Volume 24, Number 3, Fall/Winter 2024, Pages 385–566 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 Generalized Cayley maps and their Petrie duals Robert Jajcay, Jozef Širáň, Yan Wang . . . . . . . . . . . . . . . . . . . . 385 On p-gonal fields of definition Ruben A. Hidalgo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405 An online bin-packing problem with an underlying ternary structure Helmut Prodinger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415 Variants of the domination number for flower snarks Ryan Burdett, Michael Haythorpe, Alex Newcombe . . . . . . . . . . . . . 427 Optimal strategies in fractional games: vertex cover and domination Csilla Bujtás, Günter Rote, Zsolt Tuza . . . . . . . . . . . . . . . . . . . . 453 The Sierpiński domination number Michael A. Henning, Sandi Klavžar, Elżbieta Kleszcz, Monika Pilśniak . . 473 On the minisymposium problem Peter Danziger, Eric Mendelsohn, Brett Stevens, Tommaso Traetta . . . . . 489 Two kinds of partial Motzkin paths with air pockets Jean-Luc Baril, Paul Barry . . . . . . . . . . . . . . . . . . . . . . . . . . 509 Independent coalition in graphs: existence and characterization Mohammad Reza Samadzadeh, Doost Ali Mojdeh . . . . . . . . . . . . . . 535 Spectra of signed graphs and related oriented graphs Zoran Stanić . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553 Volume 24, Number 3, Fall/Winter 2024, Pages 385–566 xix ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.01 / 385–403 https://doi.org/10.26493/1855-3974.2408.42e (Also available at http://amc-journal.eu) Generalized Cayley maps and their Petrie duals* Robert Jajcay † Comenius University, Bratislava, Slovakia Jozef Širáň ‡ Slovak University of Technology, Bratislava, Slovakia Yan Wang § Yan Tai University, Yan Tai, China Received 12 August 2020, accepted 22 October 2023, published online 6 June 2024 Abstract Cayley maps are embeddings of Cayley graphs in orientable surfaces which possess a group of orientation preserving automorphisms acting regularly on the vertices. We gen- eralize the concept of a Cayley map by considering embeddings of Cayley graphs in both orientable and non-orientable surfaces and by requiring a group of automorphisms acting regularly on vertices that does not have to consist entirely of orientation preserving auto- morphisms. This leads to new families of maps in both the orientable and non-orientable cases. Since the Petrie dual operator preserves the property of being a generalized Cayley map, throughout the paper we consider the action of this operator on our maps. Keywords: Cayley map, regular embedding, automorphism group of a map. Math. Subj. Class. (2020): 05C10, 05C25, 05E18 *We wish to thank our referees for their repeated careful readings of our article and their helpful and knowl- edgeable suggestions that helped considerably in improving our article and making it more focused. †Corresponding author. Supported by VEGA Research Grant 1/0437/23 and APVV Research Grant 19-0308. ‡Supported by APVV Research Grants 19-0308 and 22-0005, and VEGA Research Grants 1/0567/22 and 1/0069/23. §Supported by NSFS No. ZR2020MA044. E-mail addresses: robert.jajcay@fmph.uniba.sk (Robert Jajcay), jozef.siran@stuba.sk (Jozef Širáň), wang−yan@pku.org.cn (Yan Wang) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 386 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 1 Introduction A map is a cellular embedding of a connected graph on some surface; the map is orientable if the surface is. By a widely adopted definition (cf. [16]), an orientable map M is a Cayley map if the group of orientation-preserving automorphisms of M contains a subgroup acting regularly on the vertex set of M. This way of introducing Cayley maps has been motivated by Sabidussi’s well known characterization of Cayley graphs in terms of a subgroup of automorphisms of the graph acting regularly on its vertex set [18]. Apart from the history of the development of the theory of maps, however, when carrying Sabidussi’s characteri- zation over to maps there is no reason to restrict to orientation-preserving automorphisms, or even to orientable surfaces. The latter may have been first realized by Tucker [19], and in this paper we argue for the return of his way of treating Cayley maps. In order not to clash with customary terminology we will say that a map M on an arbitrary surface (orientable or not) is a generalized Cayley map if the automorphism group of M contains a subgroup H that is a regular permutation group on the vertex set of M. In our view, Tucker’s definition by means of existence of a regular action of a group of automorphisms has two major advantages: it is a generalization of the currently used defi- nition that is simple, and in line with the original definition of a Cayley map. To the best of our knowledge there have been two other attempts at defining non-orientable Cayley maps: the one by Abas [1] via strategically placed cross-cups on the carrier surface, and another one by Kwak and Kwon [12] who developed a full theory of generalized Cayley maps (equivalent to Tucker’s and our approach) but starting off from introducing generalized Cayley maps in a way that appears to be hard to use in applications. The main difference between (original and orientable) Cayley maps and generalized Cayley maps is that the carrier surfaces of the latter may be non-orientable, and even in the orientable case the subgroup required to act regularly on vertices does not have to be orientation-preserving. Maps admitting such subgroups appear frequently in the context of regular maps with automorphism groups acting quasi-primitively on the vertices [6], and our study of such maps was the main source of motivation for the present paper. At several places, we employ the Petrie dual operator (e.g., [22]) which proves par- ticularly useful since it preserves the class of generalized Cayley maps. For example, we consider orientable generalized Cayley maps whose Petrie dual is again an orientable gen- eralized Cayley map, as well as non-orientable generalized Cayley maps whose Petrie dual is orientable. The outline