denoted by kerφ, is the subset {a ∈ B | π(a) = 1} of B. By definition, φ is an automorphism of B if and only if kerφ = B. In the case of proper skew morphisms kerφ is not equal to B, but it is always a subgroup of B; see [18, Lemma 4]. Since kerφ is a subgroup of B and also φ(ab) = φ(a)φ(b) for all a, b ∈ kerφ, it follows that φ restricts to a group isomorphism from the kernel to its image. In particular, if kerφ is preserved by φ set-wise, then φ restricts to an automorphism of kerφ. In [11] this was shown to always be true for abelian groups. We also have the following. 330 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 Lemma 2.2 ([18]). Let φ be a skew morphism of a group B with power function π. Then two elements a, b ∈ B belong to the same right coset of the subgroup kerφ in B if and only if π(a) = π(b). Theorem 2.3 ([12]). Every skew morphism of a non-trivial group has non-trivial kernel. An immediate consequence of Theorem 2.3 is that every skew morphism of a group of prime order is an automorphism. The following facts about kernels of skew morphisms will be useful, too. Lemma 2.4 ([12]). Let G be a skew product group for a group B with skew complement C, and let c be a generator of C. If φ is the skew morphism induced by (B, c), then kerφ is the largest subgroup H of B for which cHc−1 ⊆ B. In particular, φ is an automorphism of B if and only if B is normal in G. Proposition 2.5 ([12]). Let φ be a skew morphism of a group B, and let N be a subgroup of kerφ that is normal in B and preserved by φ. Then the mapping φ∗N : B/N → B/N given by φ∗N (x) = Nφ(x) is a well-defined skew morphism of B/N . Lemma 2.6 ([12]). Let φ be a skew morphism of a finite abelian group A with power function π. Also suppose that N is any non-trivial subgroup of kerφ preserved by φ, let i be the exponent of N , and let φ∗N be the skew morphism of A/N induced by φ. If a is an element of A such that Na lies in the kernel of φ∗N , then iπ(a) ≡ i (mod ord(φ)) and, in particular, if gcd(i, ord(φ)) = 1, then a ∈ kerφ. Note that if φ is a skew morphism of Zn, then kerφ is normal in Zn and φ restricts to an automorphism of kerφ. Moreover, since Zn is cyclic, so is (its subgroup) kerφ. Noting that every automorphism of a cyclic group preserves all of its subgroups, we deduce that φ preserves all subgroups of kerφ. Hence it follows from Proposition 2.5 that φ∗N is a well defined skew morphism of Zn/N for each subgroup N of kerφ. In the case when N = kerφ, we will write simply φ∗ instead of φ∗N . We proceed with some useful facts about orders of skew morphisms. Proposition 2.7 ([20]). Let φ be a skew morphism of a group B, and let T be an orbit of ⟨φ⟩. If ⟨T ⟩ = B, then ord(φ) = |T |. Theorem 2.8 ([12]). The order of a skew morphism of a non-trivial group B is less than the order of B. Theorem 2.9 ([20]). If φ is a skew morphism of a group Zn, then ord(φ) is a divisor of nϕ(n). Moreover, if gcd(ord(φ), n) = 1, then φ is an automorphism of Zn. The periodicity of a skew morphism φ of a group B, denoted by pφ, is the smallest positive integer such that π(a) = π(φpφ(a)) for all a ∈ B. Similarly, the periodicity of a ∈ B (with respect to φ) is the smallest positive integer pa such that π(a) = π(φpa(a)). Note that if kerφ is preserved by φ (which is always true if B is abelian), then the periodicity of φ can be defined equivalently as the order of φ∗. Recall that φ is coset-preserving if π(a) = π(φ(a)) for all a ∈ B. Equivalently, coset- preserving skew morphisms can be viewed as skew morphism that preserves all right cosets of kerφ inB, or as skew morphisms with the periodicity equal to 1. The following theorem about periodicities underlines the importance of coset-preserving skew morphisms in the study of skew morphisms of abelian (and, in particular, cyclic) groups. M. Bachratý: Quotients of skew morphisms of cyclic groups 331 Theorem 2.10 ([5]). If φ is a skew morphism of an abelian group A, then φpφ is a coset- preserving skew morphism of A. Moreover, if A is cyclic, b is a generator of A, and φ is non-trivial, then pφ = pb and pφ < ord(φ). The following fact will be helpful too. Lemma 2.11 ([12]). Let φ be any skew morphism of a finite group G, and let H be any finite group. Then φ can be extended to a skew morphism θ of G×H , such that θ ↾G= φ and ker θ = kerφ×H . We conclude this section with two well-known facts about skew morphisms that are easy exercises, but we include their proofs for completeness. Lemma 2.12. Let φ be a skew morphism of an abelian group A. If φ(a) = a for some a ∈ A, then a ∈ kerφ. Proof. Let a′ be any element of A. Since a is fixed by φ, we have φ(aa′) = aφπ(a)(a′). On the other hand, we have φ(aa′) = φ(a′a) = φ(a′)φπ(a ′)(a) = φ(a′)a, and hence φ(a′) = φπ(a)(a′) for all a′ ∈ A. It follows that π(a) = 1, and therefore a ∈ kerφ. Lemma 2.13. Let φ be a skew morphism of a group B with power function π, and let a ∈ B. Then: φ(ai) = φ(a)φπ(a)(a)φπ(a 2)(a) . . . φπ(a i−1)(a) for all i ∈ N. Proof. The assertion is trivially true for i = 1. Next, if it holds for some positive integer j, then φ(aj+1) = φ(aja) = φ(aj)φπ(a j)(a) = φ(a)φπ(a)(a) . . . φπ(a j−1)(a)φπ(a j)(a), and so it is also true for j + 1. Hence the proof follows by induction. 3 Skew morphisms of abelian groups Several useful facts about skew morphisms of abelian groups (which we also apply in this paper) were proved in [12]. During the preparation of this paper, however, we noticed that three assertions in [12] do not hold. In this section, we list the incorrect findings and provide a counterexample for each of them. We also propose and prove weaker versions of the original statements. Finally, we discuss all proofs in [12] that use at least one of the flawed statements, and show that in all cases it is sufficient to replace the flawed statements by our modifications. For the rest of the section, we let φ be a skew morphism of an abelian group A with power function π. Also, to distinguish between references to this paper and references to [12], we put an asterisk after each numbered reference in the latter case. The first flawed assertion in [12] is part (b) of Lemma 5.3∗. It states that if N is a non- trivial subgroup of kerφ preserved by φ, and φ is not an automorphism of A, then ord(φ) has a non-trivial divisor in common with the exponent of N . To show that this is not true, note that according to [9] the cyclic group of order 12 admits a proper skew morphism ψ of order 3 with kernel (which is preserved by ψ) of order 6. Since kerψ is cyclic and preserved by ψ, so is its unique subgroup of order 2. But the exponent of this subgroup, which is 2, does not have a non-trivial divisor in common with ord(ψ). We propose the following modification: Lemma 3.1. If φ is a proper skew morphism of an abelian group A, then ord(φ) has a non-trivial divisor in common with the exponent of kerφ. 332 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 Proof. Let K = kerφ, let e be the exponent of K, and let L/K be the kernel of the skew morphism φ∗ of A/K induced by φ. Since φ is proper we know that A/K is a non-trivial group, and by Theorem 2.3 it follows that L/K is non-trivial. In particular, there exists an element a of A such that a /∈ K and a ∈ L. It follows that π(a) ̸≡ 1 (mod ord(φ)), by Lemma 2.6 we have eπ(a) ≡ e (mod ord(φ)), and the rest follows. Part (b) of Lemma 5.3∗ is used in the proofs of six theorems presented in [12]. Proofs of Theorem 5.4∗, Theorem 5.10∗, Theorem 6.2∗ and Theorem 6.4∗ are all easily fixable, since in each case Lemma 5.3(b)∗ is applied for N = K, and hence it is sufficient to replace it with Lemma 3.1. The proof of Theorem 6.1∗ can also be corrected. Here Lemma 5.3(b)∗ is used to show that if φ is proper and A ∼= Zn, then gcd(ord(φ), n) ̸= 1. Since A is cyclic, the exponent of kerφ is equal to the order of kerφ (which necessarily divides n), so this is an easy consequence of Lemma 3.1. Finally, since Theorem 7.3∗ was already proved previously in [20], there is no need to fix its alternative proof presented in [12]; although we believe that it is possible. Another flawed assertion in [12] is Theorem 5.6∗. It states that if L/(kerφ) is the kernel of the skew morphism φ∗ of A/(kerφ) induced by φ, and p is a prime that divides |L| but not |kerφ|, then p < q for every prime divisor q of |kerφ|. Again, we provide a counterexample that shows that this is not true. According to [9] the cyclic group of order 42 admits a proper skew morphism ρ of order 7 with kernel of order 14. Since Z42/(ker ρ) is isomorphic to Z3 and Z3 does not admit any proper skew morphism, it follows that the kernel L/(ker ρ) (of the skew morphism of Z42/(ker ρ) induced by ρ) is equal to Z42/(ker ρ). Therefore, |L| = |Z42| and, in particular, 3 divides |L|. But 3 is greater than 2, and 2 is a prime divisor of |ker ρ|. We propose the following modification: Theorem 3.2. Let φ be a skew morphism of the finite abelian group A, let K = kerφ, and let q be any prime divisor of |K|. Also let N be a subgroup of K consisting of the identity and all elements of order q, and let L/N be the kernel of the skew morphism φ∗N of A/N induced by φ. If p is a prime that divides |L| but not |K|, then p < q. Proof. Suppose that such a prime p exists. Since K is abelian, we know that N is a subgroup of K of exponent q that is invariant under φ. Next, let a be any element of order p in L, let m = ord(φ), and let π be the power function of φ. Since L/N is the kernel of φ∗N , we know by Lemma 2.6 that q(π(a) − 1) ≡ 0 (mod m). If q is relatively prime to m, then π(a) ≡ 1 (mod m) and so a ∈ K, which is impossible since K has no element of order p. Thus q divides m and π(a) − 1 ≡ 0 (mod m/q). In particular, π(a) = 1 + i(m/q) where 1 ≤ i ≤ q − 1, so there are at most q − 1 possibilities for π(a). The same holds for every non-trivial power of a. So now if p > q, then by the pigeon- hole principle two different powers of a will have the same value under π, in which case they lie in the same coset of N . But that cannot happen since K ∩ ⟨a⟩ is trivial. Thus p < q. The only application of the flawed original version of Theorem 5.6∗ is Corollary 5.7∗. This final problematic assertion in [12] states that every prime divisor of |kerφ| is greater than every prime that divides |A| but not |kerφ|. To see that this is not true, take the skew morphism ρ of Z42 with kernel of order 14 discussed earlier. Since 2 divides |ker ρ| and 3 divides |Z42| but not |ker ρ|, the statement is clearly not true. The following, however, still holds. M. Bachratý: Quotients of skew morphisms of cyclic groups 333 Corollary 3.3. Let A be a non-trivial finite abelian group, and let p be the largest prime divisor of |A|. Then the order of the kernel of every skew morphism of A is divisible by p when p is odd, or by 4 when p = 2. Proof. Let φ be any skew morphism of A, let K = kerφ, and suppose to the contrary that p does not divide |K|. Also let q be any prime divisor of |K|, let N be a subgroup of K consisting of the identity and all elements of order q, and let L/N be the kernel of the skew morphism φ∗N of A/N induced by φ. If p divides |L/N |, then by Theorem 3.2 we know that p is smaller than q, so this cannot happen. It follows that p does not divide |L/N |, so we can repeat the same argument for the skew morphism φ∗N of A/N with kernel L/N . (Note that p divides |A/N |.) Since A is finite, this will eventually terminate for some groups A′, N ′ and L′ with |L′/N ′| = 1. Then since the kernel L′/N ′ is trivial, it follows by Theorem 2.3 thatA′/N ′ is a trivial group, and hence |A′| = |N ′|. But this is impossible, since p divides |A′| but not |N ′|. The second part for p = 2 follows from the original proof of [12, Corollary 5.7]. Both applications of Corollary 5.7∗, namely the proofs of Theorem 6.2∗ and Theo- rem 9.1∗, only use the fact that the order of kerφ is divisible by the largest prime divisor of |A|. Since this follows by Corollary 3.3, both theorems still hold, and their proofs can be corrected by minor changes in their wording. 4 Quotients and their properties In this section, we will show that if BC is a complementary product of two cyclic groups with C core-free in BC, then not only C corresponds to some skew morphism of B (of order |C|), but also B corresponds to some skew morphism of C (of order smaller than |B|). Let φ be a skew morphism of a cyclic group B, let G = B⟨φ⟩, and let C = ⟨φ⟩. Also let b be a generator of B. To distinguish between φ as a permutation of B and φ as a generator of the cyclic group C, we use c in the latter case. Since C is core-free in G, by Theorem 2.1 we have |G : B| = |C| < |G : C| = |B| , so (again by Theorem 2.1) we find that B has a non-trivial core in G. Let K denote the core of B in G, and for every X ≤ G let X denote XK/K (∼= X/(X ∩ K)). Next, let H be a subgroup of B such that cHc−1 ⊆ B. Then, since B is cyclic, H is the unique subgroup of B of order |H|, and so cHc−1 = H . It follows that H is normal in G, and so it is contained in K. Moreover, since K is the core of B in G, we have cKc−1 = K ⊆ B, and hence by Lemma 2.4 we find that K = kerφ. Now we look closely at the product G = B C. First, noting that K ∩ C = {1} we have C ∼= C (and so C is cyclic, and hence abelian) and B ∩ C = {1}. Since B is cyclic, so is its quotient B, and by the definition of K we deduce that B is core-free in G. Now it is straightforward to check that the bijection that maps every element d ∈ C to the unique element d′ ∈ C such that d b = bjd′ defines a skew morphism of C, with power function π given by π(d) = j. (We choose b to be the image of b ∈ B under the natural homomorphism from B to B; since b is a generator of B, it follows that b is a generator of B.) A skew morphism of C (∼= C) constructed in the way described in the previous para- graph is called the quotient of φ (with respect to b), and will be denoted by φ. Since a cyclic group B can be generated by different elements, φ can have more than one quotient. 334 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 (In fact, by [4, Remark 5.1] this is always true unless B has a unique generator.) In the case of additive notation B = Zn, and unless otherwise specified, by the quotient of φ we understand the quotient with respect to 1. The above construction (proposed in the author’s PhD thesis [3]) was also introduced independently in [15], where it was noted that if φ is a skew morphism of Zn and φ is a quo- tient of φ, then (φ,φ) is an (ord(φ), n)-reciprocal pair of skew morphisms. A pair (φ, ρ) of skew morphisms of Zn and Zm with power functions π and τ is called (m,n)-reciprocal if ord(φ) dividesm, ord(ρ) divides n, and the congruences π(i) ≡ ρi(1) (mod ord(φ)) and τ(j) ≡ φj(1) (mod ord(ρ)) hold for each i ∈ Zn and j ∈ Zm. We note that while every pair (φ,φ) gives an (ord(φ), n)-reciprocal pair of skew morphisms, not every reciprocal pair arises in this way. For example, there exist (m,n)-reciprocal pairs of skew morphisms with m = n, but by Theorem 2.8 we know that ord(φ) is always strictly smaller than n. The following observation is an easy consequence of the fact that a skew morphism (of a cyclic group) and its quotient always give a reciprocal pair of skew morphisms. Lemma 4.1. Let φ be a skew morphism of Zn with power function π, and let φ be the quotient of φ with power function π. Then for every i ∈ N: (a) π(i) = φ i(1), so in particular ord(φ) = n/|kerφ|; and (b) φi(1) ≡ π(i) (mod n/|kerφ|). Proof. First, since (φ,φ) is an (ord(φ), n)-reciprocal pair of skew morphisms, we have π(i) ≡ φ i(1) (mod ord(φ)). Hence, since both π and φ are mappings into Zord(φ), it follows that π(i) = φ i(1). The second part of (a) follows from the fact that n/|kerφ| is the smallest non-zero integer in kerφ. Finally, (b) follows easily as π(i) ≡ φi(1) (mod ord(φ)) and ord(φ) = n/|kerφ|. Next we provide a lemma which shows that quotients can be used to check whether a skew morphism of a cyclic group is an automorphism or a coset-preserving skew morphism. Lemma 4.2. A skew morphism φ of Zn is coset-preserving if and only if the quotient φ of φ is an automorphism. Moreover, φ is proper if and only if φ is non-trivial. Proof. Let φ be a coset-preserving skew morphism of Zn. This is equivalent with φ(1) ≡ 1 (mod n/|kerφ|), which by Lemma 4.1(b) happens if and only if π(1) = 1. This proves the first part. The second part follows easily by Lemma 4.1(a) as φ ∈ Aut(Zn) if and only if π(1) = 1, and φ is trivial if and only if φ(1) = 1. We note that a part of the Lemma 4.2 was proved in [17], where it was shown that if φ is an automorphism, then φ is coset-preserving, but not the other way around. Also note that since the only skew morphism of Z2 is the identity mapping, it follows immediately from Lemma 4.2 that if a skew morphism of a cyclic group has order 2, then it must be an automorphism. (This is also an easy consequence of Lemma 2.4.) Another consequence of Lemma 4.2 is the fact that a proper skew morphism of Zn is coset-preserving if and only if the quotient of its quotient is the identity mapping. Somewhat interestingly, Lemma 4.2 also implies that by taking quotients every skew morphism of a cyclic group can be reduced to a non-trivial automorphism of the same group. M. Bachratý: Quotients of skew morphisms of cyclic groups 335 5 Using quotients to generate skew morphisms In this section, we describe an algorithm for finding recursively skew morphisms of cyclic groups based on various observations about the quotients of skew morphisms. 