of the paper is as follows. In order to emphasize parallels between the original Cayley maps on orientable surfaces and the generalized Cayley maps, the three sections following this Introduction are devoted to reviewing the background and context of highly symmetric maps, Cayley maps, and the Petrie dual operator; with Sections 5 and 6 containing the new results. We begin in Section 2 by reviewing some of the most important properties of regular maps, followed in Section 3 by reviewing algebraic machinery for dealing with orientable Cayley maps in their original setting. Section 4 discusses Petrie duals of maps in some detail. We discuss generalized Cayley maps on orientable surfaces in Section 5 and generalized Cayley maps on non-orientable surfaces in Section 6. We can summarize the results obtained in our paper as follows. A generalized Cayley map may be both orientable and non-orientable. The orientable generalized Cayley maps come in two kinds. First, there are the original orientable Cayley maps in which the group acting regularly on the vertices of the map consists entirely of orientation preserving auto- morphisms. Then there are orientable generalized Cayley maps with bipartite underlying R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 387 graphs in which the group acting regularly on the vertices of the map contains a subgroup of index 2 consisting of orientation preserving automorphisms while the other half of the regularly acting group consists of orientation reversing automorphisms (Theorems 5.1 and 5.2). There are infinitely many examples of orientable generalized Cayley maps which are not the original Cayley maps (Lemma 5.3 and Example 5.5), and infinitely many ori- entable generalized Cayley maps which are self-Petrie-dual (Remark 5.6). In Remark 5.7 we present orientable generalized Cayley maps that simultaneously admit an orientation preserving group acting regularly on their vertices and a vertex-regular group half of which consists of orientation reversing automorphisms (hence, these maps are classical Cayley maps and orientable maps of the second kind at the same time). In Remark 6.2 we present an infinite family of non-orientable generalized Cayley maps. Taking advantage of the well-known fact that the Petrie dual of an orientable map M is orientable if and only if the underlying graph of M is bipartite (Theorem 4.2 in our paper), we argue the existence of infinitely many examples of the original orientable Cayley maps whose Petrie duals are orientable (Example 6.6) and infinitely many examples of the original orientable Cayley maps whose Petrie duals are non-orientable (any orientable Cayley map with a non-bipartite underlying graph). The Petrie dual of a non-orientable generalized Cayley map may be both orientable and non-orientable. We classify the non- orientable generalized Cayley maps whose Petrie dual is orientable (Corollary 6.5) and exhibit infinitely many examples of both non-orientable generalized Cayley maps whose Petrie dual is orientable and non-orientable generalized Cayley maps whose Petrie dual is non-orientable (Remark 6.2). 2 Regular, orientably-regular, reflexible and chiral maps For a map M, regions (faces) of its barycentric subdivision are the flags of the map. Every flag is a triangular region; informally, its three ‘corners’ are a vertex, the ‘midpoint’ of an edge incident with the vertex, and the ‘centre’ of a face incident with both the vertex and the edge. As long as no face of the map contains two occurrences of an edge on its boundary (no maps with such a degeneracy will be considered here), a flag can be identified with a triple (v, e, F ), where v is a vertex, e an edge incident with v, and F is a face incident with both v and e. A pair of distinct flags (v, e, F ) and (v′, e′, F ′) of M are incident if they share a segment of the skeleton of the barycentric subdivision; in case when no face of the map contains two occurrences of an edge on its boundary, two flags are incident if precisely two of the three equalities v = v′, e = e′, F = F ′ hold. Let r0, r1 and r2 be involutory fixed-point-free permutations of the flag set of a map M formed by two-cycles consisting of incident flags, with r0 swapping faces that share a face-center-to-edge-midpoint segment, r1 swapping faces that share the vertex-to-face- center segment, and r2 swapping faces that share the vertex-to-edge-midpoint segment; again, if no face of the map contains two occurrences of an edge on its boundary, the flag (v, e, F ) is mapped by r0 to (v′, e, f), v ̸= v′, by r1 to (v, e′, F ), e′ ̸= e, and by r2 to (v, e, F ′), F ′ ̸= F . The (flag-transitive) permutation group generated by r0, r1 and r2 is the monodromy group Mon(M) of the map. The three generators of the monodromy group can be interpreted as ‘gluing instructions’ to assemble the map from its set of flags; in other words, knowledge of Mon(M) in terms of the action of the generators r0, r1 and r2 on the set its flags is equivalent to knowing the map M. Note that r0 and r2 commute; if the map is finite, then the orders of r0r1 and r1r2 are equal to the least common multiples of face 388 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 lengths and vertex valencies, respectively. It is well known that the carrier surface of M is orientable if and only if ⟨r1r2, r0r2⟩ = ⟨r1r2, r0r1⟩ is a subgroup of index 2 in Mon(M). In such a case we speak about an orientable map; letting ρ = r1r2 and λ = r0r2 one may consider the ‘orientable part’ Mon+(M) = ⟨ρ, λ⟩, the index-2-subgroup of Mon(M). Here, Mon+(M) can be re- garded as a permutation group acting on the dart set D of the map (i.e., on the set of directed edges of the map). Then ρ is a permutation that cyclically permutes, at each vertex v, the darts emanating from v in accord with a chosen orientation of the carrier surface of the map, and λ is an an involution interchanging the two darts belonging to the same edge. An automorphism of a map M with monodromy group Mon(M) = ⟨r0, r1, r2⟩ is any permutation of the flag set F of M that preserves incidence of flags. Equivalently, a permutation of F is an