5.1 Skew morphisms with a given quotient First, we explain how to find all skew morphisms of a cyclic group with a given quotient. The following observation about quotients of skew morphisms of cyclic groups will be useful. Proposition 5.1. Let φ be a skew morphism of Zn with power function π, and let φ be the quotient of φ with power function π. Then: (a) the periodicity pφ of φ is the smallest generator of kerφ; (b) ord(φ) = |kerφ|pφ; and (c) the orbit T of ⟨φ⟩ that contains 1 is expressible in the form T = (x1, . . . , xpφ , ψ(x1), . . . , ψ(xpφ), . . . , ψ ord(ψ)−1(x1), . . . , ψ ord(ψ)−1(xpφ)), (5.1) for some coset-preserving skew morphism ψ of Zn such that ord(ψ) = ord(φ)/pφ andψ(1) ≡ 1 (mod n/|kerφ|), and with x1 = 1 and xi ≡ π(i−1) (mod n/|kerφ|) for each i ∈ {2, . . . , pφ}. Proof. First, by Theorem 2.10 we have pφ = p1. Then, since the values taken by π at any two elements of Zn are equal if and only if they belong to the same right coset of kerφ in Zn, it follows that p1 is the smallest positive integer such that 1 ≡ φp1(1) (mod n/|kerφ|), which by Lemma 4.1(b) is equivalent with 1 ≡ π(p1) (mod n/|kerφ|). Noting that n/|kerφ| = ord(φ), we deduce that p1 is the smallest positive integer such that π(p1) = 1, and (a) follows. Moreover, since pφ is the smallest non-trivial element of kerφ, which is a subgroup of Zord(φ), it follows that ord(φ) = |kerφ|pφ, which proves (b). To prove the final assertion, let T denote the orbit of ⟨φ⟩ that contains 1, and let ψ = φp1 . By Theorem 2.10 we know that ψ is a coset-preserving skew morphism of Zn, and by the definition of the periodicity we have ψ(1) ≡ 1 (mod n/|kerφ|). By Proposition 2.7 we know that the size of T is equal to ord(φ). Moreover, p1 divides |T | (see [6, Lemma 3.1]), and hence ord(ψ) = ord(φp1) = ord(φ)/p1, and the effect of ψ on T induces p1 cycles, each of length ord(φ)/p1. Finally, by Lemma 4.1(b) we have xi = φ i−1(1) ≡ π(i− 1) (mod n/|kerφ|) and the rest follows. Let ρ be a skew morphism of a cyclic group Zm. We will provide a detail explanation of the method for finding all skew morphisms φ of Zn with quotient ρ.1 First, we find the smallest positive integer j such that j ∈ ker ρ. Then by Propo- sition 5.1(a) we have p1 = pφ = j. Next, since ρ is a quotient of φ, it follows by 1It is important to emphasise here that this method finds all skew morphisms with a particular quotient only for a given cyclic group. If we do not restrict ourselves to a specific group, then in some cases we can find infinitely many skew morphisms with a given quotient. For example, using the classification of skew morphisms for cyclic p-groups of odd order (presented in [21]) it can be shown that each cyclic 3-group of order at least 9 admits a skew morphism whose quotient is the skew morphism (1, 3, 5) of Z6. 336 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 Lemma 4.1(a) that ord(ρ) = n/|kerφ|, and hence kerφ is the unique subgroup of Zn of or- der n/ord(ρ). As a next step we find all coset-preserving skew morphisms ψ of Zn satisfy- ing ord(ψ) = m/pφ andψ(1) ≡ 1 (mod n/|kerφ|) (note here thatm = |Zm| = ord(φ)). This allows us to identify all possible candidates for the orbit T of ⟨φ⟩ that contains 1, using (5.1). (For each choice of ψ, we have at most |kerφ|p1−1 = (n/ord(ρ))p1−1 candi- dates; the number of candidates could be smaller since sometimes (5.1) does not define a cycle on B.) Next, suppose that φ is a skew morphism, and let φ1 be the cyclic permutation of T induced by φ. Then by Lemma 4.1(a) we have π(i) = ρi(1), and hence by Lemma 2.13 we find that φ(i) = φ1(1) + φ ρ(1) 1 (1) + φ ρ2(1) 1 (1) + · · ·+ φ ρi−1(1) 1 (1) for all i ∈ N. (5.2) As a final step, for each candidate for φ1 we use (5.2) to define a function φ : Zn → Zn, and then check whether φ is a skew morphism of Zn, and T an orbit of ⟨φ⟩. It can be easily verified that if this is true, then ρ is the quotient of φ. Remark 5.2. Note that to use (5.2), we do not need to know the complete orbit T , but only the elements φ ρ i(1) 1 (1) for each i ∈ {0, 1, . . . , ord(ρ)}. In fact, since for every positive integer j we have φ j+p11 (1) = ψ(φ j 1 (1)), only elements of the form φ e 1 (1) with e ≡ ρi(1) (mod p1) are needed to define φ. In some cases, this significantly reduces the number of possible candidates for φ1. 5.2 Algorithm for finding all skew morphisms of a cyclic group We are ready to describe our algorithm for finding all skew morphisms of cyclic groups up to any order. This algorithm is recursive in the sense that it takes the sets of all skew morphisms of the groups Zm for m ∈ {2, 3, . . . , n − 1} as input and outputs all skew morphisms of Zn. Since the only skew morphism of Z2 is the identity permutation, the algorithm can be easily initialised. As the first step, we use the method presented in [6] to find all coset-preserving skew morphisms of Zn. Further details on this method are available in Section 6.3. Next, let φ be a skew morphism of Zn that is not coset-preserving, and let φ be the quotient of φ. Then by Lemma 4.2 we know that φ is a proper skew morphism of Zord(φ). Moreover, by Theorem 2.8 and Theorem 2.9 we find that ord(φ) < n, and that ord(φ) divides nϕ(n), and gcd(ord(φ), n) ̸= 1. Since we know all skew morphisms of Zm for each m < n, and we also know all coset-preserving skew morphisms of Zn, it follows that we can simply apply the method explained in the previous subsection to find all skew morphisms of Zn that are not coset-preserving. Together with the coset-preserving skew morphisms of Zn (which include all automorphisms and were found earlier), this gives all skew morphisms of Zn. A MAGMA [7] implementation of the described algorithm succeeded in finding all skew morphisms of cyclic groups of order up to 161. (The file listing all of these skew morphisms is available at [2].) This significantly improves the previous largest complete list [9] which goes up to the order 60. In Table 1 we summarise the information obtained about skew morphisms of cyclic groups Zn for n ≤ 161. We include a group in the table if and only if it admits a proper skew morphism. Moreover, if a listed group admits a proper skew morphism that is not coset-preserving, then the order of the group is preceded by M. Bachratý: Quotients of skew morphisms of cyclic groups 337 n Skew Classes n Skew Classes n Skew Classes 6 2 + 2 1 58 28 + 28 1 114 148 + 36 7 8 2 + 4 1 60 80 + 16 17 116 112 + 56 3 ∗9 4 + 6 2 62 30 + 30 1 ∗117 88 + 72 11 10 4 + 4 1 ∗63 44 + 36 7 118 58 + 58 1 12 4 + 4 2 ∗64 268 + 32 42 120 208 + 32 43 14 6 + 6 1 66 60 + 20 13 ∗121 900 + 110 90 16 12 + 8 4 68 64 + 32 3 122 60 + 60 1 ∗18 24 + 6 6 70 72 + 24 11 124 60 + 60 2 20 16 + 8 3 ∗72 156 + 24 36 ∗125 1568 + 100 152 21 12 + 12 1 74 36 + 36 1 ∗126 348 + 36 34 22 10 + 10 1 ∗75 96 + 40 24 ∗128 1132 + 64 114 24 16 + 8 7 76 36 + 36 2 129 84 + 84 1 ∗25 48 + 20 12 78 104 + 24 9 130 144 + 48 17 26 12 + 12 1 80 152 + 32 26 132 120 + 40 26 ∗27 64 + 18 20 ∗81 676 + 54 110 134 66 + 66 1 28 12 + 12 2 82 40 + 40 1 ∗135 256 + 72 80 30 24 + 8 7 84 104 + 24 14 136 228 + 64 10 ∗32 60 + 16 14 86 42 + 42 1 138 132 + 44 25 34 16 + 16 1 88 80 + 40 15 140 240 + 48 29 ∗36 48 + 12 12 ∗90 216 + 24 36 142 70 + 70 1 38 18 + 18 1 92 44 + 44 2 ∗144 552 + 48 96 39 24 + 24 1 93 60 + 60 1 146 72 + 72 1 40 44 + 16 9 94 46 + 46 1 ∗147 960 + 84 68 42 52 + 12 7 ∗96 272 + 32 58 148 144 + 72 3 44 20 + 20 2 ∗98 480 + 42 38 ∗150 648 + 40 74 ∗45 16 + 24 8 ∗99 40 + 60 20 152 144 + 72 23 46 22 + 22 1 ∗100 512 + 40 42 ∗153 64 + 96 32 48 64 + 16 20 102 96 + 32 19 154 180 + 60 17 ∗49 180 + 42 30 104 132 + 48 13 155 120 + 120 1 ∗50 152 + 20 18 105 48 + 48 4 156 352 + 48 22 52 48 + 24 3 106 52 + 52 1 158 78 + 78 1 ∗54 246 + 18 33 ∗108 492 + 36 66 ∗160 616 + 64 84 55 40 + 40 1 110 168 + 40 9 56 48 + 24 11 111 72 + 72 1 57 36 + 36 1 112 192 + 48 36 Table 1: Skew morphisms of cyclic groups of order n. the asterisk character (∗). All included groups are listed by their orders, and for each of them we provide the total number of skew morphisms (written as the sum of the numbers of proper skew morphisms and automorphisms), and the number of conjugacy classes of proper skew morphisms in Aut(Zn). Note that the automorphism group of a cyclic group Zn is always abelian, and hence the conjugation action of Aut(Zn) on itself is trivial. It follows that the number of conjugacy classes of Aut(Zn) is equal to |Aut(Zn)|. For this reason, we list the number of conjugacy classes only for proper skew morphisms. We also note that the numbers of skew morphisms in Table 1 for n ≤ 60 coincide with the numbers of skew morphisms in [9]. 5.3 Remarks concerning Table 1 An inspection of Table 1 suggests various interesting questions regarding skew morphisms of cyclic groups. Possibly the most natural question to ask here is which values n actually appear in Table 1 or, equivalently, which cyclic groups admit a proper skew morphism. 338 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 This was answered in [20] for cyclic groups, and later in [12] for all other abelian groups. Specifically, if an abelian group A does not admit any proper skew morphism, then A is cyclic of order n where n = 4 or gcd(n, ϕ(n)) = 1, or A is an elementary abelian 2-group. Next, we look at values n such that Zn admits (up to conjugacy in Aut(Zn)) only one proper skew morphism. In [20] this was shown to be true for all cases where n is a product of two distinct primes and gcd(n, ϕ(n)) > 1. The only other value n that appears in Table 1 and has this property is n = 8. An interesting question raised in this context is whether this covers all such values n, or if there are others. We are also interested in those cyclic groups that admit only coset-preserving skew morphisms. Unlike general skew morphisms, coset-preserving skew morphisms are well understood for cyclic groups, and, in particular, we can list all coset-preserving skew mor- phisms of Zn in polynomial time; see [6]. Thus, if for some n we can show that all skew morphisms of Zn are coset-preserving, then we can find all skew morphisms of Zn much faster than by using the algorithm explained in Section 5.2. In the following section we completely solve this problem by characterising all cyclic groups that admit only coset- preserving skew morphisms. 6 Cyclic groups that admit only coset-preserving skew morphisms In this section we focus on cyclic groups admitting only coset-preserving skew morphisms. Our main theorem is the following: Theorem 6.1. All skew morphisms of Zn are coset-preserving if and only if n = 2em with e ∈ {0, 1, 2, 3, 4} and m odd and square-free. Note that Theorem 6.1 include all groups Zn that does not admit any proper skew morphism, as in that case either n = 4, or (n, ϕ(n)) = 1, which forces n to be square-free (for if some prime square p2 divides n, then p is a common factor of n and ϕ(n)). In what follows, we will say that the positive integer n is resolvable if it is expressible in the form n = 2em with e ∈ {0, 1, 2, 3, 4} and m odd and square-free. The proof of Theorem 6.1 is split into two parts; in Section 6.1 we show that if n is not resolvable, then Zn admits a non-coset-preserving skew morphism, and in Section 6.2 we show that if n is resolvable, then Zn does not admit a non-coset-preserving skew morphism. Then in Section 6.3 we use Theorem 6.1 (and further facts about coset-preserving skew morphisms of cyclic groups) to enumerate all skew morphisms for many finite cyclic groups for which no such enumeration was available to date. Finally, in Section 6.4 we give an example that demonstrates how Theorem 6.1 can be applied to find a precise formula for the number of skew morphisms of Zn in the case when n is resolvable and has a relatively small number of prime factors. 6.1 Cyclic groups admitting non-coset-preserving skew morphisms Here we show that if the order of a cyclic group is divisible by 32 or by the square of an odd prime, then this group admits a skew morphism that does not preserve the cosets of its kernel. To do this, we will use some facts about skew morphisms of Zn that give rise to a regular Cayley map. First, we have the following: Proposition 6.2 ([18]). A skew morphism φ of a finite group B gives rise to a regular Cayley map for B if and only if the set of elements of some orbit of ⟨φ⟩ is closed under taking inverses and generates B. M. Bachratý: Quotients of skew morphisms of cyclic groups 339 We say that a skew morphism φ of a group B is t-balanced if its kernel has index 2 in B. The value t is given by t = π(a), where a is any element of B not contained in kerφ. In the special case when t = ord(φ)− 1 we say that φ is anti-balanced. For further information on t-balanced skew morphisms we refer the reader to [11]. The following observation shows that every coset-preserving skew morphism of Zn that gives rise to a regular Cayley map is either an automorphism of Zn, or a t-balanced skew morphism of Zn. Lemma 6.3. If φ is a coset-preserving skew morphism of Zn that gives rise to a regular Cayley map, then the index of kerφ in Zn is at most two. In particular, if n is odd, then kerφ = Zn and φ is an automorphism of Zn. Proof. Since φ gives rise to a regular Cayley map, by Proposition 6.2 there exists some orbit T of ⟨φ⟩ that is closed under taking inverses and generates Zn. Further, by [20, Corollary 3.3] we know that T contains some element t such that ⟨t⟩ = Zn, and since T = −T , we also have −t ∈ T . Next, from the fact that φ is coset-preserving we deduce that t and −t are both in the same coset of kerφ in Zn. It follows that 2t ∈ kerφ, and noting that t is a generator of Zn, we also have gcd(n, t) = 1. Hence 2 ∈ kerφ, and the rest follows. Throughout the proof of the following proposition we repeatedly refer to the classifica- tion of regular Cayley maps for cyclic groups given in [13]. Proposition 6.4. If a positive integer n is divisible by 32 or p2 for some odd prime p, then Zn admits a skew morphism that is not coset-preserving. Proof. First, assume that n is odd and divisible by p2 for some odd prime p. Then there exists a regular Cayley map for Zn with non-balanced representation (see [13, Section 8]), and hence there exists a proper skew morphism φ of Zn that gives rise to this Cayley map. Since φ is proper and n is odd, by Lemma 6.3 we deduce that φ is not coset-preserving. Next, if n is even and divisible by p2 for some odd prime p, then we have Zn = Zℓ×Z2e with ℓ odd. Since ℓ is clearly divisible by p2, from the previous paragraph we know that Zℓ admits a skew morphism φ that is not coset-preserving. By Lemma 2.11 there exists a skew morphism θ of Zn such that θ ↾Zℓ= φ and ker θ = kerφ×Z2e . Now it can be easily seen that since φ does not preserve the cosets of kerφ in Zℓ, the same is true for θ and cosets of ker θ in Zn. Finally, let n be even and divisible by 32, and consider the factorisation Zn = Z2e ×Zℓ with ℓ odd. Note that to show that Zn admits a non-coset-preserving skew morphism, it is sufficient to prove this for Z2e (and the rest will follow by Lemma 2.11). Let M(2m, r) be the regular Cayley map for Z2m given by [13, Definition 3.6]. This map is defined for every unit r modulo m such that if b is the largest divisor of m that is relatively prime to r − 1, then either b = 1, or r is a root of −1 modulo b of multiplicative order 2k where k is relatively prime to m/b. Let m = 2e−1, r = 2e−3 + 1, and M = M(2m, r). Note that the largest divisor of m relatively prime to r − 1 is 1, and hence b = 1. Also note that r is not a root of −1 modulo m, and since e ≥ 5 we have r2 ̸≡ 1 (mod m). It follows that M has no balanced, no t-balanced, and no anti-balanced representation; see [13, Section 8]. Since every automorphism of Z2e gives rise to a skew morphism with a balanced representation, and every skew morphism of Z2e with kernel of index 2 in Z2e gives rise to a skew morphism with either t-balanced or anti-balanced representation, we 340 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 deduce that a skew morphism φ of Z2e that gives rise toM has kernel of index greater than two in Z2e . Hence by Lemma 6.3 we find that φ is not coset-preserving. 