automorphism of M if and only if it commutes with all of r0, r1 and r2; hence the automorphism group Aut(M) of the map is simply the centralizer of the monodromy group Mon(M) in the symmetric group SF on the set F . Since Mon(M) is transitive on F , it follows that Aut(M) acts freely on the set set F . For an orientable analogue, let now M be an orientable map with dart set D and with Mon+(M) = ⟨ρ, λ⟩ being a subgroup of Mon(M) of index 2. An orientation-preserving automorphism of M is any permutation of D commuting with ρ and λ. It follows that the group Aut+(M) of all orientation-preserving automorphisms of an orientable map is the centralizer of Mon+(M) in the symmetric group SD on the set D. Again, transitivity of Mon+(M) on D implies that the group Aut+(M) acts freely on the set D. If such an orientable map M admits an automorphism commuting with λ but inverting ρ, then the automorphism is orientation-reversing. Finally, let us discuss the highest ‘level of symmetry’of maps. A map M with mon- odromy group Mon(M) is called regular if the automorphism group M acts regularly on the set F of flags. In this case, the groups Aut(M) and Mon(M) are abstractly iso- morphic, so that in the case of a finite map the order of both groups is equal to |F|. If M is a regular orientable map, then its automorphism group Aut(M) contains the group Aut+(M) of orientation-preserving automorphisms as a subgroup of index 2, its other coset in Aut(M) being the collection of all orientation-reversing automorphisms of the map. The orientation-reversing automorphisms are sometimes referred to as reflections, giving such maps the name reflexible. An orientable map M is orientably-regular if the group Aut+(M) of orientation preserving automorphisms of M acts regularly on its set of darts. An orientably regular map that is not reflexible is chiral, and in such a case Aut(M) = Aut+(M). 3 Cayley maps A Cayley map is an orientable map M that admits a group of orientation preserving auto- morphisms G acting regularly on its set of vertices. To make this definition more precise, one needs to consider the induced action of map automorphisms of M on the vertex set of the underlying graph associating the map automorphism ψ with the vertex permutation ψ mapping v to v′ whenever ψ(v, e, F ) = (v′, e′, F ′). It is easy to see that ψ is a well defined graph automorphism of the underlying graph of M defined via its action on the vertices (of the map or the graph). The induced action of Aut(M) on the vertices is almost universally faithful; with the very few exceptions listed in [14, Proposition 9]. The auto- morphisms contained in the automorphism group of a Cayley map which acts regularly on R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 389 the vertices of the map correspond in this way to automorphisms of its underlying graph forming a group acting regularly on the graph’s vertices, making it into a Cayley graph [18]. A Cayley graph C(G,X) is a graph whose vertex set can be identified with the elements of a group G generated by a set X closed under taking inverses and not containing the identity 1G, with the pairs of adjacent vertices consisting of all pairs g, gx with g ∈ G and x ∈ X . A graph Γ is isomorphic to a Cayley graph C(G,X) if and only if AutΓ contains a subgroup G acting regularly on the vertices of Γ [18], which justifies the remark we have made about the underlying graphs of Cayley maps being Cayley. Hence, the automorphism group of a Cayley map M containing a group G of orientation preserving automorphisms acting regularly on the vertices of M is necessarily contained in the automorphism group of its underlying Cayley graph C(G,X), G ≤ AutC(G,X) (after applying the necessary restriction to vertices). The action of G on C(G,X) is defined for each g ∈ G via left multiplication: Ag(h) = gh for all g, h ∈ H , and we shall denote the group {Ag | g ∈ G} byGL; it is always a subgroup of AutC(G,X). The set of darts of the Cayley map M thus takes the form {(h, hx) | h ∈ G, x ∈ X} and each Ag ∈ GL ‘extends’ to a permutation of the darts of M mapping the dart (h, hx) to (gh, ghx). If each of the extensions of Ag , g ∈ G, is to preserve the orientation of M, there must exist a cyclic permutation p of X having the property that the dart (g, gp(x)) is always the dart lying immediately next to the dart (g, gx) in the local surface neighborhood of g (sharing the face with (g, gx)), for all g ∈ G and x ∈ X [16]. Thus, we can talk about the local permutation of the darts (g, gx), x ∈ X , emanating from g, determined by the cyclic permutation p of X . This gives rise to an equivalent definition of a Cayley map as an orientable embedding of a Cayley graph C(G,X) in an orientable surface determined by the local rotation scheme with the property that the local surface ordering corresponding to the vertex g ∈ G, ρg , acting cyclically on the darts (g, x), x ∈ X , does not depend on g, and is determined by a fixed cyclic permu- tation p of X . Such Cayley map is denoted by CM(G,X, p), and the equivalence of this ‘local rotation definition’ and the definition via the existence of an orientation preserving automorphism group acting regularly on its vertices is a well established fact in the theory of Cayley maps [16]. As argued above, the group GL = {Ag | g ∈ G} (isomorphic to G) is always a subgroup of the group Aut+CM(G,X, p) of orientation preserving automorphisms, and CM(G,X, p) is orientably regular if and only if there exists an orientation preserving automorphism A ∈ Aut+CM(G,X, p) which fixes the identity 1G and maps the darts emanating from 1G in the same order as p: A((1G, x)) = (1G, p(x)), for all x ∈ X . This has been shown to be equivalent to the existence a special identity-fixing permutation φ called skew-morphism in [7], where it was first defined. Given a group G, a permutation φ : G → G of the elements of G that fixes the identity of G, φ(1G) = 1G, is said to be a skew-morphism of G with associated power function π : G→ N if the equation φ(gh) = φ(g)φπ(g)(h) (3.1) is satisfied for all g, h ∈ G. A Cayley map CM(G,X, p) is orientably regular if and only if there exists a skew-morphism φ of G with the property φ(x) = p(x), for all x ∈ X . Furthermore, a group G admits the existence of an orientably regular Cayley map CM(G,X, p) if and only if G admits the existence of a skew-morphism φ with an orbit X that generates G and is closed under taking inverses (in which case, the desired orientably regular map is the map CM(G,X,φ|X)) [7]. The classification of finite groups G that admit the existence of an orientably regular Cayley map CM(G,X, p) is a hard 390 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 problem and the topic of many articles. The orientation preserving automorphism group of an orientably regular Cayley map with skew-morphism φ and power function π takes the form Aut+(CM(G,X, p)) ∼= GL ⟨φ⟩ with the product multiplication defined by the rule aφ = φπ(a)φ(a), for all a ∈ G. A Cayley map CM(G,X, p) with the property p(x−1) = (p(x))−1 satisfied by all x ∈ X is called a balanced Cayley map. Since a map CM(G,X, p) is balanced if and only GL is normal in Aut+(CM(G,X, p)), we will break with the long line of tradition, and we will call such map a normal Cayley map. This is in line with the name given to Cayley graphs C(G,X) with GL normal in AutC(G,X). Let us complete this section with a necessary condition for a Cayley map to be regular. Even though it can be deduced from [5] and [7], we are not aware of this condition being stated explicitly before, and so we also provide a short proof. Theorem 3.1. If a Cayley map M = CM(G,X, p) is regular, there exists a pair of skew- morphisms φ,ψ of G that preserve X and the restriction of φ to X is equal to p, while the restriction of ψ to X is equal to p−1. Proof. If M = CM(G,X, p) is regular, it is also orientably regular, hence there exists a skew-morphism φ of G that preserves X and whose restriction to X is equal to p [7]. Moreover, if M is regular, it is isomorphic to its mirror reflection CM(G,X, p−1), which is therefore also regular, hence orientably regular, and there exists a skew-morphism ψ of G that preserves X and whose restriction to X is equal to p−1. It is interesting to point out that the above necessary condition is not sufficient. For example, the skew-morphism whose existence guarantees the orientable regularity of a normal Cayley map CM(G,X, p) is well-known to be a group automorphism of G [21]. The inverse of a group automorphism is always a group automorphism, and hence a skew- morphism. Thus, every orientably regular normal Cayley map admits a skew-morphism whose restriction to X is p and whose inverse is a skew-morphism (and its restriction to X is p−1). However, not every normal orientably regular Cayley map is regular. To mention just one famous example, all orientably regular embeddings of complete graphs have been shown to be normal Cayley maps, however, the only orientably regular embeddings of complete graphs which are also regular are the orientable embeddings of complete graphs of prime power order not exceeding 4 [8]. 4 Petrie dual The well-known duality operation switching the roles of vertices and faces of a map M preserves some of the most important topological characteristics of M, such as the ori- entability and the genus, but generally changes the underlying graph of M. The less well- known Petrie duality has the advantage of preserving both the embedded graph and the action of Aut(M) on it. This makes the Petrie dual operation more useful when dealing with maps with automorphism groups acting regularly on vertices of the underlying graph. The Petrie dual of a map M is the map P (M) with the same vertices and edges as M (and thus, the same underlying graph), however, the faces of P (M) are determined by the Petrie (or zig-zag) walks of M which visit vertices of M along the edges while switching sides (i.e., faces) in the middle of every next edge situated along the boundary of the face R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 391 of M. The key point of our interest in the Petrie dual lies in the fact that the two maps M and P (M) have the same vertex set and the same automorphism group. To put this claim on a more precise footing, we turn again to the monodromy groups. If ⟨r0, r1, r2⟩ is the monodromy group of a map M, the monodromy group of the Petrie dual P (M) is the group ⟨r0r2, r1, r2⟩ (which can be easily seen from the fact that the two flags associated with the opposite dart get swapped in the Petrie dual construction). Thus, P (P (M)) = M, and, since ⟨r0, r1, r2⟩ = ⟨r0r2, r1, r2⟩, we obtain Aut(M) = CSF (⟨r0, r1, r2⟩) = CSF (⟨r0r2, r1, r2⟩) = AutP (M). Therefore, the two automorphism groups act on the same set of flags in exactly the same way, and any properties of the automorphism group of M with regard to the vertices of M are shared by the automorphism group of P (M). In particular, if Aut(M) contains a subgroup acting regularly on the set of vertices of M, so does the group AutP (M). Since we make repeated use of this observation throughout our paper, we state it in the form of a theorem. Theorem 4.1. A map M is a generalized Cayley map if and only if its Petrie dual P (M) is also a generalized Cayley map. Even though the Petrie dual pair shares with the original map the same underlying graph and the same automorphism group, the two maps have usually quite distinct topological properties. It is, for example, possible (even common) that M is orientable while P (M) need not be. For an example, consider M to be the tetrahedron as an embedding of K4 on the sphere. Then P (M) is an embedding of K4 in the projective plane as a map of type {4, 3}. Since we will have to be able to distinguish between various situations with regard to the orientablity vs. non-orientability, we recall the following well-known result (e.g., [15, Remark 7]). Theorem 4.2. If M is orientable, then P (M) is orientable if and only if the underlying graph of M is bipartite. Next, recall the concept of a ribbon graph which is constructed from a map M by cutting out an open neighborhood of each face-center and keeping small bands around the edges of M and small circles around the vertices. The ribbon graph of the Petrie