6.2 Cyclic groups admitting only coset-preserving skew morphisms Next we show that if the positive integer n is resolvable, then all skew morphisms of Zn are coset-preserving. We start with the following technical lemma. Lemma 6.5. Let φ be a skew morphism of a cyclic group Zn, let N be any non-trivial subgroup of kerφ, let φ∗N be the skew morphism of Zn/N induced by φ, and let L/N be the kernel of φ∗N . Also let s be a prime factor of n/|kerφ|, let ks denote the largest power of s that divides |(kerφ)/N |, and let a = n/(s|kerφ|) be an element of Zn. If sks divides |L/N |, then a /∈ kerφ and a ∈ L. Proof. First, since a|kerφ| = n/s < n, we have a /∈ kerφ. (Note that this part is true regardless of whether sks divides |L/N |.) Next, let K = kerφ, and let m be an integer such that |K/N | = mks. (Observe that m and ks are relatively prime.) Then since K/N is a subgroup of L/N , we know that mks divides |L/N |. But |L/N | is also divisible by sks, and since gcd(m, ks) = 1, it follows that |L/N | must be divisible by msks. Hence |L| is divisible by s|K|, and therefore a ∈ L. We are now ready to prove the key part of the proof of Theorem 6.1. Proposition 6.6. Let n = 2em with e ∈ {0, 1, 2, 3, 4} and m odd and square-free. Then every skew morphism of Zn is coset-preserving. Proof. Suppose to the contrary that the assertion is not true, and let n be the smallest resolvable integer such that Zn admits a skew morphism φ that is not coset-preserving. Also let K = kerφ, let φ∗ denote the skew morphism of Zn/K induced by φ, and let φ be a quotient of φ. Since the only skew morphism of the trivial group is clearly coset- preserving, we have n > 1. Recalling that every skew morphism of a non-trivial group has a non-trivial kernel, it follows that |K| ≥ 2. In particular, it follows that |K| has at least one prime factor. We proceed by considering the two following cases: Case (a): |K| is a prime power Let p be the largest prime divisor of n. Then by Corollary 3.3 we know that p di- vides |K|. If p = 2, then n = 2, 4, 8, or 16, in which case Zn does not admit a skew morphism that is not coset-preserving; see [9]. Hence p is odd and |K| = p, and thus |Zn/K| = n/p. Then since p is the largest prime divisor of n, we know that p does not divide |Zn/K|ϕ(|Zn/K|), and it follows from Theorem 2.9 that p does not divide the order of φ∗. Further, by Lemma 3.1 we know that p divides ord(φ), and since ord(φ∗) = pφ, we deduce that p divides ord(φpφ). On the other hand, noting that φpφ preserves the cosets of K in Zn, we have ord(φpφ) ≤ |K| = p, and hence ord(φpφ) = p. Therefore ord(φ) = ppφ, and then by Proposition 5.1(b) we have |kerφ| = p. But p does not di- vide (n/p), which (by Lemma 4.1(a)) is the order of φ, and so by Lemma 3.1 we find that φ is a group automorphism. Hence by Lemma 4.2 we deduce that φ is coset-preserving, contradiction. Case (b): |K| has at least two distinct prime factors M. Bachratý: Quotients of skew morphisms of cyclic groups 341 Let k = |K|, and let d denote the integer satisfying n = kd. (Note that the elements of K are exactly the multiples of d modulo n.) Also let d = r1 . . . rℓ be a factorisation such that each factor is either an odd prime or the maximum possible power of two. Since φ does not preserve the cosets of K in Zn, we know that φ(1) ̸≡ 1 (mod d). Hence, noting that all factors ri for i ∈ {1, . . . , ℓ} are pairwise relatively prime, it follows from the Chinese Remainder Theorem that there exists r ∈ {r1, . . . , rℓ} such that φ(1) ̸≡ 1 (mod r). We will show that this cannot happen. First, assume that r is odd. It follows that r is a prime, and also that n is not divisible by r2. (And so, in particular, r does not divide |K|.) Let p be any prime factor of |K|, let N be the unique subgroup of K of order p, and let φ∗N be the skew morphism of Zn/N induced by φ. Also let L/N be the kernel of φ∗N , and suppose that |L/N | is not divisible by r. Since the order of the cyclic group Zn/N is clearly resolvable, by the assumption of minimality of n we know that φ∗N is coset-preserving. Then since r divides Zn/N but not L/N , we have φ(1) ≡ φ∗N (1) ≡ 1 (mod r). But this contradicts the fact that φ(1) ̸≡ 1 (mod r), and hence we deduce that r divides |L/N |. Now, since r does not divide |K|, it follows that r does not divide |K/N |, and thus we may use Lemma 6.5 (with s = r and ks = 1). Hence we deduce that the element a = d/r of Zn is not contained in K, but also a ∈ L, and by Lemma 2.6 it follows that pπ(a) ≡ p (mod ord(φ)). Since p was an arbitrary prime factor of |K|, the same is true also for some other prime factor q of |K|. (Here we use the assumption that the order of K has at least two distinct prime factors.) Hence we have pπ(a) ≡ p (mod ord(φ)), qπ(a) ≡ q (mod ord(φ)), for a pair of distinct primes p and q. Then since gcd(p, q) = 1, we deduce that π(a) ≡ 1 (mod ord(φ)), and consequently a ∈ K, contradiction. Next, assume that r is even. Again let p denote a prime factor of |K|, and define N , φ∗N , and L/N same as in the case when r was odd. Also let k2 denote the largest power of 2 that divides |K/N |, and suppose that |L/N | is not divisible by 2k2. Noting that K/N is a subgroup of L/N , it follows that the largest power of 2 that divides |L/N | must be k2. Then since φ∗N must be coset-preserving (due to the minimality of n) and the largest power of 2 that divides |Zn/N | is equal to rk2, we deduce that φ(1) ≡ φ∗N (1) ≡ 1 (mod r), contradicting the fact that φ(1) ̸≡ 1 (mod r). Hence it follows that 2k2 divides |L/N |, and Lemma 6.5 (in this case we take s = 2 and ks = k2) implies that the element a = d/2 of Zn satisfies a /∈ K and a ∈ L. Using the same argument as for r odd this again leads to a contradiction. Theorem 6.1 now follows directly from Propositions 6.4 and 6.6. 6.3 Enumeration In [6] it was shown that each coset-preserving skew morphism φ of Zn is uniquely de- termined by the following four parameters: the smallest non-zero element d of kerφ; the element h of Zn such that φ(1) = 1+h; the smallest positive integer s such that φ(d) = sd; and the positive integer e = π(1). (Note that s always exists since d ∈ kerφ and φ re- stricts to an automorphism of kerφ.) Using various properties of coset-preserving skew morphisms it can be checked that if φ is non-trivial, then the parameters d, h, s and e must satisfy the following properties (see [6, Section 4] for details): 342 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 (i) all four parameters are positive integers; (ii) d is a proper divisor of n; (iii) s < n/d and gcd(s, n/d) = 1; (iv) h is a multiple of d strictly smaller than n; (v) if r is the smallest positive integer such that h ∑r−1 i=0 s i ≡ 0 (mod n), then e is a (multiplicative) unit modulo r of order d and e < r; (vi) sd ≡ d−1∑ j=0 (1 + h ℓj∑ i=0 si) (mod n), where ℓj = ej − 1 mod r; and (vii) se−1 ≡ 1 (mod n/d). On the other hand, for each set of parameters (d, h, s, e) satisfying all of the above prop- erties there exists a unique non-trivial coset-preserving skew morphism of Zn (which can be constructed in a straightforward way) with this parameter set; see [6, Section 5]. This gives a one-to-one correspondence between non-trivial coset-preserving skew morphisms of Zn and the sets of parameters (d, h, s, e), and this correspondence can be used to find all coset-preserving skew morphisms of a given cyclic group in a polynomial time (in the cardinality of the group). Hence, by Theorem 6.1 we can quickly find all skew morphisms of Zn, where n is resolvable. Using the above observations, we developed an algorithm that can enumerate all skew morphisms for any given cyclic group of any resolvable order. A C++ implementation of this algorithm succeeded in enumerating all skew morphisms for cyclic groups of resolvable orders smaller than 10000 within a second, even without paral- lelisation. In comparison, the best available method to date for finding all skew morphisms for cyclic groups of general order (described in Section 5.2) is computationally feasible only up to order 161. In our enumeration, which is available at [1], we provide the total number of skew morphisms of Zn, and also the total number of automorphisms and their proportion among all skew morphisms. 6.4 Skew morphisms of Z4p Although coset-preserving skew morphisms can be generated efficiently, there is no known explicit formula for the number of coset-preserving skew morphisms of Zn for general n. Such a formula would be useful, and in the case when n is resolvable it would give the number of all skew morphisms of Zn. Resolvable integers n with the simplest structure (with respect to their prime factorisation) are primes, in which case all skew morphisms of Zn are automorphisms of Zn. The situation is also completely understood in the case when n is a product of two distinct primes p and q, in which case the number of skew morphisms of Zpq is (p−1)(q−1) if gcd(p, q−1) = gcd(p−1, q) = 1, and 2(p−1)(q−1) otherwise; see [20] for example. Here we go one step further and find a formula for the number of skew morphisms of Z4p, where p is an odd prime. We will use the following fact: Proposition 6.7. Let p be an odd prime. If φ is a proper skew morphism of Z4p, then the action of φ on its kernel is trivial. M. Bachratý: Quotients of skew morphisms of cyclic groups 343 Proof. Let π and K denote the power function and the kernel of φ, and let φ∗ be the skew morphism of Zn/K induced by φ. Note that by Theorem 6.1 we know that φ is coset- preserving, and so φ∗ must be trivial. Further, by Corollary 3.3 we know that p divides |K|, and since φ is proper it follows that the order of K is either p or 2p. We proceed by considering these two cases. In each case, we let T be the orbit of ⟨φ⟩ that contains 1. Note that by Proposition 2.7 we have |T | = ord(φ). Case (a): |K| = p Since φ is coset-preserving, we know that the cosets of K in Zn are preserved set-wise by φ. Note that only one element of the coset 1 +K does not generate Zn (either p or 3p, depending on whether p ≡ 1 (mod 4) or p ≡ 3 (mod 4)), and since by Lemma 2.12 no elements outside of K are fixed by φ, it follows that each orbit of ⟨φ⟩ on 1 +K generates Zn. Hence by Proposition 2.7 we know that all of these orbits have size ord(φ), and since |1 +K| = |K| = p, we see that ord(φ) = p. Since φ restricts to an automorphism of K, and ord(φ) = |K| = p, we deduce that the action of φ on K is trivial. Case (b): |K| = 2p In this case, since φ is proper, by Theorem 2.9 we have gcd(ord(φ), 4p) > 1. If the order of φ is odd, then we have ord(φ) = p or ord(φ) = 3p, but the latter case can be easily excluded since φ must preserve both cosets of K in Zn of size 2p. Next we deal with the case when ord(φ) is even. Note that by Lemma 4.1(a) we have π(1) = φ(1), and since φ is an automorphism of the cyclic group Zord(φ) of even order, we deduce that π(1) is odd. We will use this observation to show that both p and 3p are contained in some orbits of ⟨φ⟩ that generate Zn. Suppose to the contrary that this is not true. Since every element of 1 + K other than p and 3p generates Zn and no element of 1 +K is fixed by φ, we must have φ(p) = 3p and φ(3p) = p. Then since π(1) is odd, we have φπ(1)(p) = 3p, and therefore φ(1+p) = φ(1)+φπ(1)(p) = φ(1)+3p. On the other hand, we have φ(p+ 1) = φ(p) + φπ(p)(1) = 3p+ φπ(p)(1). But then φ(1) = φπ(p)(1), and since |T | = ord(φ) we find that π(p) = 1. This forces p ∈ K, contradicting the fact that the order of K is 2p. Hence we deduce that all orbits of ⟨φ⟩ on 1 +K generate Zn. In particular, ord(φ) divides 2p, and it follows that ord(φ) = 2p. We have shown that if |K| = 2p, then ord(φ) is equal to p or 2p. Hence by order considerations it can be easily seen that φ acts on K either trivially, or by the inversion. To exclude the latter, first note that 1 and φ(1) are in the same coset of K in Z4p, and hence φ(1)− 1 ∈ K. If we let h = φ(1)− 1, then we have φ2(1) = φ(h+1) = φ(h) +φ(1) = −h + h + 1 = 1, but then by Proposition 2.7 we have ord(φ) = 2, which is impossible. Hence we again conclude that the action of φ on K is trivial. Using Theorem 6.1 and Proposition 6.7 we can now easily enumerate all proper skew morphisms of Z4p. Theorem 6.8. If p is an odd prime, then the number of skew morphisms of Z4p is{ 6p− 6 if p ≡ 1 (mod 4) 4p− 4 if p ≡ 3 (mod 4). Proof. Throughout this proof we refer to the properties (i) to (vii) and the parameters d, h, s and e of coset-preserving skew morphisms for cyclic groups explained in Section 6.3. 344 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 We know that |Aut(Z4p)| = 2p − 2, so we proceed by counting proper skew morphisms of Z4p. Let φ be a proper skew morphism of Z4p, and recall that by Theorem 6.1 it is coset-preserving. Let d, h, s and e be the four defining parameters of φ, and note that by Proposition 6.7 we have s = 1. Since p is the largest prime divisor of 4p, by Corollary 3.3 we know that p divides |kerφ|, and it follows that d = 2 or d = 4. (The case d = 1 can be excluded as φ is proper.) First let d = 2, and let h be any positive multiple of 2 strictly smaller than 4p. If 4 divides h, then by (v) we find that r = p and e = p− 1. Since s = 1, both (iii) and (vii) are trivially true, and (vi) holds as (1+h)+(1+h(p−1)) ≡ 2+hp ≡ 2 (mod 4p). If 4 does not divide h and h ̸= 2p, then by (v) we have r = 2p and e = 2p− 1. Again both (iii) and (vii) hold trivially, and (vi) is also true since (1+h)+(1+h(2p−1) ≡ 2+2hp ≡ 2 (mod 4p). If h = 2p, then it can be easily verified that ord(φ) = r = 2, contradicting the fact that φ is proper. Since for all but one choice of h we obtain exactly one coset-preserving skew morphism, it follows that in this case we have exactly 2p− 2 skew morphism of Z4p. Next let d = 4, and let h be any positive multiple of 4 strictly smaller than 4p. Then by (v) we deduce that r = p, and e must be a fourth root of unity modulo p. It follows that necessarily p ≡ 1 (mod 4), in which case there are two possible candidates for e. Note that for either candidate we have e2 ≡ −1 (mod p) and e3 ≡ −e (mod p). Again (iii) and (vii) hold trivially, and (vi) holds as well since (1 + h) + (1 + he) + (1 + h(p− 1)) + (1 + h(p − e)) ≡ 4 + 2hp ≡ 4 (mod 4p). Hence, if p ≡ 1 (mod 4), then every choice of h gives two coset-preserving skew morphisms (one for each choice of e), which gives a total of 2p− 2 skew morphisms of Z4p. ORCID iDs Martin Bachratý https://orcid.org/0000-0002-4300-7507 References [1] M. Bachratý, Enumeration of skew morphisms of cyclic groups admitting only coset- preserving skew morphisms up to order 100000, https://drive.google.com/file/ d/1HpumLCS-hJW-LCdWbYsUJ-PCj3pVes4g. [2] M. Bachratý, List of skew-morphisms of cyclic groups up to order 161, https://drive. google.com/file/d/1rpFTIa961JkxHY5yuyN2gm2fUovQSuar. [3] M. Bachratý, Skew morphisms and skew product groups of finite groups, Ph.D. thesis, Univer- sity of Auckland, 2020, https://researchspace.auckland.ac.nz/bitstream/ handle/2292/53164/Bachraty-2020-thesis.pdf. [4] M. Bachratý, M. Conder and G. Verret, Skew product groups for monolithic groups, Algebr. Comb. 5 (2022), 785–802, doi:10.5802/alco.206, https://doi.org/10.5802/alco. 206. [5] M. Bachratý and R. Jajcay, Powers of skew-morphisms, Symmetries in Graphs, Maps, and Polytopes (2016), 1–25, doi:10.1007/978-3-319-30451-9_1, https://doi.org/10. 1007/978-3-319-30451-9_1. [6] M. Bachratý and R. Jajcay, Classification of coset-preserving skew-morphisms of finite cyclic groups, Australas. J. Comb. 67 (2017), 259–280, https://ajc.maths.uq.edu.au/ pdf/67/ajc_v67_p259.pdf. M. Bachratý: Quotients of skew morphisms of cyclic groups 345 [7] W. Bosma, J. Cannon and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), 235–265, doi:10.1006/jsco.1996.0125, https://doi.org/ 10.1006/jsco.1996.0125. [8] J. Chen, S. Du and C. H. Li, Skew-morphisms of nonabelian characteristically simple groups, J. Comb. Theory Ser. A 185 (2022), 105539, doi:10.1016/j.jcta.2021.105539, https://doi. org/10.1016/j.jcta.2021.105539. [9] M. Conder, List of skew-morphisms for small cyclic groups, https://www.math. auckland.ac.nz/~conder/SkewMorphisms-SmallCyclicGroups-60.txt. [10] M. Conder, R. Jajcay and T. W. Tucker, Regular Cayley maps for finite abelian groups, J. Algebr. Comb. 25 (2007), 343–364, doi:10.1007/s10801-006-0037-0, https://doi.org/ 10.1007/s10801-006-0037-0. [11] M. Conder, R. Jajcay and T. W. Tucker, Regular t-balanced Cayley maps, J. Comb. Theory Ser. B 97 (2007), 453–473, doi:10.1016/j.jctb.2006.07.008, https://doi.org/10.1016/j. jctb.2006.07.008. [12] M. Conder, R. Jajcay and T. W. Tucker, Cyclic complements and skew morphisms of groups, J. Algebra 453 (2016), 68–100, doi:10.1016/j.jalgebra.2015.12.024, https://doi.org/10. 1016/j.jalgebra.2015.12.024. [13] M. Conder and T. W. Tucker, Regular Cayley maps for cyclic groups, Trans. Amer. Math. Soc. 336 (2014), 3585–3609, doi:10.1090/s0002-9947-2014-05933-3, https://doi.org/10. 1090/s0002-9947-2014-05933-3. [14] S. Du, K. Hu. and A. Lucchini, Skew-morphisms of cyclic 2-groups, J. Group The- ory 22 (2019), 617–635, doi:10.1515/jgth-2019-2046, https://doi.org/10.1515/ jgth-2019-2046. [15] Y.-Q. Feng, K. Hu, R. Nedela, M. Škoviera and N.-E. Wang, Complete regular dessins and skew-morphisms of cyclic groups, Ars Math. Contemp. 18 (2020), 289–307, doi:10.26493/ 1855-3974.1748.ebd, https://doi.org/10.26493/1855-3974.1748.ebd. [16] K. Hu, Y. S. Kwon and J.-Y. Zhang, Classification of skew morphisms of cyclic groups which are square roots of automorphisms, Ars Math. Contemp. 21 (2021), 2–01, 23 pp., doi:10.26493/ 1855-3974.2129.ac1, https://doi.org/10.26493/1855-3974.2129.ac1. [17] K. Hu, R. Nedela, N.-E. Wang and K. Yuan, Reciprocal skew morphisms of cyclic groups, Acta Math. Univ. Comenian. 88 (2019), 305–318, http://www.iam.fmph.uniba.sk/ amuc/ojs/index.php/amuc/article/view/1006. [18] R. Jajcay and J. Širáň, Skew-morphisms of regular Cayley maps, Discrete Math. 244 (2002), 167–179, doi:10.1016/S0012-365X(01)00081-4, https://doi.org/10.1016/ S0012-365X(01)00081-4. [19] I. Kovács and Y. S. Kwon, Regular Cayley maps for dihedral groups, J. Comb. Theory Ser. B 148 (2021), 84–124, doi:10.1016/j.jctb.2020.12.002, https://doi.org/10.1016/j. jctb.2020.12.002. [20] I. Kovács and R. Nedela, Decomposition of skew-morphisms of cyclic groups, Ars Math. Contemp. 4 (2011), 329–349, doi:10.26493/1855-3974.157.fc1, https://doi.org/10. 26493/1855-3974.157.fc1. [21] I. Kovács and R. Nedela, Skew-morphisms of cyclic p-groups, J. Group Theory 20 (2017), 1135–1154, doi:10.1515/jgth-2017-0015, https://doi.org/10.1515/ jgth-2017-0015. [22] A. Lucchini, On the order of transitive permutation groups with cyclic point-stabilizer, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl. 9 (1998), 241–243. 346 Ars Math. Contemp. 24 (2024) #P2.08 / 327–346 [23] N.-E. Wang, K. Hu, K. Yuan and J.-Y. Zhang, Smooth skew morphisms of the dihedral groups, Ars Math. Contemp. 16 (2019), 527–547, doi:10.26493/1855-3974.1475.3d3, https: //doi.org/10.26493/1855-3974.1475.3d3. [24] J.-Y. Zhang and S. Du, On the skew-morphisms of dihedral groups, J. Group The- ory 19 (2016), 993–1016, doi:10.1515/jgth-2016-0027, https://doi.org/10.1515/ jgth-2016-0027. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.09 / 347–354 https://doi.org/10.26493/1855-3974.2751.81f (Also available at http://amc-journal.eu) Finite simple groups on triple systems* Xiaoqin Zhan , Xuan Pang , Suyun Ding † School of Science, East China JiaoTong University, Nanchang, 330013, People’s Republic of China Received 3 December 2021, accepted 14 May 2023, published online 22 November 2023 Abstract Let D be a triple system, and let G be a finite simple group. In this paper we almost determine all possibilities of D admitting G as its flag-transitive automorphism group. Keywords: Triple system, flag-transitivity, finite simple group. Math. Subj. Class. (2020): 05B07, 20B25, 05B25 1 Introduction A 2-(v, k, λ) design is a pair D = (P,B) where P is a set of v points and B is a collection of b k-subsets (blocks) of P with the property that every 2-subset of P occurs in λ blocks of B. If no blocks are identical, then D is called simple. An automorphism of a design D is a permutation of P which leaves B invariant. The full automorphism group of D, denoted by Aut(D), is the group consisting of all automor- phisms of D. A flag of D is a point-block pair (α,B) such that α ∈ B. For G ≤ Aut(D), G or D is called flag-transitive if G acts transitively on the set of flags, and point-primitive if G acts primitively on P . A set of blocks of D is called a set of base blocks with respect to an automorphism group G of D if it contains exactly one block from each G-orbit on the block set. In particular, if G is a flag-transitive automorphism group of D, then any block B is a base block of D. *The authors would like to express their gratitude to the referee who made very helpful comments and sugges- tions that improved our paper. This work is supported by the National Natural Science Foundation of China (Grant Nos. 12361004 and 11961026) and the Natural Science Foundation of Jiangxi Province (Grant Nos. 20224BAB211005 and 20224BAB201005). †Corresponding author. E-mail addresses: zhanxiaoqinshuai@126.com (Xiaoqin Zhan), p1443202623@163.com (Xuan Pang), dingsy2017@163.com (Suyun Ding) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 348 Ars Math. Contemp. 24 (2024) #P2.09 / 347–354 In this paper, we focus on simple 2-(v, 3, λ) designs also known as simple triple sys- tems, which can be denoted by TS(v, λ). One possibility is to take all possible 3-subsets of P however such designs are called complete and will be ignored. A triple system is a Steiner triple system, or STS(v), when λ = 1. Let r be the number of the blocks through a given point. For a TS(v, λ), it is well known that a necessary and sufficient condition for the existence of a TS(v, λ) is v ̸= 2 and λ ≡ 0 (mod (v − 2, 6)), and 3b = vr; (1.1) r = λ(v − 1) 2 ; (1.2) b = λv(v − 1) 6 ; (1.3) b ≥ v. (1.4) A 2-(v, k, 1) design is also called a finite linear space. A classic result is that of Higman and McLaughlin [8] who proved that for a finite linear space, flag-transitivity implies point- primitivity. Then Buekenhout, Delandtsheer and Doyen in [1] proved that if G acts flag- transitively on a linear space, then G is of affine or almost simple type. In 1990, the six-person team [2] classified all flag-transitive linear spaces apart from those with an one- dimensional affine automorphism group. For 2-(v, k, 1) designs with small values of k, one of the first classifications was for Steiner triple systems in [4], which considered what happens when the action was block- transitive but not 2-transitive on points. It is described in [11] what happens when the action on points is 2-transitive. This result depends on the classification of all finite simple groups and is subsumed into the general results proved by Kantor in [10]. Let G be a flag-transitive automorphism group of a TS(v, λ). It is shown in [6, 2.3.7(c), (e)] that G is point-primitive. Moreover, we can easily prove that G is 2-homogeneous (see Lemma 2.2 below). This result makes it possible to classify all flag-transitive triple systems using the classification of the finite 2-transitive permutation groups. Our main purpose is to give a classification of all triple systems admitting a simple flag-transitive automorphism group. We now state the main result of this paper: Theorem 1.1. Let D be a triple system, and let G be a finite simple group. If G acts flag-transitively on D, then one of the following LINES of Table 1 holds. Remark 1.2. • All but the triple systems listed in LINES 20 and 21 exist. • If G = PSU(3, q) with q = 5, then there are only two flag-transitive triple systems corresponding to LINES 19 and 20. • The existence of triple systems with 3 ∤ q and q ̸= 5 corresponding to LINES 20 and 21 is in doubt. X. Zhan et al.: Finite simple groups on triple systems 349 Table 1: G and corresponding triple systems. LINE G D Notes 1 A7 TS(15, 1) 2 TS(15, 12) 3 PSL(2, 11) TS(11, 3) 4 TS(11, 6) 5 HS TS(176, 12) 6 TS(176, 72) 7 TS(176, 90) 8 Co3 TS(276, 112) 9 TS(276, 162) 10 PSp(2d, 2) TS(2d−1(2d + 1), 22d−2) d ≥ 3 11 TS(2d−1(2d + 1), 2(2d−1 − 1)(2d−2 + 1)) 12 PSp(2d, 2) TS(2d−1(2d − 1), 22d−2) d ≥ 3 13 TS(2d−1(2d − 1), 2(2d−1 + 1)(2d−2 − 1)) 14 PSL(d, q) TS( q d−1 q−1 , q − 1) d ≥ 3 15 TS( q d−1 q−1 , qd−1 q−1 − q − 1) 16 PSL(2, q) TS(q + 1, q−1 2 ) q ≡ 1(mod 4) 17 Ree(q) TS(q3 + 1, 2(q − 1)) q = 32e+1 > 3 18 TS(q3 + 1, q − 1) 19 PSU(3, q) TS(q3 + 1, q − 1) q ≥ 3 20 TS(q3 + 1, q 2−1 (3,q+1) ) 21 TS(q3 + 1, 2(q 2−1) (3,q+1) ) 2 Useful lemmas The notation and terminology used is standard and can be found in [5, 6] for design theory and in [7, 9] for group theory. In particular, if G is a permutation group on a set Ω, and {α, β} ⊆ ∆ ⊆ Ω, then Gα denotes the stabilizer of a point α in G, and Gαβ denotes the pointwise stabilizer of two points α and β in G, and G∆ denotes the setwise stabilizer of ∆ in G. The following result about flag-transitive 2-designs is well-known. Lemma 2.1. Let D = (P,B) be a 2-(v, k, λ) design, and let G be an automorphism group of D. For any α ∈ P and B ∈ B, G is flag-transitive if and only if G is point-transitive and Gα is transitive on the pencil P (α) (the set of blocks through α), if and only if G is block-transitive and GB is transitive on the points of B. Lemma 2.2. Let D = (P,B) be a triple system, and let G be a flag-transitive automor- phism group of D. If G is a simple group, then G acts 2-transitively on P . Proof. Let {α, β} and {γ, δ} be arbitrary two unordered pairs of P . By the definition of a triple system, there are two points ε and θ such that B1 = {α, β, ε} and B2 = {γ, δ, θ} are two blocks of D. The flag-transitivity of G implies that there is a g ∈ G such that (ε,B1) g = (εg, Bg1 ) = (θ,B2), and so {α, β}g = {γ, δ}. Thus G is 2-homogeneous. If G is a simple group, then G acts 2-transitively on P by [7, Theorem 9.4B]. 350 Ars Math. Contemp. 24 (2024) #P2.09 / 347–354 Lemma 2.3. Let D = (P,B) be a triple system, and let G ≤ Aut(D) be a 2-transitive group on P . Then the following conditions are equivalent: (i) G acts flag-transitively on D. (ii) If B = {α, β, γ} ∈ B, then {{α, β, γi} | γi ∈ γG{α,β}} is the set of all blocks through points α and β. Proof. (i) ⇒ (ii): Let B(α, β) = {B1, B2, . . . , Bλ} be the set of blocks through points α and β, where Bi = {α, β, γi}, γi ∈ P \ {α, β}. Clearly, B(α, β)G{α,β} = B(α, β). If G acts flag-transitively on D, then for any two flags (γi, Bi) and (γj , Bj), there is a g ∈ G such that (γi, Bi)g = (γj , Bj), so γ g i = γj and g ∈ G{α,β}. Thus G{α,β} acts transitively on B(α, β) and hence {γ1, . . . , γλ} = γ G{α,β} i . (ii) ⇒ (i): Let (γ,B) and (ϵ, C) be two flags of D with B = {α, β, γ}, C = {δ, η, ϵ}. By the 2-transitivity of G, there exists g1 ∈ G such that {α, β}g1 = {δ, η}, thus Bg1 = {δ, η, γg1} is a block containing δ and η. Since {{δ, η, ϵi} | ϵi ∈ ϵG{δ,η}} is the set of all blocks through δ and η, there exists g2 ∈ G{δ,η} such that γg1g2 = ϵ, and then (γ,B)g1g2 = (ϵ, C). Therefore, G acts flag-transitively on D. Corollary 2.4. Let G be a 2-transitive group on a point set P with |P| = v, and let λ1, λ2, . . . , λk be all sizes of orbits of Gαβ on P \ {α, β}. If λi ̸= λj for i ̸= j, then there exist k different flag-transitive TS(v, λi). Proof. Without loss of generality, let ∆ = γGαβ with |∆| = λ1, where γ ∈ P \ {α, β}. Since Gαβ ⊴ G{α,β}, the group Gαβ acts 12 -transitively on γ G{α,β} , that is, Gαβ-orbits on γG{α,β} have the same length. The uniqueness of the Gαβ-orbit with size λ1 implies that γGαβ = γG{α,β} . Thus G{α,β} has a unique orbit with size λ1. Let B = {α, β, γ} and B = BG. We shall prove below that D = (P,B) is a TS(v, λ1) admitting G as its flag-transitive automorphism group. Since G is 2-transitive, for any pair {δ, η}, there exists g ∈ G such that {α, β}g = {δ, η}. So G{δ,η} has a unique orbit ∆g = (γg)G{δ,η} with |∆g| = |∆| = λ1. Let B(δ, η) be the set of elements of B containing δ, η with |B(δ, η)| = λ. It is easy to see that Λ = {{δ, η, ϵ} | ϵ ∈ ∆g} ⊆ B, so we have λ ≥ λ1. On the other hand, for C = {δ, η, θ} ∈ B(δ, η), there exists h ∈ G such that C = Bh. As |γGαβ | = |αGγβ | = |βGαγ | = λ1, we may assume that θ = γh. Then |θG{δ,η} | = |γhG{δ,η} | = |γG{α,β}h| = |∆h| = λ1, it implies λ1 ≥ λ. Thus, λ = λ1 and B(δ, η) = Λ. Hence D is a TS(v, λ1), and G is a flag-transitive automorphism group of D by Lemma 2.3(ii). Lemma 2.5. Let G be a 2-transitive group on a point set P with |P| = v, and let ∆ = {α, β, γ} be a 3-subset of P . If Gαβ is a cyclic group of order λ and |γGαβ | = λ, then (i) D = (P,∆G) is a flag-transitive TS(v, λ) if and only if G∆∆ ∼= S3, or (ii) D = (P,∆G) is a flag-transitive TS(v, 2λ) if and only if G∆∆ ∼= Z3. Proof. Here we only prove case (i), and case (ii) can be proved by same procedure. Since Gαβ is a cyclic group for any points α and β, we have that G∆ = G∆∆. Let D = (P,∆G). If D is a flag-transitive TS(v, λ), then using Lemma 2.1 and Equation (1.3), we have that b = λv(v − 1) 6 = |∆G| = [G : G∆] = [G : Gαβ ][Gαβ : G∆]. X. Zhan et al.: Finite simple groups on triple systems 351 By 2-transitivity of G and |Gαβ | = λ, we obtain |G∆| = 6. The flag-transitivity of G implies that G∆ acts transitively on the points of ∆ by Lemma 2.1. Thus G∆ ∼= S3. If G∆ ∼= S3, then G{α,β}γ ∼= Z2 and |∆G| = [G : Gαβ ][Gαβ : G∆] = λv(v−1)6 . Thus, D is a TS(v, λ) as G acts 2-transitively on P . Clearly, |γG{α,β} | = [G{α,β} : G{α,β}γ ] = λ, where G{α,β}γ = G{α,β} ∩Gγ . Therefore, G acts flag-transitively on D by Corollary 2.4. Lemma 2.6. Let G = Ree(q) act 2-transitively on Ω, where |Ω| = q3+1 and q = 32e+1 > 3. Then there exist subsets ∆, Σ of size 3 such that G∆∆ = Z3, GΣΣ = S3. Proof. Let Q be a Sylow 3-subgroup of G. Then |Q| = q3, and there exists α ∈ Ω such that Q is regular on Ω\{α}. Thus each subgroup of Q is semiregular on Ω\{α}. Let x, y ∈ Q such |x| = |y| = 3, x /∈ Z(Q) and y ∈ Z(Q), where the centre Z(Q) is elementary abelian of order q. Let ∆ be an orbit of ⟨x⟩. Then |∆| = 3 and G∆∆ = Z3 or S3. Further, since x is not conjugate to x−1 in G (reference [12]), we have G∆∆ ∼= ⟨x⟩ ∼= Z3. Consider y acting on Ω \ {α}. Since y is in the centre Z(Q), there is an involution z ∈ Gα such that yz = y−1, and the subgroup H = ⟨y, z⟩ ∼= S3. Since ⟨y⟩ is semiregular on Ω \ {α}, the set Ω \ {α} is divided into 13q 3 orbits of ⟨y⟩: ∆1,∆2, . . . ,∆m, where m = 13q 3 is odd. Since each H-orbit Σ contains a ⟨y⟩-orbit, the cardinality |Σ| = 3 or 6. As the number 13q 3 of ⟨y⟩-orbits is odd, it follows that there is at least one H-orbit Σ on Ω \ {α} has length 3. Therefore, GΣΣ = HΣΣ = S3 with |Σ| = 3. 3 Proof of Theorem 1.1 Let D = (P,B) be a TS(v, λ), and let G be a simple group acting flag-transitively on D. Then G acts 2-transitively on P by Lemma 2.2. Since we neglect the case D is complete, we may assume that G is not 3-homogeneous group on P . Thus, all such groups are known and we can find a classification in [3] and we have that G must be one of the following Table 2. We will prove Theorem 1.1 by analyzing the 11 cases in Table 2 one by one. Proof of Theorem 1.1. Let α and β be two points of P . For Cases 1 – 7, we have the following facts by the proof of [10, Theorem 1]: If G = A7 and v = 15, then Gαβ has orbit-lengths 1 and 12 on P \ {α, β}. If G = PSL(2, 11) and v = 11, then Gαβ has orbit-lengths 3 and 6 on P \ {α, β}. If G = HS and v = 176, then Gαβ has orbit-lengths 12, 72 and 90 on P \ {α, β}. If G = Co3 and v = 276, then Gαβ has orbit-lengths 112 and 162 on P \ {α, β}. If G = PSp(2d, 2) and v = 22d−1 + 2d−1, then Gαβ has orbit-lengths 2(2d−1 − 1)(2d−2 + 1) and 22d−2 on P \ {α, β}. 352 Ars Math. Contemp. 24 (2024) #P2.09 / 347–354 Table 2: 2-transitive, not 3-homogeneous simple groups. Case Group Degree Notes 1 A7 15 2 PSL(2, 11) 11 3 HS 176 4 Co3 276 5 PSp(2d, 2) 22d−1 + 2d−1 d ≥ 3 6 PSp(2d, 2) 22d−1 − 2d−1 d ≥ 3 7 PSL(d, q) (qd − 1)/(q − 1) d ≥ 3 8 PSL(2, q) q + 1 q ≡ 1 (mod 4) 9 Suz(q) q2 + 1 q = 22e+1 > 2 10 Ree(q) q3 + 1 q = 32e+1 > 3 11 PSU(3, q) q3 + 1 q ≥ 3 If G = PSp(2d, 2) and v = 22d−1 − 2d−1, then Gαβ has orbit-lengths 2(2d−1 + 1)(2d−2 − 1) and 22d−2 on P \ {α, β}. If G = PSL(d, q) with d ≥ 3 and v = q d−1 q−1 , Gαβ has orbit-lengths q − 1 and qd−1 q−1 − q − 1 on P \ {α, β}. It follows from Corollary 2.4 that D is one of triple systems corresponding LINES 1-15 in Table 1. Case 8: G = PSL(2, q) with q ≡ 1 (mod 4) and v = q + 1. In this case, there are exactly two G-orbits on 3-subsets of q + 1 points with size q(q 2−1) 12 . Also, Gαβ ∼= Z q−1 2 has two orbits with length q−12 on P \ {α, β}, denoted by Γ1 and Γ2. Suppose that Γ1 = {α1, α2, . . . , α q−1 2 }, Γ2 = {β1, β2, . . . , β q−1 2 }. For i ∈ {1, 2}, let Di = (P,∆Gi ) where ∆i = {α, β, γi} and γi ∈ Γi. It is easy to calculate that |G∆i | = 6, and hence G∆i ∼= S3. By Lemma 2.5(i), both D1 and D2 are TS(q + 1, q−12 ). Let g = (α, β)(α1, β1) · · · (α q−1 2 , β q−1 2 ). Clearly, g is an isomorphism from D1 to D2, that is D1 ∼= D2. Thus, D is a TS(q+1, q−12 ). Case 9: G = Sz(q) and v = q2 + 1. Since G acts flag-transitively on D, then 3 | |G| by Lemma 2.1. But this contracts the fact that 3 ∤ |G| (see [9, Theorem 3.6]). Therefore, there is no triple system admitting Sz(q) as its flag-transitive automorphism group. Case 10: G = Ree(q) and v = q3 +1 with q = 32e+1 > 3. From Lemmas 2.5 and 2.6, we have that D is one of triple systems corresponding LINES 17 and 18 in Table 1. Case 11: G = PSU(3, q) and v = q3 + 1. Since Gαβ ∼= Z q2−1 (3,q+1) has a unique orbit O with size q − 1 and q(3, q + 1) orbits with size q 2−1 (3,q+1) . Similar to proof of Lemma 2.4, we can prove that there exists a unique TS(q3 +1, q− 1) admitting G as its flag-transitive automorphism group. If q = 3e ≥ 3, there exist subsets ∆, Σ of size 3 such that G∆∆ = Z3, GΣΣ = S3 by the same proof as Lemma 2.6. In this case, D is one of triple systems corresponding LINES 20 and 21 in Table 1 from Lemma 2.5. X. Zhan et al.: Finite simple groups on triple systems 353 If q = 5 then D can only be a flag-transitive TS(126, 8) in addition to TS(126, 4) by a simple calculation. This means that there is no flag-transitive TS(126, 16) in this case. Unfortunately, we don’t know whether Lemma 2.6 holds when 3 ∤ q. Thus the existence of TS(q3 + 1, q 2−1 (3,q+1) ) (or TS(q 3 + 1, 2(q 2−1) (3,q+1) )) with 3 ∤ q and q ̸= 5 is in doubt. This completes the proof of Theorem 1.1. Conjecture 3.1. Let D be a triple system TS(q3 + 1, λ), and let G = PSU(3, q) act flag-transitively on D with 3 ∤ q and q ̸= 5. If λ ̸= q − 1 then one of following holds: (i) If q is even, then λ = 2(q 2−1) (3,q+1) . (ii) If q is odd, then λ = q 2−1 (3,q+1) or 2(q2−1) (3,q+1) . In fact, using MAGMA, we have already proved that the conjecture holds when q ≤ 100. ORCID iDs Xiaoqin Zhan https://orcid.org/0000-0003-0669-6419 Xuan Pang https://orcid.org/0000-0003-2500-9741 Suyun Ding https://orcid.org/0000-0002-6564-4427 References [1] F. Buekenhout, A. Delandtsheer and J. Doyen, Finite linear spaces with flag-transitive groups, J. Comb. Theory, Ser. A 49 (1988), 268–293, doi:10.1016/0097-3165(88)90056-8, https: //doi.org/10.1016/0097-3165(88)90056-8. [2] F. Buekenhout, A. Delandtsheer, J. Doyen, P. B. Kleidman, M. W. Liebeck and J. Saxl, Linear spaces with flag-transitive automorphism groups, Geom. Dedicata 36 (1990), 89–94, doi:10. 1007/bf00181466, https://doi.org/10.1007/bf00181466. [3] P. J. Cameron, Finite permutation groups and finite simple groups, Bull. Lond. Math. Soc. 13 (1981), 1–22, doi:10.1112/blms/13.1.1, https://doi.org/10.1112/blms/13.1.1. [4] P. C. Clapham, Steiner triple systems with block-transitive automorphism groups, Discrete Math. 14 (1976), 121–131, doi:10.1016/0012-365x(76)90055-8, https://doi.org/10. 1016/0012-365x(76)90055-8. [5] C. J. Colbourn and J. H. Dinitz (eds.), The CRC Handbook of Combinatorial Designs, Discrete Math. Appl. (Boca Raton), Chapman & Hall/CRC, Boca Raton, FL, 2nd edition, 2007, doi: 10.2307/3618812, https://doi.org/10.2307/3618812. [6] P. Dembowski, Finite Geometries, Classics in Mathematics, Springer-Verlag, New York, 1968, doi:10.1007/978-3-642-62012-6, https://doi.org/10.1007/ 978-3-642-62012-6. [7] J. D. Dixon and B. Mortimer, Permutation Groups, Graduate Texts in Mathematics, Springer-Verlag, New York, 1996, doi:10.1007/978-1-4612-0731-3, https://doi.org/ 10.1007/978-1-4612-0731-3. [8] D. G. Higman and J. E. McLaughlin, Geometric ABA-groups, Ill. J. Math. 5 (1961), 382–397, doi:10.1215/ijm/1255630883, https://doi.org/10.1215/ijm/1255630883. [9] B. Huppert and N. Blackburn, Finite Groups II, Springer-Verlag, New York, 1982, doi:10.1007/ 978-3-642-67994-0, https://doi.org/10.1007/978-3-642-67994-0. 354 Ars Math. Contemp. 24 (2024) #P2.09 / 347–354 [10] W. M. Kantor, Homogeneous designs and geometric lattices, J. Comb. Theory, Ser. A 38 (1985), 66–74, doi:10.1016/0097-3165(85)90022-6, https://doi.org/10.1016/ 0097-3165(85)90022-6. [11] J. Key and E. Shult, Steiner triple systems with doubly transitive automorphism groups: A corollary to the classification theorem for finite simple groups, J. Comb. Theory Ser. A. 36 (1984), 105–110, doi:10.1016/0097-3165(84)90082-7, https://doi.org/10.1016/ 0097-3165(84)90082-7. [12] H. N. Ward, On Ree’s series of simple groups, Bull. Am. Math. Soc. 69 (1963), 113–114, doi:10.1090/S0002-9904-1963-10885-X, https://doi.org/10.1090/ S0002-9904-1963-10885-X. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P2.10 / 355–383 https://doi.org/10.26493/1855-3974.2763.1e6 (Also available at http://amc-journal.eu) There is a unique crossing-minimal rectilinear drawing of K18* Bernardo M. Ábrego , Silvia Fernández–Merchant Departament of Mathematics, California State University at Northridge, CA, United States Oswin Aichholzer Institute for Software Technology, University of Technology, Graz, Austria Jesús Leaños † Academic Unit of Mathematics, Autonomous University of Zacatecas, Mexico Gelasio Salazar Institute of Physics, Autonomous University of San Luis Potosi, Mexico Received 8 December 2021, accepted 15 June 2023, published online 14 February 2024 Abstract We show that, up to order type isomorphism, there is a unique crossing-minimal recti- linear drawing of K18. It is easily verified that this drawing does not contain any crossing- minimal drawing of K17. Therefore this settles, in the negative, the following question from Aichholzer and Krasser: is it true that, for every integer n ≥ 4, there exists a crossing- minimal drawing of Kn that contains a crossing-minimal drawing of Kn−1? Keywords: Rectilinear crossing number, complete graphs, k-edges. Math. Subj. Class. (2020): 05C10, 05C60 *We thank an anonymous referee for carefully reading an earlier version of this paper, and providing several insightful comments, corrections, and suggestions. †Corresponding author. E-mail addresses: bernardo.abrego@csun.edu (Bernardo M. Ábrego), silvia.fernandez@csun.edu (Silvia Fernández–Merchant), oaich@ist.tugraz.at (Oswin Aichholzer), jleanos@uaz.edu.mx (Jesús Leaños), gsalazar@ifisica.uaslp.mx (Gelasio Salazar) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 356 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 1 Introduction The rectilinear crossing number cr(G) of a graphG is the minimum number of edge cross- ings in a rectilinear drawing of G in the plane, i.e., a drawing of G in the plane where the vertices are points in general position and the edges are straight line segments. A drawing of G with exactly cr(G) crossings is crossing-minimal. Determining the rectilinear crossing number cr(Kn) of the complete graph Kn is a well-known open problem in combinatorial geometry (see for instance [5, 11]). In [9] Aichholzer et al. determined the exact values of cr(Kn) for 13 ≤ n ≤ 17. In that paper also the following question was raised. Question 1.1. Is it true that, for every integer n ≥ 4, there exists a crossing-minimal drawing of Kn that contains a crossing-minimal drawing of Kn−1? The exact value of cr(Kn) is known for n ≤ 27 and n = 30 (see [3, 7, 8, 9, 10]). The value of cr(K18) = 1029 was established in [8]. Crossing-minimal rectilinear drawings of Kn for this range of values of n can be found in [2] and [6]. In particular, from [6], we know that there are at least 37269 non-isomorphic crossing-minimal drawings of K17. Let θ denote the counterclockwise rotation of 2π/3 around the origin, and let W := {(−51, 113), (6, 834), (16, 989), (18, 644), (18, 1068), (22, 211)}. From [2], we know that the 18-point set W ∪ θ(W ) ∪ θ2(W ) induces a crossing-minimal drawing of K18. See Figure 1 for an illustration of such a point set. Our main result is the following. Theorem 1.2. Up to order type isomorphism, there is a unique 18-point set whose induced rectilinear drawing of K18 has cr(K18) crossings. Let D be the (unique, in view of Theorem 1.2) crossing-minimal rectilinear drawing of K18. It is easily verified that every subdrawing of D with 17 points has more than cr(K17) = 798 crossings. This settles Question 1.1 in the negative. In the next section, we introduce the necessary notation and additional concepts re- quired for the proof of Theorem 1.2. In Section 4 we prove Theorem 1.2. 2 k-edges, (≤ k)-edges, and 3-decomposability Throughout this section, Q is a set of n ≥ 3 points in general position in the plane. If p and q are distinct points of Q, then we denote by pq the directed line spanned by p and q, directed from p towards q. Furthermore, pq+ and pq− denote the set of points in Q on the right and left, respectively, of pq. Thus Q = pq− ∪ {p, q} ∪ pq+ for all p, q ∈ Q with p ̸= q. Let k ∈ {0, 1, . . . , ⌊n/2⌋−1}. A k-edge ofQ is a directed line spanned by two distinct points of Q, which leaves exactly k points of Q on one side. A (≤ k)-edge (respectively, a (> k)-edge) is an i-edge of Q with 0 ≤ i ≤ k (respectively, k < i ≤ ⌊n/2⌋ − 1). Let Ek(Q), E≤k(Q), and E>k(Q) denote, respectively, the set of k-edges, (≤ k)-edges and (> k)-edges of Q. We use ek(Q), e≤k(Q), and e>k(Q) to denote, respectively, the number of elements in Ek(Q), E≤k(Q), and E>k(Q). Then e≤k(Q) = ∑k j=0 ej(Q) and e>k(Q) = ( n 2 ) − e≤k(Q). The vector E≤k(Q) := (e≤0(Q), e≤1(Q), . . . , e≤⌊n/2⌋−1(Q)) is the (≤ k)-edges vec- tor of Q. Finally, e≤k(n) denotes the minimum of e≤k(P ) taken over all n-point sets P B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 357 Figure 1: This is the 18-point set produced by the union of W = {(−51, 113), (6, 834), (16, 989), (18, 644), (18, 1068), (22, 211)}, θ(W ) and θ2(W ). It is not difficult to see that P produces a crossing-minimal rectilinear drawing ofK18. The triangle and the six straight line segments show that P is 3-decomposable. in the plane in general position. The exact determination of e≤k(n) is another well known open problem in combinatorial geometry (see for instance [3, 4, 7, 8]). The number of crossings in a rectilinear drawing of Kn and the number of k- and (≤ k)-edges in its underlying n-point set P are closely related by the following equality, independently proved in [4] and [12]: cr(P ) = ⌊n/2⌋−2∑ k=0 (n− 2k − 3) e≤k (P )− 3 4 ( n 3 ) + ( 1 + (−1)n+1 ) 1 8 ( n 2 ) . (2.1) This equality allows us to fully determine the (≤ k)-edges vector of any 18-point set whose induced drawing attains the rectilinear crossing number of K18. Proposition 2.1. If P is an 18-point set such that cr(P ) = cr(K18), then E≤k(P ) = (3, 9, 18, 30, 45, 63, 87, 120, 153). Proof. Let Q be an 18-point set in the plane in general position. It is known (see [3] or [7]) that E≤k(Q) = (e≤0(Q), e≤1(Q), . . . , e≤8(Q)) is bounded below entry-wise by (3, 9, 18, 30, 45, 63, 87, 120, 153). On the other hand, from (2.1) we know that cr(Q) = −612 + 15 · e≤0 (Q) + 13 · e≤1 (Q) + · · ·+ 1 · e≤7 (Q) . From the coefficients of this equation and the fact that e≤8 (Q) = 153, it follows that if e≤k (Q) is greater than the k-th component in the vector (3, 9, 18, 30, 45, 63, 87, 120, 153), then cr(Q) > 1029. 358 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Finally, we introduce a concept that captures a property shared by all known crossing- minimal rectilinear drawings ofKn, for n a multiple of 3. A point setQ is 3-decomposable if it can be partitioned into three equal-size sets A,B and C, such that (i) there exists a triangle T enclosing the point set Q; and (ii) the orthogonal projection of Q onto the three sides of T shows A between B and C on one side, B between C and A on the second side, and C between A and B on the third side. In such a case, we say that {A,B,C} is a 3-decomposition of Q. For instance, {W, θ(W ), θ2(W )} is a 3-decomposition of the 18-point set shown in Figure 1. As in [2], if {A,B,C} is a 3-decomposition of Q, we define two types of edges. Let p and q be distinct points of Q. If p, q ∈ A, p, q ∈ B or p, q ∈ C then we call pq monochromatic; otherwise, pq is bichromatic. Let Emonk (Q) and E bi k(Q) denote the set of monochromatic and bichromatic k-edges of Q, respectively. As before, we use emonk (Q) and ebik(Q) to denote |Emonk (Q)| and |Ebik(Q)|, respectively. Note that ek(Q) = emonk (Q)+ ebik(Q). Now we partition the monochromatic edges of Q into three types. If p, q ∈ A, then we say that pq is an edge of type aa. Similarly, we define the edges of types bb and cc. For x ∈ {a, b, c}, we denote the number of monochromatic k-edges of type xx by exxk (Q). Then emonk (Q) = e aa k (Q) + e bb k (Q) + e cc k (Q). 3 Overview of the proof of Theorem 1.2 For the rest of this paper, P is an 18-point set in the plane in general position where the rectilinear crossing number of K18 is attained. That is, cr(P ) = cr(K18). The first step in the proof, carried out in Section 4.1, consists of giving an algo- rithm that yields a canonical, unambiguous labelling of the points in P . The 18 points in P get labelled x0, x1, . . . , x5, y0, y1, . . . , y5, z0, z1, . . . , z5. Thus P gets naturally par- titioned into three sets X = {x0, x1, x2, x3, x4, x5}, Y = {y0, y1, y2, y3, y4, y5}, and Z := {z0, z1, z2, z3, z4, z5}. As we shall prove shortly afterwards, {X,Y, Z} happens to be a 3-decomposition of P . Once we have laid out the foundation by giving a canonical labelling of the points of P , the rest of the proof consists of showing the following: Lemma 3.1 (Implies Theorem 1.2). For each pair of distinct points p, q ∈ P , the set pq+ is uniquely determined. Clearly Lemma 3.1 implies Theorem 1.2: if the lemma holds, then the unambiguity of the labelling of the points in P implies that P is unique up to order type isomorphism. First we establish the lemma for the case in which pq is a (≤5)-edge. This is actually done in Section 4.1, where we give the algorithm to label the points in P . Indeed, the unambiguity in the labelling of the points in P is established in Proposition 4.3(1), and in order to prove this we need to prove simultaneously Proposition 4.3(2), which in particular implies Lemma 3.1 for the case in which pq is a (≤5)-edge. We then move on to proving Lemma 3.1 for the case in which pq is a (>5)-edge, that is, when pq is either a 6-edge, or a 7-edge, or an 8-edge. As we shall see, even if this follows from elementary observations, the investigation of these cases is remarkably more involved than the case in which pq is a (≤5)-edge. The first step towards the investigation of (>5)-edges is given in Section 4.2, where we prove that {X,Y, Z} is a 3-decomposition of P . This allows us to classify each edge of P as either monochromatic or bichromatic, as we explained at the end of Section 2. Also B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 359 in Section 4.2 we show that for each k ∈ {6, 7, 8} it is easy to determine the number of bichromatic k-edges and the number of monochromatic k-edges. After proving these elementary properties of P we move on to Section 4.3. This is the most technical and long part of the paper, and its purpose is to establish a collection of structural properties of P . On a first read it may be advisable to skip this section, and only come back to it whenever its main results are invoked in Sections 4.4 and 4.5. Finally, in Section 4.4 (respectively, Section 4.5) we prove Lemma 3.1 for the case in which pq is a monochromatic (respectively, bichromatic) 6-edge, 7-edge, or 8-edge. As we shall see, using the structural results from Section 4.3 these tasks are reduced to a relatively straightforward case analysis. For completeness, the conclusion of the proof is presented in Section 4.6. 4 Proof of Theorem 1.2 We recall that throughout this paper, P is an 18-point set in the plane in general position such that cr(P ) = cr(K18). 4.1 The algorithm to label the 18 points in P , and proof of Lemma 3.1 when pq is a (≤ 5)-edge It follows from Proposition 2.1 that the convex hull of P has exactly 3 vertices. Without loss of generality (the whole set P may be rotated, if necessary) we may assume that all three vertices have distinct x-coordinates. Let x0 denote the vertex with the smallest x- coordinate. As we travel counterclockwise along the convex hull starting from x0, let y0 be the first vertex we find, and let z0 be the other vertex. See Figure 2. Figure 2: The convex hull of P . 360 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Observation 4.1. E0(P ) = {x0y0, y0z0, x0z0}. □ We have already unambiguously determined a labelling for the three convex hull ver- tices of P . It remains to unambiguosly determine a labelling for the remaining 15 points of P . For j ∈ {0, . . . , 5}, let x↷j denote the j-th point in P that we find as we rotate the line y0x0 clockwise around y0 (we consider x0 to be the 0-th point in P hit by the rotating line, so that x↷0 = x0). We define y ↷ j and z ↷ j similarly, using z0y0 and x0z0 as the clockwise rotating lines, around z0 and x0, respectively. See Figure 3(a). In an analogous manner, we let x↶j denote the j-th point in P that we find as we rotate the line z0x0 counterclockwise around z0 (again, we consider x0 to be the 0-th point in P hit by the rotating line, so that x↶0 = x0). We define y ↶ j and z ↶ j similarly, using x0y0 and y0z0 as the counterclockwise rotating lines, around x0 and y0, respectively. See Figure 3(b). Figure 3: (a) As we rotate the line y0x0 clockwise around y0, the third point in P we find is labelled x↷3 . The points x ↷ 1 and x ↷ 2 are also indicated. (b) As we rotate the line z0x0 counterclockwise around z0, the third point in P we find is labelled x↶3 . The points x ↶ 1 and x↶2 are also indicated. By definition x ↷ 0 = x ↶ 0 = x0. Note that in this example x ↷ i = x ↶ i for i = 0, 1, 2, 3. Observation 4.2. For each j ∈ {0, . . . , 5}, y0x↷j , z0x↶j , z0y↷j , x0y↶j , x0z↷j , and y0z↶j are all j-edges. The next statement is our first major result on the structure of P . In particular, it yields a labelling of all the points of P . As it happens, this proposition simultaneously establishes Lemma 3.1 for the case in which pq is a (≤ 5)-edge. Proposition 4.3. Let j ∈ {0, . . . , 5}. Then: (1) For u ∈ {x, y, z}, u↷j and u↶j are the same point, which will be denoted uj; (2) For all nonnegative integers m,n such that m+ n = j, we have that (a) Ej(P ) = {umvn ∣∣m+n = j and uv ∈ {xy, yz, zx}}. Moreover, for such values of m,n, and j the following holds: B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 361 (b) umv+n = {ui ∣∣ i < m} ∪ {vi∣∣ i < n} for any uv ∈ {xy, yz, zx}. Proof. We prove (1) and (2) by induction on j. Since x↷0 = x ↶ 0 = x0, y ↷ 0 = y ↶ 0 = y0, and z↷0 = z ↶ 0 = z0, it follows from Observations 4.1 and 4.2 that (1) and (2) are true for j = 0. Now we let t ∈ {0, 1, 2, 3, 4} be an integer such that (1) and (2) hold for every j such that 0 ≤ j ≤ t (in particular, the points xj , yj , zj are already defined for 0 ≤ j ≤ t). We complete the proof by showing that then (1) and (2) hold for j = t+ 1. Let Xt := {x0, . . . , xt}, Yt := {y0, . . . , yt}, Zt := {z0, . . . , zt}, and Pt := Xt ∪ Yt ∪ Zt. From the definitions involved, it follows that |Pt| = 3(t + 1). First we establish an injection ψ : Pt → Et+1(P ). Consider any point xi ∈ Xt. It follows from the induction hypothesis that xiyt−i is a t-edge, and that xiy+t−i = {xr ∣∣ r < i} ∪ {yr ∣∣ r < t− i}. Let xi be the first point that we find as we rotate the line xiyt−i counterclockwise around xi. It is easy to see that the induction hypothesis implies that the rotating line hits xi with its head, and so xix+i = {xr ∣∣ r < i}∪{yr ∣∣ r < t−i+1}. We define ψ(xi) = xixi. In an analogous manner we define ψ(yi) and ψ(zi) for all yi ∈ Yt and zi ∈ Zt. Since ψ defines a one-to-one relation and |Pt| = 3(t+ 1), it follows that |ψ(Pt)| = 3(t+ 1). Let E′ := {x0z↷t+1, y0x↷t+1, z0y↷t+1}. Observation 4.2 implies that E′ ⊂ Et+1(P ). We note that ψ(x0) = x0y↶t+1, ψ(y0) = y0z ↶ t+1, and ψ(z0) = z0x ↶ t+1. Using these observations and that {x↷t+1, y↷t+1, z↷t+1} ∩ Pt = ∅, it follows that E′ ∩ ψ(Pt) = ∅. On the other hand, from Proposition 2.1 it follows that |Et+1(P )| = 3(t + 2). Thus Et+1(P ) is the disjoint union of ψ(Pt) and E′. By way of contradiction, suppose that x↷t+1 ̸= x↶t+1. Then each point of {x1, . . . , xt} is contained in the interior of the quadrilateral bounded by y0x↷t+1, z0x ↶ t+1, z0x0, and x0y0 (see Figure 4). From the induction hypothesis it follows that x↷t+1x ↶ t+1 /∈ E≤t(P ). This and the fact that x↷t+1x ↶ t+1 /∈ ψ(Pt)∪E′ imply that |x↷t+1x↶t+1−| ≥ t+2. Then the interior of the triangle T bounded by y0x↷t+1, z0x ↶ t+1, and x ↷ t+1x ↶ t+1 is nonempty, and, moreover, it contains every element of Q := x↷t+1x ↶ t+1 − \ {x0, x1, . . . , xt}. Let p be the first point of Q that z0x↶t+1 finds as it is rotated counterclockwise around x ↶ t+1. Then x ↶ t+1p must be a (t + 1)-edge of P . On the other hand, it is immediately seen that x↶t+1p /∈ ψ(Pt) ∪ E′, contradicting that Et+1 = ψ(Pt) ∪ E′. This contradiction shows that x↷t+1 and x ↶ t+1 are the same point. Analogous arguments show that y↷t+1 and y ↶ t+1 are the same point, and that z ↷ t+1 and z ↶ t+1 are the same point. This proves (1) for j = t+ 1. Now we show that (2) holds for j = t + 1. Note that at this point xt+1, yt+1, zt+1 are all well-defined. For each m ∈ {0, 1, . . . , t + 1}, we let Xm := {xi ∣∣ i ≤ m}, Ym := {yi ∣∣ i ≤ m}, and Zm := {zi ∣∣ i ≤ m}. Let m ∈ {0, 1, 2, . . . , t+ 1}. We shall show that (i) xmyt+1−m is in Et+1(P ); and (ii) xmy+t+1−m = Xm−1 ∪ Yt−m. By symmetry, analogous arguments show that: (i’) ymzt+1−m is in Et+1(P ); (ii’) ymz+t+1−m = Ym−1 ∪ Zt−m; (i”) zmxt+1−m is in Et+1(P ); and (ii”) zmx + t+1−m = Zm−1 ∪ Xt−m. Note that these six assertions, together with the fact that |Et+1(P )| = 3(t+ 2), imply (2). Thus we complete the proof by showing (i) and (ii). 