dual of an orientable embedding of a bipartite graph is the ribbon graph of the original embedding with all bands twisted (cut off at one of the end vertices and glued back after being rotated by 180 degrees) [4]. Remark 4.3. There is a large number of examples of Cayley maps throughout the litera- ture. For example, four of the five Platonic solids (all but the dodecahedron) are Cayley maps. All orientably regular embeddings of complete graphs are also Cayley maps [8]. Since none of the complete graphs butK2 are bipartite, their Petrie duals are non-orientable generalized Cayley maps. On the other hand, the underlying graph of the dodecahedron is not a Cayley graph, and hence the dodecahedron and its Petrie dual are neither Cayley maps nor generalized Cayley maps. 392 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 5 Orientable generalized Cayley maps and their Petrie duals Let us begin by pointing out that the underlying graph of any generalized Cayley map (orientable or not) is a Cayley graph. This observation follows from the same line of arguments as that included in our section on Cayley maps, namely, it follows from the fact that the group of automorphisms of the underlying graph induced by the group G of map automorphisms acting regularly on the vertices of the map acts regularly on the vertices of the graph. Thus, as is well-known, the set of vertices of the underlying graph of a generalized Cayley map can be identified with the elements of G, and the action of these automorphisms on the elements of G is that of left-multiplication in G. It follows that generalized Cayley maps M are precisely embeddings of Cayley graphs C(G,X) into (orientable or non-orientable) surfaces satisfying the property that all left multiplications in G viewed as permutations of the elements of G ‘extend’ into automor- phisms of the embedding, and that the order of the automorphism group of M acting regularly on the vertices of M must be equal to its number of vertices, i.e., must be equal to |G|. Thus, orientable generalized Cayley maps come in two kinds. First, there are the em- beddings M of Cayley graphs C(G,X) in orientable surfaces having the property that all left multiplications in G extend into orientation preserving automorphisms of M, which are precisely the classical Cayley maps M = CM(G,X, p) [16]. An orientable general- ized Cayley map M that is not of this kind, i.e., that is not a Cayley map, must then be an embedding of a Cayley graph C(G,X) in an orientable surface having the property that at least one of the left multiplications in G extends to an orientation-reversing automorphism of M. Recall that any automorphism group of an orientable map that contains at least one orientation reversing automorphism must contain an index 2 subgroup of orientation preserving automorphisms. This means that orientable generalized Cayley maps which are not Cayley maps are embeddings of Cayley graphs C(G,X) in which a subgroup HL of GL of index 2 extends into orientation preserving automorphisms and the other coset of this subgroup in GL extends into orientation reversing automorphisms. Recall that an orientable embedding of C(G,X) is determined by choosing a cyclic permutation ρg of the elements in X for each g ∈ G. In the case of Cayley maps CM(G,X, p), for every g ∈ G, the permutation ρg of the elements of X must be equal to p. This, of course, cannot be the case for orientable generalized Cayley maps which are not Cayley maps, i.e., there must exist elements f, g ∈ G such that ρf ̸= ρg . However, the distribution of local rotations in orientable generalized Cayley maps which are not Cayley maps is only slightly more complicated than that of Cayley maps. Namely, all such maps are still determined by a single cyclic permutation p of X which becomes the local rotation for the vertices contained in a subgroup H of index 2 in G and whose inverse p−1 becomes the local rotation for the rest of them. We shall denote these maps by GCM(G,H,X, p) and note that each such map is an orientable embedding of a Cayley graph C(G,X) having the properties that H is a subgroup of index 2 in G, p is a cyclic permutation of X , and the embedding of C(G,X) is determined by the local rotations ρh = p, for all h ∈ H , and ρk = p −1, for all k ∈ G \H . Using the notation introduced above, we classify in the following theorem orientable generalized Cayley maps in terms of their local orientations. Theorem 5.1. Let M be an orientable generalized Cayley map. Then M is either a Cayley map, M = CM(G,X, p), or a map M = GCM(G,H,X, p) defined above. R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 393 Conversely, everyCM(G,H,X, p), i.e., every orientable embedding of a Cayley graph C(G,X) having the property that ρh = p, for all h ∈ H , whereH is a subgroup of index 2 in G and p is a cyclic permutation of X , and ρk = p−1, for all k ∈ G \H , is an orientable generalized Cayley map. Finally, the maps GCM(G,H,X, p) and GCM(G,H,X, p−1) are isomorphic. Proof. Let M be an orientable generalized Cayley map that is not a Cayley map, let G ≤ Aut(M) act regularly on the vertices of M , and identify the vertices of M with the elements of G. Let H be the subgroup of index 2 in G of automorphisms preserv- ing the orientation in M whose existence has been argued prior to the statement of this theorem. First, let h be an element of H . This means that the associated map automorphism Ah mapping the darts (g, x), g ∈ G, x ∈ X , to the darts (hg, x) preserves the orientation in M. It follows that Ahρ(g, x) = ρAh(g, x), for all g ∈ G and x ∈ X , and thus (hg, ρg(x)) = (hg, ρhg(x)), for all g ∈ G and x ∈ X . Since multiplication by h preserves H and its coset, the above identities yield that ρh = ρh′ , for all h, h′ ∈ H , and ρk = ρk′ , for all k, k′ ∈ G \H . Taking next an element k ∈ G \H , its corresponding automorphism Ak reverses the orientation in M, and hence satisfies the identity Akρ = ρ−1Ak. Thus, (kg, ρg(x)) = (kg, ρ −1 kg (x)), for all g ∈ G and x ∈ X . It follows that ρg = ρ −1 kg , for all g ∈ G. If we denote the cyclic permutation of X assigned to the elements of H by