362 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Figure 4: If x↷t+1 ̸= x↶t+1, then the triangle T contains a point p such that x↶t+1p is a (t+1)- edge of P . Since xt+1 = x↶t+1 = x ↷ t+1, yt+1 = y ↶ t+1 = y ↷ t+1, and zt+1 = z ↶ t+1 = z ↷ t+1, it follows that (i) and (ii) hold whenever m is in {0, t + 1}. Thus it suffices to prove (i) and (ii) for 1 ≤ m ≤ t. From the induction hypothesis we have that xm−1y+t+1−m = Xm−2 ∪ Yt−m and xmy + t−m = Xm−1 ∪ Yt−m−1. Also note that Xm−1 ∪ Yt−m ⊆ xmy+t+1−m. LetB denote the triangle bounded by the lines xm−1yt+1−m, xmyt−m, and xmyt+1−m (see Figure 5). Let PB be the set of points of P contained in the interior of B. We claim that PB = ∅. By way of contradiction, suppose this is not the case. Let L = p1p2 · · · pk be the lower chain of the convex hull of PB ∪ {xm, yt+1−m}. Then p1 = xm and pk = yt+1−m, where (since B ̸= ∅) k ≥ 3. We note that pip+i+1 = Xm−1 ∪Yt−m for all i = 1, 2, . . . , k− 1. Thus each edge of L is a (t+1)-edge. We recall that Et+1(P ) = ψ(Pt)∪E′. It is readily seen that no edge of L is in E′, and so every edge of L is in ψ(Pt). In particular, the line p2p3 is in ψ(Pt). Recall that every edge in ψ(Pt) is obtained by starting with a line viwt−i (for v, w ∈ {x, y, z}, v ̸= w), counterclockwise rotating it around vi, and recording the first point p in P hit by the rotating line: ψ(vi) is then the line vip. Thus, in particular p2p3 is obtained in this way. Now if we reverse the process and clockwise rotate p2p3 around p2, the first point hit by the rotating line must be yt−m. This implies that p2 = xm, contradicting that p1 = xm. We therefore conclude that PB = ∅. Finally, note that PB = ∅ immediately implies that ψ(xm) = xmyt+1−m. Thus xmyt+1−m is a (t + 1)- edge. This proves (i). Moreover, as we observed above, Xm−1 ∪ Yt−m ⊆ xmy+t+1−m. Since |Xm−1 ∪ Yt−m| = t+ 1, then Xm−1 ∪ Yt−m = xmy+t+1−m. Thus (ii) follows. In view of Proposition 4.3(1), we have achieved our goal to unambiguously identify B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 363 Figure 5: If the interior of the triangle B is nonempty, then every edge of the convex chain p1, p2, . . . , pk is a (t+ 1)-edge. 364 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 (and label) all 18 points of P . For the rest of the paper, we letX := {x0, x1, . . . , x5}, Y := {y0, y1, . . . , y5}, and Z := {z0, z1, . . . , z5}, where xj , yj , and zj are as in Proposition 4.3, for j = 0, 1, . . . , 5. 4.2 {X,Y, Z} is a 3-decomposition of P If we rotate the line x0z0 clockwise along x0, then for j = 1, 2, . . . , 5, the j-th point hit by the rotating line is zj . If we rotate the line x0y0 counterclockwise along x0, then for j = 1, 2, . . . , 5, the j-th point hit by the rotating line is yj . It follows that the sixth point hit by the clockwise rotating line ℓx is in X , and the sixth point hit by the counterclockwise rotation line ℓ′x is also in X (see Figure 6). These two points in X are obviously distinct (since |X| > 2), and so they define an infinite cone CX with vertex x0 (here by cone with vertex p we mean a pair of distinct directed rays, both with startpoint p). Note that CX is the smallest infinite cone with vertex x0 that contains X . See Figure 6. We similarly find infinite cones CY (with vertex y0) and CZ (with vertex z0). Figure 6: The sets X,Y, and Z are contained in the indicated (closed) shaded regions. The shaded region containing X is ∆X,Y ∩∆X,Z . Now CX ∪ CY divide the plane into several regions, three of which are bounded. Two of these bounded regions are triangles: one triangle ∆X,Y with x0 as a vertex and another triangle ∆Y,X with y0 as a vertex; the other one is a quadrilateral. The entire set X is contained in ∆X,Y , and the entire set Y is contained in ∆Y,X . By considering the pair CX , CZ (respectively, CY , CZ), we obtain triangles ∆X,Z and ∆Z,X (respectively, ∆Y,Z and ∆Z,Y ). Thus X ⊆ ∆X,Y ∩ ∆X,Z , Y ⊆ ∆Y,X ∩ ∆Y,Z , and Z ⊆ ∆Z,X ∩ ∆Z,Y . Hence the situation is as illustrated in Figure 6. In this figure, each of ∆X,Y ∩ ∆X,Z , ∆Y,X ∩∆Y,Z , and ∆Z,X ∩∆Z,Y is a quadrilateral, although it is easy to see that any (or all) of them may be a triangle. In view of this, it follows immediately that there is a triangle that witnesses the follow- ing: Proposition 4.4. P is 3-decomposable, with 3-decomposition {X,Y, Z}. □ As we mentioned in Section 2, knowing that {X,Y, Z} is a 3-decomposition of P B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 365 allows us to classify each edge of P as either monochromatic or bichromatic: for p, q ∈ P , the edge pq is monochromatic if p and q belong to the same set of the 3-decomposition {X,Y, Z}. Otherwise, pq is bichromatic. We close this section by noting that using the 3-decomposability of P it is easy to determine the number of bichromatic and monochromatic k-edges in P , for each k ∈ {0, . . . , 8}. Indeed, since P is 3-decomposable it follows from [2, Claim 1] that ebi≤k(P ) = 3 ( k+2 2 ) for k ∈ {0, . . . , 5}, ebi≤6(P ) = 81, and ebi≤7(P ) = 99. Also note that ebi≤8(P ) is the total number of bichromatic edges of P , namely 3 · 6 · 6 = 108. Using that ebij(P ) = ebi≤j(P )− ebi≤j−1(P ) for j ∈ {1, . . . , 8}, we obtain the following. Proposition 4.5. ebik (P ) = 3(k + 1) for k ∈ {0, . . . , 5}, ebi6 (P ) = 18, ebi7 (P ) = 18, and ebi8 (P ) = 9. To obtain emon6 (P ), e mon 7 (P ) and e mon 8 (P ), we note that Proposition 2.1 implies that e6(P ) = 24, e7(P ) = 33, and e8(P ) = 33. Since ej(P ) = ebij (P ) + e mon j (P ) for j = 0, . . . , 8, Proposition 4.5 implies the following. Corollary 4.6. emonk (P ) = 0 for k ∈ {0, . . . , 5}, emon6 (P ) = 6, emon7 (P ) = 15, and emon8 (P ) = 24. 4.3 Structural properties of P 4.3.1 Determination of euuk (P ) for any u ∈ {x, y, z} and any k ∈ {0, . . . , 8} Let u ∈ {x, y, z}, i ∈ {1, 2, . . . , 5}, and ℓu, ℓ′u be the directed rays forming the cone CU with vertex u0 mentioned in the arguments leading to Proposition 4.4. See Figure 6. From now on, we shall use ui0 to denote the i-th point of P that ℓu finds when it is rotated clockwise around of u0 until it reaches ℓ′u. Clearly, {u10, u20, . . . , u50} = {u1, u2, . . . , u5}. Our next observation is evident, but useful. Observation 4.7. Let v1, v2, and v3 be three distinct points in P , and let ℓ := v1v2 and ℓ′ := v1v3. Let P1 := ℓ− ∩ ℓ′+, P2 := ℓ+ ∩ ℓ′+, and P3 := ℓ− ∩ ℓ′−. Then P1, P2, and P3 are pairwise disjoint subsets of P . See Figure 7. For i = 1, 2, 3, let ri be the number of points in Pi. If P \ {v1, v2, v3} is the disjoint union of P1, P2, and P3, and pi is the i-th point of P1 that ℓ finds when it is rotated counterclockwise around v1 until it reaches ℓ′, then v1pi is a j-edge of P for j = min{r2 + i, 16− (r2 + i)}. The next observation is immediate from the definition of ui0 and Observation 4.7. Observation 4.8. Let u ∈ {x, y, z}. Then (1) u0u10 and u0u 5 0 are both 6-edges, (2) u0u20 and u0u 4 0 are 7-edges and they are the only 7-edges of the type u0u, and (3) u0u30 is an 8-edge. Claim 4 in [2] implies that euu8 (P ) ≤ 8 for each u ∈ {x, y, z}. Using this, together with Observation 4.8 and Corollary 4.6, we obtain the following. Proposition 4.9. Let u ∈ {x, y, z}. Then euuk (P ) = 0 for k ∈ {0, . . . , 5}, euu6 (P ) = 2, euu7 (P ) = 5, and e uu 8 (P ) = 8. 366 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Figure 7: The i-th point of P1 := ℓ− ∩ ℓ′+ that ℓ finds when it is rotated counterclockwise around v1 is a j-edge for j = min{r2 + i, 16− (r2 + i)}, where r2 denotes the number of points of P2 := ℓ+ ∩ ℓ′+. The next corollary follows immediately from Observation 4.8(1) and Proposition 4.9. Corollary 4.10. Emon6 (P ) = {x0x10, x0x50, y0y10 , y0y50 , z0z10 , z0z50}. Moreover, any other monochromatic edge must belong to Emon7 (P ) ∪ Emon8 (P ). 4.3.2 Determination of the convex hull of U for U ∈ {X,Y, Z} and related facts Let u be any element of {x, y, z}. One of the main goals in this subsection is to show that the triangle formed by u0, u4 and u5 contains in its interior the remaining u’s, namely, u1, u2 and u3. We also prove other statements about the relative position of the elements of U . Almost all these assertions will be used in the subsequent steps later on. Proposition 4.11. Let u ∈ {x, y, z}. If u1 ∈ {u20, u40}, then there are at least three 7-edges of type uu involving u1 but not u0. Proof. We prove the proposition for the case u = x. The cases u = y and u = z are handled in a totally analogous manner. Suppose that x1 = x40. Then ℓ := x0x1 leaves x 1 0, x 2 0 and x 3 0 on its left halfplane (and x50 on its right halfplane). Let z ′ be the first z that ℓ finds when it is rotated counter- clockwise around x1, and let ℓ′ = x1z′. See Figure 8. By Corollary 4.10, we know that {x1x10, x1x20, x1x30} ⊂ Emon7 (P ) ∪ Emon8 (P ). Then ℓ finds each of x10, x20 and x30 before it reaches z′. This and Observation 4.7 imply that at most one of x1x10, x1x 2 0, x1x 3 0 is an 8-edge and, by Corollary 4.10, the other two must be 7-edges. From the way in that the x′s were labelled it follows that x50 is the first x that ℓ finds when it is rotated clockwise around x1, and so x1x50 is a (≤ 7)-edge. This and Corol- lary 4.10 imply that x1x50 is the third required 7-edge. The case x1 = x 2 0 can be handled in an analogous manner (y′s play the role of z’s). Proposition 4.12. Let u ∈ {x, y, z} and {p, q} = {x, y, z} \ {u}. Suppose that {q0, . . . , q5} ⊂ u0u3−0 and that {p0, . . . , p5} ⊂ u0u 3+ 0 . Then: (A1) u1 /∈ {u10, u50}; B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 367 Figure 8: Here x0x1 is a 7-edge. (A2) there are at least two 7-edges of type uu involving u1 but not u0; (A3) u2 /∈ {u10, u50}; (A4) each of u3u4, u3u5 and u4u5 is an 8-edge; (A5) {u10, u50} = {u4, u5}; (A6) the triangle formed by u0, u4 and u5 is the convex hull of U ; and (A7) if u5 ∈ u0u+4 , then u0u − 4 = {q0, . . . , q5} and u0u + 5 = {p0, . . . , p5}. Otherwise, u0u + 4 = {p0, . . . , p5} and u0u − 5 = {q0, . . . , q5} . Proof. By rotating P if necessary, and exchanging appropriately the labels x, y, and z, we can assume, without any loss of generality, that u = x, p = y, q = z and that X,Y and Z are placed as in Figure 9. (A1): Seeking a contradiction, suppose that x10 = x1. Let v be the first point that x0x1 finds when it is rotated clockwise around x1 as shown in Figure 9(a). Note that v ∈ Y , as otherwise v ∈ {x2, x3, x4, x5} and x1v− = Z. Then x1v is a 6-edge, contradicting Corollary 4.10. Let x′ be the last element of {x2, x3, x4, x5} that x1v finds when it is rotated clockwise around x1. Since v ∈ Y , then x1x′ must be a (≤ 6)-edge, contradicting Proposition 4.9. The case x50 = x1 can be handled in an analogous manner (with the roles of Z and Y interchanged). (A2): From (A1) we know that x0x1 leaves at least one x in each side. By definition of x1, the points x2, x3, x4 and x5 must be contained in X ′ := X ∩ x1z+0 ∩ x1y − 0 , see Figure 9(b). Let x′ be the last element of {x2, x3, x4, x5} that x0x1 finds when it is rotated clockwise around x1 as shown in Figure 9(b). Note that x1x′ must be a (≤ 7)-edge. Since x′ ̸= x0, then Proposition 4.9 implies that x1x′ must be a 7-edge. Similarly, if we rotate x0x1 in the other direction, then we can find the other 7-edge involving x1 but not x0. (A3): Seeking a contradiction, suppose that x2 = x10. Then x0x2 is a 6-edge, and x3, x4 and x5 are contained in X ′′ := X ∩x10y−0 . See Figure 10(b). From Observation 4.7, we know that at most one of x2x3, x2x4, x2x5 is an 8-edge. This and Corollary 4.10 imply 368 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Figure 9: (a) x0x1 cannot be a 6-edge. (b) There are at least two 7-edges of type xx involving x1 but not x0. that at least two of x2x3, x2x4, x2x5 are 7-edges. This together with Observation 4.8(2) and (A2) imply that exx7 (P ) ≥ 6, which contradicts Proposition 4.9. The case x2 = x50 can be handled in an analogous manner. (A4): From Corollary 4.10, Observation 4.8(1), and (A1), we know that x0x1 is a 7- edge or an 8-edge. First suppose that x0x1 is a 7-edge. Then x1 ∈ {x20, x40}. This together with Propositions 4.9 and 4.11 and Observation 4.8(2) imply that each element of Exx7 (P ) contains at least one of x0 or x1. This fact and Proposition 4.9 imply that x3x4, x3x5 and x4x5 are 8-edges, as required. Now, we suppose that x0x1 is an 8-edge. Then x2 ∈ x0x−1 or x2 ∈ x0x + 1 . We only analyze the case x2 ∈ x0x−1 (the other case is symmetric). Then we must have that X ′ := X ∩ x0x+1 ∩ x2y − 0 contains exactly two elements x ′, x′′ of {x3, x4, x5}, see Figure 10(a). Now we rotate x2y0 clockwise around x2 until it be parallel to x0x1. See Figure 10(a). From Observation 4.7 and Corollary 4.10, we know that at least one of x2x ′, x2x ′′ is a 7-edge. Such a 7-edge plus the four 7-edges provided by Observation 4.8(2) and (A2) give us, 5, the total number of 7-edges of P . This and Proposition 4.9 imply that x3x4, x3x5, x4x5 are 8-edges, as required. (A5): Seeking a contradiction, suppose that {x10, x50} ̸= {x4, x5}. Then (A1) and (A3) imply that x3 = x10 or x3 = x 5 0. Again, by symmetry it is enough to analyze the case x3 = x 1 0. Clearly, both x1 and x2 are contained in the triangle formed by x0x3, x0x 5 0 and x3y0; and x4, x5 are contained in X ′′ := X ∩ x10y−0 , see Figure 10(b). Now we rotate x0x3 clockwise around x3 until it reaches x3y0, and note that such a rotation hits x4 and x5. From Observation 4.7 we know that at most one of x3x4 or x3x5 is 8-edge, contradicting (A4). (A6): This follow directly from (A5) and the way in that the x’s were labelled. (A7): Suppose that x5 ∈ x0x+4 . Then (A5) implies that x10 = x4 and x50 = x5. The required equalities follow from the definition of x10 and x 5 0 and the hypotheses Z ⊂ x0x3−0 and Y ⊂ x0x3+0 . Similarly, we can deduce that x0x − 5 = Z and x0x + 4 = Y whenever x5 ∈ x0x−4 . B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 369 Figure 10: (a) x3x4, x3x5, x4x5 are 8-edges. The dotted straight line containing x2 is parellel to x0x1. (b) {x10, x50} = {x4, x5}, and hence x0x4 and x0x5 are 6-edges. 4.3.3 Determination of the position of u5 with respect to u0u4 for each u ∈{x, y, z} Our main goal in this subsection is to show that u5 ∈ u0u+4 for each u ∈ {x, y, z}. In order to prove this, we need to establish some auxiliary statements that will also be used later on. Let u, v ∈ {x, y, z} with u ̸= v. We will say that u4u5 splits the v’s to mean that u4u5 separates {u0, . . . , u3} ∪ {v0, . . . , v3} from the rest of the points of P \ {u4, u5}. Proposition 4.13. Let {u, v, w} = {x, y, z}, and suppose that u0u5 separates u4 from the v’s. Then u4u5 splits the v’s. Proof. By rotating and/or reflecting P along u0u4, if necessary, we can assume that u = x, v = y, w = z and that X,Y and Z are placed as in Figure 6. Since x0x5 separates x4 from the y’s, then x5 ∈ x0x+4 . From Proposition 4.3 we know that x4y+0 = {x0, x1, x2, x3}. Then (A6) implies that x4y + 0 ⊆ x4x + 5 . If we rotate x4y0 counterclockwise around x4 until it reaches x0x4, then x5 ∈ x0x+4 , (A6), and Observa- tion 4.7 together imply that at most one element of {x4x5} ∪ {x4y ∣∣y ∈ Y } is an 8-edge. This and (A4) imply that such an 8-edge must be x4x5. Then (A6) implies that x4x5 leaves exactly four y’s on its right. Moreover, from Proposition 4.3 it is easy to see that y0 and y1 are in x4x+5 . Let yi and yj be two elements of Y in x4x − 5 . Without loss of generality, we can assume that i < j. Then 2 ≤ i < j ≤ 5. From (A6) we know that the triangle formed by y0, y4, and y5 is the convex hull of Y . This implies that j ∈ {4, 5}. Seeking a contradiction, suppose that i ∈ {2, 3}. Let T be the triangle formed by x0x5, x0yi and x4x5. See Figure 11(b). By the way y’s were labelled, we know that if yr ∈ T then r ∈ {i+1, . . . , 5}\{j}. Let y′ be the first point in Y ∩ T that yix0 finds when it is rotated clockwise around yi. See Figure 11(b). Then yiy ′ is a (≤ i + 4)-edge because the points of P lying in the left side of yiy′ is a subset of {x0, . . . , x3, y0, . . . , yi−1}. If i = 2, then yiy′ is a (≤ 6)-edge which does not involve y0, contradicting Corollary 4.10. Finally, if i = 3, then yiy′ is a (≤ 7)-edge, contradicting (A4). Observation 4.14. From Proposition 4.13 we know that if {u, v, w} = {x, y, z}, then u4u5 splits the v’s or the w’s. 370 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Figure 11: (a) If x4 ∈ x0x−5 , then x4x5 splits Y . (b) There are exactly two y’s, namely yi and yj , in x4x−5 . Proposition 4.15. Let u and v be two distinct elements of {x, y, z}. If u4u5 splits the v’s, and v3v5 leaves u0 and v2 on the same side, then v4v5 splits the u’s. Proof. By rotating P if necessary and exchanging appropriately the labels x, y and z, we can assume that u = x and that X,Y and Z are placed as in Figure 6. CASE 1: Suppose that x4x5 splits the y’s. Then we need to show that if y3y5 leaves x0 and y2 on the same side, then y4y5 splits the x’s. From (A4), we know that y4y5 is an 8-edge, and from (A6), that y4y5 is in the convex hull of Y . First, we show that y5 ∈ y0y−4 . By way of contradiction, suppose this is not the case. Then y5 ∈ y0y+4 and the triangle formed by y0, y4 and y5 looks like in Figure 12(a). Since x4x5 splits the y’s, then x4x5 separates y3 from y4 and y5, and so all the x’s are on the left side of both y3y4 and y3y5. In particular, x0 ∈ y3y−5 , and hence y2 ∈ y3y − 5 . Then y2 ∈ y3y−5 ∩ z0y − 3 , and so y2 is contained in the triangle Q formed by y0y4, z0y3, y3y5. See Figure 12(b). Since y3y4 is an 8-edge by (A4), and y3y4 leaves {y0, y2, x0, . . . , x5} on its left, then it leaves y1, y5 on its right. This and the fact that y1 ∈ z0y−3 imply that y1 is contained in the triangle R formed by z0y3, y3y4, y0y5. See Figure 12(b). Then y2y4 and y0y1 are 7-edges. This, together with Observation 4.8(2) and Proposition 4.11, implies eyy7 (P ) ≥ 6, the required contradiction. Thus, we can conclude that y0y4 leaves the x’s and y5 on the same side, and the desired result follows from Proposition 4.13. CASE 2: x4x5 splits the z’s. Follow the same argument as in CASE 1 with left, right, y, z,− and + in place of right, left, z, y,+ and −, respectively. B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 371 Figure 12: (a) y5 ∈ y0y+4 . (b) y2 is in Q and y1 is in R. Proposition 4.16. If u ∈ {x, y, z}, then u3u5 leaves u0 and u1 on the same side. Proof. As in Proposition 4.15, we can assume that u = x and thatX,Y andZ are placed as in Figure 6. From (A4), we know that x3x5 is an 8-edge. Seeking a contradiction, suppose that x3x5 separates x0 from x1. First, suppose that x4x5 splits the y’s. Since P is placed as in Figure 6, then x0 ∈ x3x+5 , and hence, x1 ∈ x3x−5 . Then x3x + 5 = {x0, x2}∪Y , or equivalently, x3x − 5 = {x1, x4}∪Z. This fact has two immediate consequences. The first one is that x1 = x20. This fact and Proposition 4.11 imply that there are at least three 7-edges of type xx involving x1 but not x0. The second consequence is that x2 is in the triangle X ′ (see Figure 13) formed by x1y0, x3x5 and x0x5, and hence x2x5 must be a 7-edge too. The existence of these four 7-edges together with those in Observation 4.8(2) imply exx7 (P ) ≥ 6, contradicting Proposition 4.9. Now suppose that x4x5 splits the z’s. Again, since P is placed as in Figure 6, then x0 ∈ x3x − 5 and x1 ∈ x3x + 5 . By similar arguments as above, we can deduce that x1 = x 4 0 and that x2x5 is a 7-edge. As before, x1 = x40 and Proposition 4.11 imply that there are at least three 7-edges of type xx involving x1 but not x0. The existence of these four 7-edges together with those in Observation 4.8(2) imply exx7 (P ) ≥ 6, contradicting Proposition 4.9. Proposition 4.17. There is a u ∈ {x, y, z} such that u3u5 separates u4 from the other u’s. Proof. From Proposition 4.16 we know that u3u5 leaves u0 and u1 on the same side for each u ∈ {x, y, z}. Seeking a contradiction, we suppose that u3u5 separates {u0, u1} from {u2, u4} for each u ∈ {x, y, z}. By rotating and/or reflecting P if necessary, and exchanging appropriately the labels x, y and z, we can assume that X,Y and Z are placed as in Figure 11(a), and that x4x5 splits the y’s. An immediate consequence of these assumptions and our hypothesis is that (i) x3x5 leaves to z0 and x2 on its left side. 372 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 Figure 13: Here x1 ∈ x3x−5 . Now suppose that y5 ∈ y0y+4 . Then y0y5 separates y4 from the z’s, and from Proposi- tion 4.13 we know that y4y5 splits the z’s. In particular, the triangle formed by y0, y4 and y5 must be as in Figure 12(a) and y0 ∈ y3y+5 . By supposition, y3y5 separates y0 from y2, and so y3y5 leaves to x0 and y2 on its left side. This last and Proposition 4.15 imply that y4y5 splits the x’s too, which is impossible. Thus we conclude that y5 ∈ y0y−4 . This and Proposition 4.13 imply that y4y5 splits the x’s. This fact and our supposition imply that (ii) y3y5 leaves to z0 and y2 on its right side. If z4z5 splits the x’s, then (i) and Proposition 4.15 imply that x4x5 splits the z’s, which contradicts that x4x5 splits the y’s. Similarly, if z4z5 splits the y’s, then (ii) and Proposi- tion 4.15 imply that y4y5 splits the z’s, again contradicting that y4y5 splits the x’s. Proposition 4.18. Let u be an element in {x, y, z} that satisfies the property in Propo- sition 4.17, and suppose that u4u5 splits the v’s, where v ∈ {x, y, z} \ {u}. Then the following hold: (B1) u3u5 separates v5 from the other v’s; (B2) v4v5 splits the w’s, where {w} = {x, y, z} \ {u, v}; (B3) v1v4 and v2v4 are both 7-edges; and (B4) v3v5 separates v4 from the other v’s. Proof. Without loss of generality, we can assume that u = x, v = y, and X,Y and Z are placed according to Figure 11. Indeed, we can get such requirements by rotating and/or reflecting P , and by exchanging appropriately the labels x, y and z. (B1): From our assumptions and the hypothesis we know that x4 is the only x in x3x−5 . Since x3x5 is an 8-edge, then exactly one element y∗ of Y is in x3x−5 . From (A6), we know that such a y∗ is one of y4 or y5. If y∗ = y4, then, by the way y0, . . . , y5 were labelled, we have that y5 must be contained in the triangle S of Figure 14. Then y4y5 leaves x3, x4, x5 and all the z’s on its right side. This implies that y4y5 cannot be an 8-edge, contradicting (A4). This contradiction implies that (B1) holds. B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 373 (B2): From (B1), we know that y∗ = y5 is the only element of Y in x3x−5 . If y5 ∈ y0y − 4 , then y4 must be contained in the region R of Figure 14. This implies that each element in (X ∪ Y ) \ {x4, y4, y5} lies in y4y−5 , contradicting (A4) that y4y∗ = y4y5 is an 8-edge. Thus we have that y5 ∈ y0y+4 . This fact and Proposition 4.13 imply that y4y5 splits the z’s, as required. In view of (B2) and our previous assumptions, for the rest of the proof, we may assume that the two triangles defined by {x0, x4, x5} and {y0, y4, y5} are as shown in Figure 12(a). (B3): Let ℓ be the line through y4 which is parallel to x4x5 and let M be the interior of the triangle formed by y0y5, y0y4 and x4x5. See Figure 12(a). Since y4 and y5 are the only y’s in x4x−5 , then M ∩ Y = {y1, y2, y3}. If we rotate ℓ in clockwise order around y4 until it reaches y4y0, then by Observation 4.7 we have that exactly one of {y1y4, y2y4, y3y4} is 8-edge and the other two are 7-edges. The desired assertion follows from (A4). (B4): Seeking a contradiction, suppose that y3y5 does not separate y4 from the other y’s. Then Proposition 4.16 implies that y3y5 separates {y0, y1} from {y2, y4}. On the other hand, since y3y4 is an 8-edge and x4x5 separates y3 from y4, then y3y−4 = X ∪ {y0, y2}. Thus y1 must be contained in the triangle R formed by z0y3, y3y4, y0y5. See Figure 12(b). This implies that y40 = y1. Then Proposition 4.11, Observation 4.8(2) and (B3) imply, eyy7 (P ) ≥ 6, a contradiction. Figure 14: Here x3x5 separates y∗ from the other y’s. Remark 4.19. From now on, without loss of generality, we assume that the u ∈ {x, y, z} satisfying Proposition 4.17 is x, and that x5 ∈ x0x+4 . Indeed, it is not hard to see that we can get such requirements by rotating and/or reflecting P along x0x4, and by appropriately exchanging the labels x, y and z. In particular, we assume that X,Y, Z, x0, x4 and x5 are placed as in Figure 11. Note that (B4) appears as hypothesis in Propositions 4.17 and 4.18. The following corollary is an immediate consequence of this fact. Corollary 4.20. Let σ(x) = y, σ(y) = z and σ(z) = x. The following hold for each u ∈ {x, y, z}: (C1) u3u5 separates σ(u)5 from the other σ(u)’s; (C2) σ(u)4σ(u)5 splits the σ(σ(u))’s; 374 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 (C3) σ(u)1σ(u)4 and σ(u)2σ(u)4 are both 7-edges; (C4) σ(u)3σ(u)5 separates σ(u)4 from the other σ(u)’s; and (C5) u5 ∈ u0u+4 . Proof. In view of Remark 4.19, we may assume that x4x5 splits the y’s, x3x5 separates x4 from the other x’s, and X,Y, Z, x0, x4 and x5 are placed as in Figure 11. First, we show that (C1) – (C4) hold. Proposition 4.18 states exactly (C1) – (C4) for u = x and v = σ(x) = y. In particular, (C2) and (C4) tell us that y4y5 splits the z’s and that y3y5 separates y4 from the other y’s, respectively. By applying Proposition 4.18 to the last two conclusions on y’s we have that (C1) – (C4) also hold for u = y and v = σ(y) = z. Similarly, we can conclude that (C1) – (C4) also hold for u = z and v = σ(z) = x. Now, we show (C5). For u = x the assertion holds by Remark 4.19. We first analyze the case u = y. From (A6) and Remark 4.19, we know that {x0, . . . , x5} lies on the left side of both y0y4 and y0y5. Seeking a contradiction, suppose that y5 ∈ y0y−4 . Then, from (A6) and the last two facts, it is easy to verify that x0 ∈ y3y−5 . Similarly, from y5 ∈ y0y − 4 and (C4), we can deduce that y2 ∈ y3y−5 . Thus x0, y2 ∈ y3y − 5 , and so Proposition 4.15 implies that y4y5 splits the x’s, contradicting (C2). An analogous argument shows that (C5) also holds for u = z. From Remark 4.19 and Corollary 4.20, we have that the points of P with indices 0, 3, 4 and 5 are placed as in Figure 15. Figure 15: The relative position of the points of P with indices 0, 3, 4 and 5. B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 375 4.4 Determination of pq+ when pq is a monochromatic edge. Lemma 3.1 for the case in which pq is a monochromatic edge will follow from Proposi- tions 4.21 and 4.22 below. Regarding the statements of these propositions, we recall from Remark 4.19 and Corollary 4.20 that x4x5 splits the ys, y4y5 splits the zs, and z4z5 splits the xs. Proposition 4.21. Let u ∈ {x, y, z}, and let v be the element in {x, y, z} \ {u} such that u4u5 splits the vs. Then the following hold: (D1) u0u+5 = {v0, . . . , v5}; (D2) u0u+4 = {u1, u2, u3, u5} ∪ {v0, . . . , v5}; (D3) u4u+5 = {u0, u1, u2, u3} ∪ {v0, v1, v2, v3}; and (D4) u3u+5 = {u0, u1, u2} ∪ {v0, v1, v2, v3, v4}. Proof. (D1): From (A7), we know that u0u5 separates the v’s from the w’s. Moreover, because u0u5 is an edge of the convex hull of U , then u0u5 leaves the other u’s on the same side. This and (C5) imply that {u1, u2, u3, u4} ⊂ u0u−5 . Again, from (C5) and Proposition 4.13 we have that {v0, v1, v2, v3} ⊂ u4u+5 , and hence {v0, v1, v2, v3} ⊂ u0u + 5 . This and the fact that u0u5 separates the v’s from the w’s imply that {v0, . . . , v5} ⊆ u0u+5 . We finally note that Observation 4.8(1) implies that u0u+5 = {v0, . . . , v5}, as required. (D2): From (C5), (D1), and the way in that the points of P were labelled we have that {v0, . . . , v5} = u0u+5 ⊂ u0u + 4 . On the other hand, since u0u4 is an edge of the convex hull of U , then u0u4 leaves the other u’s on the same side. This and (C5) imply that {u1, u2, u3, u5} ⊂ u0u+4 . Thus {u1, u2, u3, u5} ∪ {v0, . . . , v5} ⊂ u0u + 4 . Again, Observation 4.8(1) implies that u0u+4 = {u1, u2, u3, u5} ∪ {v0, . . . , v5}, as required. (D3): It follows immediately from (C5) and Proposition 4.13. (D4): Clearly, {v0, . . . , v5} ∩ u4u+5 ⊂ {v0, . . . , v5} ∩ u3u + 5 . This fact, together with (C1) and (D3), implies that {v0, v1, v2, v3, v4} ⊂ u3u+5 . On the other hand, from (C4) is easy to verify that {u0, . . . , u5} ∩ u3u+5 = {u0, u1, u2}. Then {u0, u1, u2} ∪ {v0, v1, v2, v3, v4} ⊂ u3u+5 . We finally note that (A4) implies that u3u + 5 = {u0, u1, u2} ∪ {v0, v1, v2, v3, v4}, as required. Proposition 4.22. Let u ∈ {x, y, z}, and let v be the element in {x, y, z} \ {u} such that u4u5 splits the vs. Then the following hold: (E1) u0u1 is an 8-edge; (E2) u0u2 and u0u3 are 7-edges; (E3) u2u3 and u2u5 are 8-edges; (E4) u2u+4 = {u5} ∪ {v0, . . . , v5}; (E5) u1u+4 = {u2, u3, u5} ∪ {v0, . . . , v5} and u3u + 4 = {u2, u5} ∪ {v0, . . . , v5}; (E6) u0u+2 = {u5} ∪ {v0, . . . , v5}; (E7) u0u+1 = {u2, u5} ∪ {v0, . . . , v5} and u0u + 3 = {u1, u2, u5} ∪ {v0, . . . , v5}; 376 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 (E8) u1u+5 = {u0} ∪ {v0, . . . , v5}; (E9) u1u2 and u1u3 are 8-edges; (E10) u1u+3 = {u2, u5} ∪ {v0, . . . , v5}; (E11) u1u+2 = {u0, u5} ∪ {v0, . . . , v5}; (E12) u2u+5 = {u0, u1} ∪ {v0, . . . , v5}; and (E13) u2u+3 = {u4, u5} ∪ {v0, . . . , v5}. Proof. (E1): From Proposition 4.9, Corollary 4.10, and (A5), we know that x0x1 is a 7- or an 8-edge. If x0x1 is a 7-edge, then x1 ∈ {x20, x40} and by Proposition 4.11, there are at least three 7-edges of type xx involving x1 but not x0. This, Observation 4.8(2), and (C3) imply that exx7 (P ) ≥ 6, a contradiction. Thus x0x1 must be an 8-edge. (E2): From Observation 4.8(2), we know that there are exactly two 7-edges of the type x0x. Since (E1) and (A6) imply that none of x0x1, x0x4, x0x5 is a 7-edge, then both x0x2 and x0x3 are 7-edges, as desired. (E3): By Corollary 4.10 and (A5), each of x2x3 and x2x5 is a 7- or an 8-edge. From (C3) and (E2), we know that each of x1x4, x2x4, x0x2, x0x3 is a 7-edge. Since (A2) guar- antees the existence of an additional 7-edge involving x1 and exx7 (P ) = 5, then x2x3 and x2x5 are 8-edges, as required. Observation 4.23. Since z4z5 separates {x4, x5} from the other x’s (see Figure 15), then xix4 leaves Z on its left for any i ∈ {0, 1, 2, 3}. (E4): From (C3), we know that x1x4 and x2x4 are 7-edges. This and Observation 4.23 imply that x2x4 leaves exactly 1 or exactly 3 points of X on its left. Suppose first that x1 ∈ x2x+4 . Since x0 ∈ x2x − 4 , then x2x + 4 = {x1, x3, x5} ∪ Y . Again, Observation 4.23 implies that when we rotate x2x4 counterclockwise around x4, the first two points that such line finds (with the tail) are x1 and x3. Since x1x4 is a 7-edge and x3x4 is an 8-edge, then the first point that such a rotation finds must be x3, and hence x1 ∈ x3x+4 . This and the way in which the x′s were labelled imply that x40 = x1. But then x0x1 is a 7-edge, contradicting (E1). Then x0, x1 ∈ x2x−4 and hence x2x4 leaves exactly three points of X on its left, namely x0, x1 and x3. The desired equality follows from the last conclusion, Observation 4.23, and (C3). (E5): From (A4), we know that x3x4 is an 8-edge, and from (C3) that x1x4 is a 7-edge. Since x1, x3 ∈ x2x−4 , by (E4), then when we rotate x2x4 clockwise around x4, the first two points that such line finds (with the tail) are precisely x1 and x3. Since x1x4 is a 7-edge and x3x4 is an 8-edge, then such a rotation finds first x3 and then x1. The desired conclusions are immediate from this fact and (E4). (E6): From (E4) and the way the x’s were labelled, we know that when we rotate x2x4 clockwise around x2, the first point that such line finds (with the tail) is one of x0 or x1. Since x0x2 is a 7-edge, by (E2), then such a point must be x0 and the desired result follows from (E4). (E7): From (E6), we know that the first two points that we find when we rotate x0x2 counterclockwise around x0 are x1 and x3. Since x0x1 is an 8-edge, by (E1), then we B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 377 have that such a rotation finds x1 and then x3. These together with (E6) imply the desired results. (E8): (D4) implies that x3 ∈ x1x−5 . If x2 ∈ x1x + 5 , then {x1, x3, x4} ∪ Z ⊂ x2x − 5 . This would imply that x2x5 is not an 8-edge, contradicting (E3). Then we can assume that x2 ∈ x1x−5 . Thus, x0 is the only x on the right side of x1x5 and hence x1x5 is a (≤ 7)- edge. From Proposition 4.9 and Corollary 4.10, we have that x1x5 must be a 7-edge. This implies that Y ⊂ x1x+5 , and hence (E8) holds. (E9): (E4), (E5), (E6), (E7), and (E8) imply, respectively, that x2x4, x1x4, x0x2, x0x3 and x1x5 are 7-edges. Then Proposition 4.9 implies that these five are all the monochro- matic edges of type xx. This and Corollary 4.10 imply that x1x2 and x1x3 must be 8-edges. (E10): The first assertion of (E7) implies that x3, x4 ∈ x0x−1 and x2, x5 ∈ x0x + 1 . From the first assertion of (E5) we know that x2, x3 ∈ x1x+5 . Then when we rotate x0x1 counterclockwise around x1, the first point that such line finds must be x3, and so the desired result follows immediately from this and the first assertion of (E5). (E11): From the first assertion of (E7), we have that x3, x4 ∈ x0x−1 and x2, x5 ∈ x0x + 1 . From (E8), we know that x2, x3 ∈ x1x−4 . Then when we rotate x0x1 clockwise around x1, the first point that we find is x2, and so the desired result follows from the first assertion of (E7). (E12): (E8) implies {x2, x3, x4} ∪ Z = x1x−5 . From (D4) and the second assertion of (E3), we know that when we rotate x1x5 clockwise around x5, the first point that we find must be x2, and so the desired result follows from (E8). (E13): (E4) implies that Z ∪ {x0, x1, x3} = x2x−4 . From (E10), we know that x2 ∈ x1x + 3 . This and the first assertion of (E3) imply that when we rotate x2x4 counterclockwise around x2, the first point that we find must be x3, and so the desired result follows from (E4). 