p, using the fact that the multiplication by k swaps H and its coset yields that the elements in G\H are assigned the local permutation p−1. This completes the proof of the first statement of our theorem. The proof of the second part of the theorem is essentially the reverse of the above. If one chooses the local orientations as described in the theorem, it is easy to see that left multiplication by the elements of H preserves and left multiplication by the elements from G \H reverses the orientation of the obtained map. The veracity of the final statement of our theorem can be verified by showing that the mapping φk defined on the set of darts of GCM(G,H,X, p) by mapping the dart (g, gx) to the dart (kg, kgx) in GCM(G,H,X, p−1), for any fixed k ∈ G \ H and all g ∈ G and x ∈ X , is a map isomorphism between the maps GCM(G,H,X, p) and GCM(G,H,X, p−1). This relatively simple task is left to the reader. To obtain examples of orientable generalized Cayley maps that are not Cayley maps, one needs to consider embeddings of Cayley graphs of even order. To obtain an ‘easy’ example, one could consider the generalized Cayley mapGCM(G,H,X, p) for the groups G = Z2n, n ≥ 1, H = ⟨2⟩, X = G \ {0}, and any cyclic permutation p of X . All of these maps are embeddings of the complete graphs K2n in orientable surfaces. However, it is easy to see that the cases n = 1 and n = 2 result in Cayley maps, and even for n ≥ 3, it is hard to show that the resulting maps are not Cayley. Thus, to construct an infinite family of orientable generalized Cayley maps which are provably not Cayley maps, one might rely on the Petrie dual operator again. Specifically, in what follows, we shall consider orientable generalized Cayley maps whose Petrie dual is orientable again. Because of Theorem 4.2, each such map must be bipartite (which is also a sufficient condition). Thus, we obtain the following: Theorem 5.2. Let M be an orientable generalized Cayley map. The Petrie dual of M is orientable if and only if M = CM(G,X, p) is a Cayley map for a group G that contains 394 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 a subgroup H of index 2 for which X ⊆ G \H or if M = GCM(G,H,X, p) where H is a subgroup of G of index 2 for which X ⊆ G \H . Furthermore, if the Petrie dual of an orientable generalized Cayley map M is ori- entable, the two maps CM(G,X, p) and GCM(G,H,X, p) are mutual Petrie duals. Proof. In order for a Cayley graph C(G,X) to be bipartite, it is easy to see (and well known) that it must be a Cayley graph of an even order group G that possesses a subgroup H of index 2, and X must be a subset of the non-trivial coset of H in G. The rest of the first part of the theorem follows from Theorem 5.1. Suppose now that M = CM(G,X, p) with X satisfying the required condition with respect to a subgroup H of G. It is again not too hard to see that the Petrie dual to M reverses the order of elements fromX for each element ofH (or for each element ofG\H; the two maps GCM(G,H,X, p) and GCM(G,H,X, p−1) are isomorphic as argued in Theorem 5.1). Clearly, an orientable generalized Cayley map M = GCM(G,H,X, p) is not simulta- neously a Cayley map if and only if the group Aut(M) contains no orientation preserving subgroup acting regularly on the vertices of M. One way to make sure no such group exists, is to find a pair of vertices u, v of M for which there is no orientation preserving automorphism mapping u to v. This is the approach we take to construct an infinite family of orientable generalized Cayley maps which are not Cayley maps. Lemma 5.3. Let G be a finite group with a subgroup H of index 2 and a generating set X ⊆ G \H . If CM(G,X, p) is chiral, its Petrie dual GCM(G,H,X, p) is a generalized Cayley map that is not a Cayley map. Proof. Recall that an orientably regular map is chiral if it admits no orientation reversing automorphisms. Note also, that since CM(G,X, p) is bipartite and connected, any auto- morphism φ ∈ AutCM(G,X, p) that maps a vertex u ∈ H to a vertex v ∈ G \H maps all elements of H onto the elements of G \ H . This means that all the automorphisms in AutGCM(G,H,X, p) mapping elements of H to elements in G \ H are orientation reversing, and no orientation preserving automorphisms of GCM(G,H,X, p) map the el- ements of H to the elements in G \H . It follows that GCM(G,H,X, p) is not a Cayley map. Remark 5.4. There are many examples of chiral bipartite Cayley maps. One good source of such maps is the paper [10] the authors of which construct infinitely many orientably regular embeddings of Kn,n, with n = pe where p is an odd prime and e ≥ 1, which are Cayley maps for cyclic and dihedral groups and are chiral. Lemma 5.3 yields that all their Petrie duals are generalized Cayley maps that are not Cayley maps. A simpler family of chiral bipartite Cayley maps, brought to our attention by one of our referees, is also formed by the torus maps {4, 4}b,c, for b, c both non-zero, b + c even, and b ̸= c to guarantee chirality, and the torus maps {6, 3}b,c. The chiral bipartite {4, 4}b,c maps have orders b2 + c2 and arise by identifying opposite sides of squares with corners (0, 0), (b, c), (b−c, b+c) and (−c, b) in a unit rectangular grid. Example 5.5. Consider the dihedral groups Dn = 〈 a, b | an = b2 = 1, bab−1 = a−1 〉 , R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 395 n ≥ 3. Taking G = Dn, H = ⟨a⟩ and X = {b, ba, . . . , ban−1} yields a pair of mutu- ally Petrie dual generalized Cayley maps CM(G,X, p) and GCM(G,H,X, p), for every cyclic permutation p of X . All such maps are orientable embeddings of the complete bi- partite graph Kn,n. Taking the specific permutation p = (b, ba, . . . , ban−1) yields two (possibly isomorphic) embeddings of Kn,n. First, the Cayley map CM(Dn, {b, ba, . . . , ban−1}, (b, ba, . . . , ban−1)) is an orientable embedding of Kn,n with n faces of length 2n and of genus (n−1)(n−2) 2 , while its Petrie dual GCM(Dn, ⟨a⟩ , {b, ba, . . . , ban−1}, (b, ba, . . . , ban−1)) is an orientable embedding of Kn,n with the same number of faces of the same length and of the same genus. The excluded case n = 2 results in an embedding of K2,2 in the sphere of the form CM(Z22, {(0, 1), (1, 1)}, ((0, 1), (1, 1))) which is its own Petrie dual because the inverse of the permutation ((0, 1), (1, 1)) is ((0, 1), (1, 1)) again. The toroidal maps for case n = 3 are pictured below. b a2 ba2 a ba 1 b a ba2 a2 ba 1 Figure 1: Maps CM(D3, {b, ba, ba2}, (b, ba, ba2)) and GCM(D3, ⟨a⟩ , {b, ba, ba2}, (b, ba, ba2)). Clearly, the two maps pictured in Figure 1 are isomorphic. Hence, the smallest maps CM(D3, {b, ba, ba2}, (b, ba, ba2)) and CM(Z22, {(0, 1), (1, 1)}, ((0, 1), (1, 1)))) are self- Petrie-dual. This might come as a surprise, as the same permutations of vertices (com- ing from left multiplications) cannot simultaneously extend into orientation preserving and orientation reversing automorphisms. It is, however, easy to notice that relabeling the vertices of CM(D3, {b, ba, ba2}, (b, ba, ba2)) changes the permutation actions of left- multiplications, and hence left multiplications with respect to one labeling may extend to orientation preserving while left multiplications with respect to another labeling may ex- tend to orientation reversing automorphisms. Since in non-bipartite cases the Petrie dual of an orientable map is non-orientable, the case of bipartite orientable maps is the only case where one may encounter orientable self-Petrie-dual generalized Cayley maps. 396 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 We devote the rest of this section to the investigation of such possibility, i.e., to the study of orientable self-Petrie-dual generalized Cayley maps. Namely, suppose an orientable generalized Cayley map M is self-Petrie-dual. Then, M is bipartite and Theorem 5.1 im- plies that the automorphism group of M contains two subgroups acting regularly on the vertices, a subgroup of orientation preserving automorphisms as well as a subgroup of au- tomorphisms half of which is orientation reversing. This yields a necessary condition for M being self-Petrie-dual, namely, M must be a Cayley map admitting at least one orienta- tion reversing automorphism. Thus, chiral bipartite Cayley maps are never self-Petrie-dual (in Example 5.4, we have already encountered infinitely many such maps constructed in [10]). In fact, it is well-known that no chiral maps are self-Petrie-dual. These observations appear to suggest that in order to construct bipartite Cayley maps that are self-Petrie-dual, one should consider maps with many orientation preserving and many orientation reversing automorphisms. In fact, searching through the literature for orientable self-Petrie-dual maps resulted only in regular orientable self-Petrie-dual maps [2, 11, 17], with the methods used in [11] or [17] relying on regularity and producing large maps of large genera. An infinite family of regular examples also arises as follows. Remark 5.6. For every positive integer n ≥ 2 that is relatively prime to φ(n) (i.e., there exists a unique group of order n), there exists exactly one orientably regular embedding of Kn,n [9]. That means that such embedding is necessarily isomorphic to its mirror re- flection, hence regular, as well as self-Petrie dual. Kwak and Kwon in [13] classified the orientable regular self-Petrie embeddings of Kn,n. Nevertheless, even in the extreme case when M is regular, it might happen that any orientation preserving subgroup K ≤ Aut(M) of order half the number of vertices of M paired with any orientation reversing automorphism ψ ∈ Aut(M) generate a group of order larger than the number of vertices of M. In that case, Aut(M) does not contain a subgroup of automorphisms containing an orientation reversing automorphism and acting regularly on the vertices. Hence, M cannot be isomorphic to any GCM(G,H,X, p), and hence cannot be self-Petrie-dual. To illustrate this possibility, in the following example we present an infinite family of bipartite regular Cayley maps none of which are self-Petrie- dual. We are thankful to Gareth Jones who pointed out this example to us. Remark 5.7. The authors of [3] considered regular embeddings of Kn,n having the prop- erty that the orientation preserving automorphism group of the embedding that does not move the bipartite sets is not metacyclic; in which case n is necessarily a power of 2. For each such n ≥ 8, they construct four non-isomorphic regular orientable embeddings of Kn,n, denoted N (n; k, l), k, l ∈ {0, 1}. They prove that the maps N (n; 0, 0) and N (n; 1, 0) are self-Petrie-dual, while N (n; 0, 1) and N (n; 1, 1) are Petrie duals of each other. They also show that all four maps are Cayley maps. This means, in particular, that each of the maps N (n; 0, 1) and N (n; 1, 1) is both a Cayley map and an orientable gener- alized Cayley map, and as such, they both admit both an orientation preserving as well as an orientation reversing automorphism group acting regularly on the vertices of the maps. Hence, not even the existence of both types of vertex-regular groups is sufficient for making a generalized Cayley map self-Petrie. It is hard to say whether these non-metacyclic maps are in any way typical, and whether a regular bipartite Cayley map is more likely to be self-Petrie-dual or not. The following highly specialized result appears to be of some relevance toward answering this question: R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 397 Theorem 5.8 ([17]). There are no regular, self-dual, self-Petrie-dual, normal Cayley maps with odd vertex degree. 