4.5 Determination of pq+ when pq is a bichromatic (>5)-edge. We are finally ready to prove Lemma 3.1 for the remaining case, namely when pq is a bichromatic (> 5)-edge. This is achieved in the next statement. We recall from Re- mark 4.19 and Corollary 4.20 that x4x5 splits the ys, y4y5 splits the zs, and z4z5 splits the xs. Proposition 4.24. Let u ∈ {x, y, z}, and let v be the element in {x, y, z} \ {u} such that u4u5 splits the vs. Then the following hold: (F1) u5v+4 = {u0, u1, u2, u3} ∪ {v0, v1, v2, v3}; (F2) u5v+5 = {u0, u1, u2} ∪ {v0, v1, v2, v3, v4}; (F3) u5v+3 = {u0, u1, u2, u3, u4} ∪ {v0, v1, v2}; (F4) u5v+2 = {u0, u1, u2, u3, u4} ∪ {v0, v1}; (F5) u5v+1 = {u0, u1, u2, u3, u4} ∪ {v0}; (F6) u4v+4 = {u0, u1, u2, u3, u5} ∪ {v0, v1, v2, v3}; (F7) u4v+5 = {u0, u1, u2, u3, u5} ∪ {v0, v1, v2, v3, v4}; 378 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 (F8) u4v+3 = {u0, u1, u2, u3} ∪ {v0, v1, v2}; (F9) u4v+2 = {u0, u1, u2, u3} ∪ {v0, v1}; (F10) u3v+5 = {u0, u1, u2, u5} ∪ {v0, v1, v2, v3, v4}; (F11) u3v+4 = {u0, u1, u2} ∪ {v0, v1, v2, v3}; (F12) u3v+3 = {u0, u1, u2} ∪ {v0, v1, v2}; (F13) u2v+5 = {u0, u1} ∪ {v0, v1, v2, v3, v4}; (F14) u2v+4 = {u0, u1} ∪ {v0, v1, v2, v3}; (F15) u1v+5 = {u0} ∪ {v0, v1, v2, v3, v4}. Proof. In view of Remark 4.19 and Corollary 4.20, we may assume that the points of P are placed as in Figure 15. Moreover, by symmetry, we only need to verify the case u = x and v = y. For brevity, form ∈ {0, . . . , 5}, we letXm := {xi ∣∣i ≤ m} and Ym := {yi∣∣i ≤ m}. (F1): From (D3), we know that x4x+5 = X3 ∪ Y3. We claim that the first point p ∈ P that x4x5 finds when it is rotated counterclockwise around x5 is y4. From (D4), we know that x3x+5 = X2 ∪ Y4. Then p = y4, and so x5y + 4 = X3 ∪ Y3, as required. (F2): We know that x3x+5 = X2 ∪ Y4 by (D4). We claim that the first point p ∈ P that x3x5 finds when it is rotated counterclockwise around x5 is y5. Indeed, from (D1), (E8), and (E12), we know that y5 ∈ xjx+5 for j = 0, 1, 2, respectively. These imply that p /∈ X2, and so p = y5. Then x5y+5 = X2 ∪ Y4, as required. (F3): We know that x4x+5 = X3 ∪ Y3 by (D3). We claim that the first point p ∈ P that x4x5 finds when it is rotated clockwise around x5 is y3. Indeed, by applying (E7), (E10), and (E13) to j = 0, 1 and j = 2 (with u = y and v = z), respectively, we have that X ⊂ yjy−3 , and so x5 ∈ yjy − 3 . These imply that p /∈ Y2. This and the fact that x5y + 0 = X4 imply that p = y3. Then x5y+3 = X4 ∪ Y2, as required. (F4): We know that x5y+3 = X4 ∪ Y2 by (F3). We claim that the first point p ∈ P that x5y3 finds when it is rotated clockwise around x5 is y2. Indeed, by applying (E6) and (E11) to j = 0 and j = 1 (with u = y and v = z), respectively, we have that X ⊂ yjy−2 , and so x5 ∈ yjy−2 . These imply that p /∈ Y1. This and the fact that x5y + 0 = X4 imply that p = y2. Then x5y+2 = X4 ∪ Y1, as required. (F5): We know that x5y+2 = X4 ∪ Y1 by (F4). We claim that the first point p ∈ P that x5y2 finds when it is rotated clockwise around x5 is y1. Indeed, by taking u = y in (E7), we have that x5 ∈ y0y−1 , and so p ̸= y0. This and the fact that x5y + 0 = X4 imply that p = y1. Then x5y+1 = X4 ∪ Y0, as required. (F6): We know that x4x+5 = X3 ∪ Y3 by (D3). We claim that the first point p ∈ P that x4x5 finds when it is rotated counterclockwise around x4 is y4. Indeed, by taking u = y and v = z in D3) we get x4 ∈ y4y−5 , and so p ̸= y5. This and the fact that x0x + 4 = {x1, x2, x3, x5} ∪ Y imply that p = y4. Then x4y + 4 = X3 ∪ Y3 ∪ {x5}, as required. (F7): We know that x4y+4 = X3 ∪ Y3 ∪ {x5} by (F6). We claim that the first point p ∈ P that x4y4 finds when it is rotated counterclockwise around x4 is y5. Indeed, from (D2), we know that x0x+4 = {x1, x2, x3, x5}∪Y . These imply that p /∈ X , and so p = y5. Then x4y+5 = X3 ∪ Y4 ∪ {x5}, as required. B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 379 (F8): We know that x4x+5 = X3 ∪ Y3 by (D3). We claim that the first point p ∈ P that x4x5 finds when it is rotated clockwise around x4 is y3. Indeed, by applying E7), E10), and E13) to j = 0, 1 and j = 2 (with u = y and v = z), respectively, we have that X ⊂ yjy−3 , and so x4 ∈ yjy−3 . These imply that p /∈ Y2. From Proposition 4.3(2), we know that X3 ⊂ x4y+1 , and so p /∈ X3. All these facts imply that p = y3, and so x4y + 3 = X3 ∪Y2, as required. (F9): We know that x4y+3 = X3 ∪ Y2 by (F8). We claim that the first point p ∈ P that x4y3 finds when it is rotated clockwise around x4 is y2. Indeed, by applying (E6) and (E11) to j = 0 and j = 1 (with u = y and v = z), respectively, we have that X ⊂ yjy−2 , and so x4 ∈ yjy−2 . These imply that p /∈ Y1. From this and (D3) it follows that p = y2, and so x4y+2 = X3 ∪ Y1, as required. (F10): We know that x3x+5 = X2∪Y4 by (D4). We claim that the first point p ∈ P that x3x5 finds when it is rotated counterclockwise around x3 is y5. Indeed, by applying (E7), (E10), and (E13) to j = 0, 1 and j = 2 (with u = x and v = y), respectively, we have that Y ⊂ xjx+3 , and so y5 ∈ xjx + 3 . These imply that p /∈ X2. From this and E5) it follows that p = y5, and so x3y+5 = X2 ∪ Y4 ∪ {x5}, as required. (F11): We know that x3x+5 = X2 ∪ Y4 by D4). We claim that the first point p ∈ P that x3x5 finds when it is rotated clockwise around x3 is y4. Indeed, by applying (D2), (E5), (E4), and (E5) to j = 0, 1, 2 and j = 3 (with u = y and v = z), respectively, we have that X ⊂ yjy−4 , and so x3 ∈ yjy − 4 . These imply that p /∈ Y3. From Proposition 4.3(2), we know that x4 ∈ x3y−2 , and so p ̸= x4. All these facts imply that p = y4, and so x3y + 4 = X2 ∪ Y3, as required. (F12): We know that x3y+4 = X2 ∪ Y3 by (F11). We claim that the first point p ∈ P that x3y4 finds when it is rotated clockwise around x3 is y3. Indeed, by applying (E7), (E10), and (E13) to j = 0, 1 and j = 2 (with u = y and v = z), respectively, we have that X ⊂ yjy−3 , and so x3 ∈ yjy − 3 . These imply that p /∈ Y2. By taking u = x in (E5) and (D4), we have that y4 ∈ x3x+4 and y4 ∈ x3x + 5 , respectively. These imply that p /∈ {x4, x5}. All these facts imply that p = y3, and so x3y+3 = X2 ∪ Y2, as required. (F13): From Proposition 4.3(2), we know that x2y+3 = X1 ∪Y2. We claim that the first point p ∈ P that x2y3 finds when it is rotated counterclockwise around x2 is y4. Indeed, by applying (E6) and (E11) to j = 0 and j = 1 (with u = x and v = y), respectively, we have that Y ⊂ xjx+2 , and so p /∈ {x0, x1}. Finally, by applying (E4), (E12), and (E13) to j = 4, 5 and j = 3 (with u = x and v = y), we have that p ̸= x4, p ̸= x5 and p ̸= x3, respectively. All these facts imply that p = y4, and so x2y+4 = X1 ∪ Y3, as required. (F14): We know that x2y+4 = X1 ∪ Y3 by (F13). We claim that the first point p ∈ P that x2y4 finds when it is rotated counterclockwise around x2 is y5. As in (F13) we can deduce from (E6) and (E11) that p /∈ {x0, x1}. Again, as in (F13) we can deduce from (E4), (E12), and (E13) that p ̸= x4, p ̸= x5 and p ̸= x3, respectively. All these facts imply that p = y5, and so x2y+5 = X1 ∪ Y4, as required. (F15): We know that x2y+5 = X1 ∪ Y4 by F14). We claim that the first point p ∈ P that x2y5 finds when it is rotated counterclockwise around y5 is x1. Indeed, from Proposi- tion 4.3(2), we know that y5z+0 = Y4. From the last two equations we have that p ∈ X1 = {x0, x1}. Again, from Proposition 4.3(2), we know that x0y+5 = Y4, and so p = x1. Then x1y + 5 = X0 ∪ Y4, as required. 380 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 4.6 Conclusion of the proof of Lemma 3.1 In Tables 1 and 2 we give a summary of the results in Propositions 4.3(2)(b), 4.21, 4.22, and 4.24. These tables assume that u ∈ {x, y, z} and v = σ(u), where σ is the automor- phism of {x, y, z} defined in Corollary 4.20, namely x σ7→ y, y σ7→ z, and z σ7→ x. In particular, for each u ∈ {x, y, z} and m,n ∈ {0, . . . , 5} with m < n, the set umu+n is given in Table 1. This also determines the set unu+m, since unu + m = umu − n , and umu − n is evidently determined from umu+n . Thus the information in Table 1 suffices to determine pq+ whenever pq is a monochromatic edge of P . Now for each u ∈ {x, y, z} and each m,n ∈ {0, . . . , 5}, the set umv+n is given in Table 2. This also determines the set vnu+m, since vnu + m = umv − n , and umv − n is evidently determined from umv+n . Thus the information in Table 2 suffices to determine pq + when- ever pq is a bichromatic edge of P . □ uiu + j for each uiuj ∈ Emonk (P ) Classification of uiuj Equality stated in u0u + 5 = {v0, . . . , v5} 6-edge (D1) u0u + 4 = {u1, u2, u3, u5} ∪ {v0, . . . , v5} 6-edge (D2) u0u + 3 = {u1, u2, u5} ∪ {v0, . . . , v5} 7-edge (E7) u0u + 2 = {u5} ∪ {v0, . . . , v5} 7-edge (E6) u0u + 1 = {u2, u5} ∪ {v0, . . . , v5} 8-edge (E7) u1u + 5 = {u0} ∪ {v0, . . . , v5} 7-edge (E8) u1u + 4 = {u2, u3, u5} ∪ {v0, . . . , v5} 7-edge (E5) u1u + 3 = {u2, u5} ∪ {v0, . . . , v5} 8-edge (E10) u1u + 2 = {u0, u5} ∪ {v0, . . . , v5} 8-edge (E11) u2u + 5 = {u0, u1} ∪ {v0, . . . , v5} 8-edge (E12) u2u + 4 = {u5} ∪ {v0, . . . , v5} 7-edge (E4) u2u + 3 = {u4, u5} ∪ {v0, . . . , v5} 8-edge (E13) u3u + 5 = {u0, u1, u2} ∪ {v0, . . . , v4} 8-edge (D4) u3u + 4 = {u2, u5} ∪ {v0, . . . , v5} 8-edge (E5) u4u + 5 = {u0, . . . , u3} ∪ {v0, . . . , v3} 8-edge (D3) Table 1: All the monochromatic edges of P . 5 Concluding remarks In this work we finally have given the full proof of Theorem 1.2, which was announced at the EuroComb’11 conference [1]. As we mentioned in the Introduction, the exact rectilinear crossing number of Kn is known only for n ≤ 27 and n = 30 [3, 7, 8, 9, 10]. In [2] and [6] we can find non- isomorphic crossing-minimal rectilinear drawings of Kn for both n = 24 and n = 30. On the other hand, from [6] and the main result of this work, now we know that there is a unique (up to order type isomorphism) crossing-minimal rectilinear drawing of Kn, for n = 6, 12, 18. Thus, a plausible conjecture is that K6m has several crossing-minimal rectilinear drawings for each integer m ≥ 4. We close this paper with a discussion on a question raised by an anonymous reviewer of an earlier version of this paper: to what degree would it be possible to get a computer- assisted proof of Theorem 1.2? B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 381 uv+ for each uv ∈ Ebik (P ) Classification of uv Equality stated in u0v + 5 = {v0, . . . , v4} 5-edge Proposition 4.3(2) u0v + 4 = {v0, v1, v2, v3} 4-edge Proposition 4.3(2) u0v + 3 = {v0, v1, v2} 3-edge Proposition 4.3(2) u0v + 2 = {v0, v1} 2-edge Proposition 4.3(2) u0v + 1 = {v0} 1-edge Proposition 4.3(2) u0v + 0 = ∅ 0-edge Proposition 4.3(2) u1v + 5 = {u0} ∪ {v0, . . . , v4} 6-edge (F15) u1v + 4 = {u0} ∪ {v0, v1, v2, v3} 5-edge Proposition 4.3(2) u1v + 3 = {u0} ∪ {v0, v1, v2} 4-edge Proposition 4.3(2) u1v + 2 = {u0} ∪ {v0, v1} 3-edge Proposition 4.3(2) u1v + 1 = {u0} ∪ {v0} 2-edge Proposition 4.3(2) u1v + 0 = {u0} 1-edge Proposition 4.3(2) u2v + 5 = {u0, u1} ∪ {v0, . . . , v4} 7-edge (F14) u2v + 4 = {u0, u1} ∪ {v0, v1, v2, v3} 6-edge (F13) u2v + 3 = {u0, u1} ∪ {v0, v1, v2} 5-edge Proposition 4.3(2) u2v + 2 = {u0, u1} ∪ {v0, v1} 4-edge Proposition 4.3(2) u2v + 1 = {u0, u1} ∪ {v0} 3-edge Proposition 4.3(2) u2v + 0 = {u0, u1} 2-edge Proposition 4.3(2) u3v + 5 = {u0, u1, u2, u5} ∪ {v0, . . . , v4} 7-edge (F10) u3v + 4 = {u0, u1, u2} ∪ {v0, v1, v2, v3} 7-edge (F11) u3v + 3 = {u0, u1, u2} ∪ {v0, v1, v2} 6-edge (F12) u3v + 2 = {u0, u1, u2} ∪ {v0, v1} 5-edge Proposition 4.3(2) u3v + 1 = {u0, u1, u2} ∪ {v0} 4-edge Proposition 4.3(2) u3v + 0 = {u0, u1, u2} 3-edge Proposition 4.3(2) u4v + 5 = {u0, u1, u2, u3, u5} ∪ {v0, . . . , v4} 6-edge (F7) u4v + 4 = {u0, u1, u2, u3, u5} ∪ {v0, . . . , v3} 7-edge (F6) u4v + 3 = {u0, u1, u2, u3} ∪ {v0, v1, v2} 7-edge (F8) u4v + 2 = {u0, u1, u2, u3} ∪ {v0, v1} 6-edge (F9) u4v + 1 = {u0, u1, u2, u3} ∪ {v0} 5-edge Proposition 4.3(2) u4v + 0 = {u0, u1, u2, u3} 4-edge Proposition 4.3(2) u5v + 5 = {u0, u1, u2} ∪ {v0, . . . , v4} 8-edge (F2) u5v + 4 = {u0, . . . , u3} ∪ {v0, . . . , v3} 8-edge (F1) u5v + 3 = {u0, . . . , u4} ∪ {v0, v1, v2} 8-edge (F3) u5v + 2 = {u0, . . . , u4} ∪ {v0, v1} 7-edge (F4) u5v + 1 = {u0, . . . , u4} ∪ {v0} 6-edge (F5) u5v + 0 = {u0, u1, u2, u3, u4} 5-edge Proposition 4.3(2) Table 2: All the bichromatic edges of P . 382 Ars Math. Contemp. 24 (2024) #P2.10 / 355–383 At the beginning of this project we asked ourselves the same question, but we are con- vinced that a traditional proof might be easier to verify. It is worth mentioning that a heav- ily computer-assisted proof seems to be out of reach, most likely involving several hundred million CPU hours. On the other hand, we believe that a partially computer-assisted proof would be more difficult to follow and perhaps also less reliable. Using computer-assisted proofs needs a very careful preparation and description of what is done, and proofs of the correctness of the results. The code must be explained in full detail, as well as how the program can be executed (including the operating system, compiler versions, etc.). We believe that in this particular case the task of verifying all this information would end up being more taxing on the reader than the current purely theoretical proof. ORCID iDs Bernardo M. Ábrego https://orcid.org/0000-0003-4695-5454 Silvia Fernández–Merchant https://orcid.org/0000-0003-2080-106X Oswin Aichholzer https://orcid.org/0000-0002-2364-0583 Jesús Leaños https://orcid.org/0000-0002-3441-8136 Gelasio Salazar https://orcid.org/0000-0002-8458-3930 References [1] B. M. Ábrego, O. Aichholzer, S. Fernández-Merchant, J. Leaños and G. Salazar, There is a unique crossing-minimal rectilinear drawing of K18, in: Extended abstracts of the sixth Euro- pean conference on combinatorics, graph theory and applications, EuroComb 2011, Budapest, Hungary, August 29 – September 2, 2011, Elsevier, Amsterdam, pp. 547–552, 2011, doi: 10.1016/j.endm.2011.09.089, https://doi.org/10.1016/j.endm.2011.09.089. [2] B. M. Ábrego, M. Cetina, S. Fernández-Merchant, J. Leaños and G. Salazar, 3-symmetric and 3-decomposable geometric drawings of Kn, Discrete Appl. Math. 158 (2010), 1240–1258, doi: 10.1016/j.dam.2009.09.020, https://doi.org/10.1016/j.dam.2009.09.020. [3] B. M. Ábrego, M. Cetina, S. Fernández-Merchant, J. L. nos and G. Salazar, On ≤ k- edges, crossings, and halving lines of geometric drawings of kn, Discrete Comput. Geom. 48 (2012), 192–215, doi:10.1007/s00454-012-9403-y, https://doi.org/10.1007/ s00454-012-9403-y. [4] B. M. Ábrego and S. Fernández-Merchant, A lower bound for the rectilinear crossing number, Graphs Comb. 21 (2005), 293–300, doi:10.1007/s00373-005-0612-5, https://doi.org/ 10.1007/s00373-005-0612-5. [5] B. M. Ábrego, S. Fernández-Merchant and G. Salazar, The rectilinear crossing num- ber of Kn: closing in (or are we?), in: Thirty Essays on Geometric Graph Theory, Springer, Berlin, pp. 5–18, 2013, doi:10.1007/978-1-4614-0110-0 2, https://doi.org/ 10.1007/978-1-4614-0110-0_2. [6] O. Aichholzer, http://www.ist.tugraz.at/aichholzer/research/rp/ triangulations/crossing/. [7] O. Aichholzer, J. Garcia, D. Orden and P. Ramos, New lower bounds for the number of (≤ k)- edges and the rectilinear crossing number of Kn, Discrete Comput. Geom. 38 (2007), 1–14, doi: 10.1007/s00454-007-1325-8, https://doi.org/10.1007/s00454-007-1325-8. [8] O. Aichholzer, J. Garcı́a, D. Orden and P. Ramos, New results on lower bounds for the num- ber of (⩽ k)-facets, in: Proceedings of the 4th European conference on combinatorics, graph B. M. Ábrego et al.: There is a unique crossing-minimal rectilinear drawing of K18 383 theory and applications, EuroComb’07, Seville, Spain, September 11–15, 2007, Elsevier, Am- sterdam, pp. 189–193, 2007, doi:10.1016/j.endm.2007.07.033, https://doi.org/10. 1016/j.endm.2007.07.033. [9] O. Aichholzer and H. Krasser, Abstract order type extension and new results on the recti- linear crossing number, Comput. Geom. 36 (2007), 2–15, doi:10.1016/j.comgeo.2005.07.005, https://doi.org/10.1016/j.comgeo.2005.07.005. [10] M. Cetina, C. Hernández-Vélez, J. Leaños and C. Villalobos, Point sets that minimize (≤ k)- edges, 3-decomposable drawings, and the rectilinear crossing number of K30, Discrete Math. 311 (2011), 1646–1657, doi:10.1016/j.disc.2011.03.030, https://doi.org/10.1016/ j.disc.2011.03.030. [11] R. K. Guy, A combinatorial problem, (Nabla) Bull. Malays. Math. Sci. Soc. 7 (1960), 68–72. [12] L. Lovász, K. Vesztergombi, U. Wagner and E. Welzl, Convex quadrilaterals and k-sets, in: Towards a Theory of Geometric Graphs, American Mathematical Society (AMS), Providence, RI, pp. 139–148, 2004. Author Guidelines Before submission Papers should be written in English, prepared in LATEX, and must be submitted as a PDF file. The title page of the submissions must contain: • Title. The title must be concise and informative. • Author names and affiliations. For each author add his/her affiliation which should include the full postal address and the country name. If avilable, specify the e-mail address of each author. Clearly indicate who is the corresponding author of the paper. • Abstract. A concise abstract is required. The abstract should state the problem stud- ied and the principal results proven. • Keywords. Please specify 2 to 6 keywords separated by commas. • Mathematics Subject Classification. Include one or more Math. Subj. Class. (2020) codes – see https://mathscinet.ams.org/mathscinet/msc/msc2020.html. After acceptance Articles which are accepted for publication must be prepared in LATEX using class file amcjoucc.cls and the bst file amcjoucc.bst (if you use BibTEX). If you don’t use BibTEX, please make sure that all your references are carefully formatted following the examples provided in the sample file. All files can be found on-line at: https://amc-journal.eu/index.php/amc/about/submissions/#authorGuidelines Abstracts: Be concise. As much as possible, please use plain text in your abstract and avoid complicated formulas. Do not include citations in your abstract. All abstracts will be posted on the website in fairly basic HTML, and HTML can’t handle complicated formulas. It can barely handle subscripts and greek letters. Cross-referencing: All numbering of theorems, sections, figures etc. that are referenced later in the paper should be generated using standard LATEX \label{. . . } and \ref{. . . } commands. See the sample file for examples. Theorems and proofs: The class file has pre-defined environments for theorem-like state- ments; please use them rather than coding your own. Please use the standard \begin{proof} . . . \end{proof} environment for your proofs. Spacing and page formatting: Please do not modify the page formatting and do not use \medbreak, \bigbreak, \pagebreak etc. commands to force spacing. In general, please let LATEX do all of the space formatting via the class file. The layout editors will modify the formatting and spacing as needed for publication. Figures: Any illustrations included in the paper must be provided in PDF format, or via LATEX packages which produce embedded graphics, such as TikZ, that compile with PdfLATEX. (Note, however, that PSTricks is problematic.) Make sure that you use uniform lettering and sizing of the text. If you use other methods to generate your graphics, please provide .pdf versions of the images (or negotiate with the layout editor assigned to your article). xiii Subscription Yearly subscription: 150 EUR Any author or editor that subscribes to the printed edition will receive a complimentary copy of Ars Mathematica Contemporanea. Subscription Order Form Name: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . E-mail: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Postal Address: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . I would like to subscribe to receive . . . . . . copies of each issue of Ars Mathematica Contemporanea in the year 2024. I want to renew the order for each subsequent year if not cancelled by e-mail: □ Yes □ No Signature: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Please send the order by mail, by fax or by e-mail. By mail: Ars Mathematica Contemporanea UP FAMNIT Glagoljaška 8 SI-6000 Koper Slovenia By fax: +386 5 611 75 71 By e-mail: info@famnit.upr.si xiv Printed in Slovenia by IME TISKARNE