6 Non-orientable generalized Cayley maps and their Petrie duals Even though constructing non-orientable generalized Cayley maps is relatively easy – it is enough to consider the Petrie dual of any non-bipartite orientable generalized Cayley map – a more explicit construction of such maps is needed. When considering only the Petrie duals of orientable generalized Cayley maps, one would never encounter a non-orientable generalized Cayley map whose Petrie dual is not orientable. To see that such maps might exist, one just needs to realize that the Petrie dual of a non-orientable generalized Cayley map whose underlying graph is bipartite (we will construct such maps) is necessarily non- orientable. The present section differs from the previous section which was concerned with maps whose groups are transitive on arcs (regular and orientably regular). In this section, we abandon the focus on such highly symmetric generalized Cayley maps and seek only to provide examples of various non-orientable generalized Cayley maps. To specify an embedding of a graph in a non-orientable surface, one needs to spec- ify the local orientation of outgoing darts around every vertex as well as to specify for each edge whether the two local orientations associated with its end-vertices are consistent or not. More precisely, for any given edge e incident with vertices u and v, let U be an open neighborhood of e which includes both u and v and all of e, but no other vertices and no other complete edges. Then U is homeomorphic to an open disk, and is therefore an orientable topological object. By calling the two local orientations associated with the end-vertices of e consistent, we mean that the local orientations around u and v within U are both clockwise or both anti-clockwise. We call them inconsistent otherwise. Refer- ring to the ribbon graph associated with the map, the edges with inconsistent end-vertex orientations are sometimes also called twisted, as the strip containing the edge e and the vertices u and v is twisted before being attached to the rest of the ribbon graph through the vertices u and v. To indicate whether the end-vertex orientations of an edge are consistent or not, one usually assigns 1 or −1 to the edge, respectively (or, in a visualization of the graph embedding using local rotations, an edge with inconsistent end-vertex orientations is marked by an ‘x’). Perhaps the biggest disadvantage of this description of an embedding of a graph is the fact that one does not have a guarantee that the resulting embedding is indeed non-orientable. For example, the bipartite orientable maps GCM(G,H,X, p) defined in the previous section can also be described using this ‘non-orientable’ description by saying that GCM(G,H,X, p) is the map in which the local rotation at each vertex is equal to p and all edges have inconsistent end-vertex orientations. As explained in the previous sections, a generalized Cayley map must be an embedding of a Cayley graph C(G,X) with the property that each left multiplication by the elements of G extends into a map automorphism of the embedding. One of the finer points to be dealt with in the previous section was the fact that the extensions might be either orientation preserving or orientation reversing. We do not have two kinds of map automorphisms in non-orientable maps which makes the situation a bit easier. Let us start by describing general embeddings of Cayley graphs. Let C(G,X) be a Cayley graph, let ρg denote the local rotation of the elements of X around g, and let ι{g,gx} ∈ {1,−1} be the label assigned to the edge {g, gx}. The three defining involu- 398 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 tions r0, r1, r2 are then defined as follows: (g, c{g,gx}, cF ) r0 = { (gx, c{g,gx}, cF ), if ι{g,gx} = 1, (gx, c{g,gx}, cF ′), if ι{g,gx} = −1 , (g, c{g,gx}, cF ) r1 = (g, c{g,gρι{g,gx}g (x)} , cF ), and (g, c{g,gx}, cF )r2 = (g, c{g,gx}, cF ′), where F ′ is the face adjacent to F across the edge {g, gx}. The following characterization of generalized Cayley maps can be already found in both [19] and [12]. As stated above, it covers both the orientable and the non-orientable generalized Cayley maps. Theorem 6.1 ([12, 19]). A (orientable or non-orientable) map M is a generalized Cayley map if and only if it is an embedding of a Cayley graph C(G,X) with all local cyclic per- mutations ρg , g ∈ G, equal to a fixed cyclic permutation p of X , and the twist distribution ι satisfying the property ι{g,gx} = ι{g′,g′x}, for all g, g′ ∈ G and x ∈ X . The simplest interpretation of the last condition is that all edges labelled by the same generator (or its inverse) must be either all simultaneously un-twisted or all twisted, and hence, from now on, we will assume that ι acts on the set X , ι : X → {−1, 1}, mapping x ∈ X to ι{g,gx} (thus, in particular, ι(x) = ι(x−1)), for all x ∈ X . As each generalized Cayley map is determined by the four-tuple (G,X, p, ι), we will denote these maps by GCM(G,X, p, ι). The authors of [12] generalized this simplification even further and introduced the fol- lowing convenient notation. Let M = GCM(G,X, p, ι) be a generalized Cayley map. Without loss of generality, we may assume that X = {x0, x1, . . . , xd−1}, while p(xi) = xi+1, for all i ∈ [d] = {0, 1, 2, . . . d−1} (with the addition performed modulo d). If we denote the distribution of inverses in X by the function κ : [d] → [d], satisfying the property (xi)−1 = xκ(i), and use the fact that ι assigns the same value to all edges arising from right mutiplication by xi to define (with just a hint of abuse of the notation) ι : [d] → {−1, 1}, ι(i) = ι{1g,xi} (note that ι(κ(i)) = ι(i), for all i ∈ [d]), we may associate the flags of M with the ordered triples from G × [d] × {−1, 1}, with the neighboring flags (g, c(g,gxi), cF ) and (g, c(g,gxi), cF ′) corresponding to the triples (g, i, 1) and (g, i,−1). The defining involutions take then the particularly simple form: (g, i, j)r0 = (gxi, κ(i), ι(i)j), (g, i, j) r1 = (g, i+ j, j), and (g, i, j)r2 = (g, i,−j). Remark 6.2. It is now easy to construct infinitely many non-orientable generalized Cayley maps. As is well-known (see e.g., [4]), an embedding of a graph determined by local ro- tations and a twisting function is non-orientable if and only if it contains at least one cycle with an odd number of twisted edges. Hence, for example, it is easy to construct infinitely many non-orientable embeddings of the complete bipartite graphs Kn,n. To get such an embedding, one can take a bipartite C(G,X) with G containing a subgroup H of index 2, X = G \H = {x0, x1, . . . , x|G|/2−1}, satisfying the only additional condition that x20 ̸= x−12 x −1 1 , choose any cyclic permutation p of X , twist the edges labeled by x0, and observe that the 4-cycle consisting of the vertices 1G, x0, x0x1, x0x1x2, x0x1x2(x0x1x2)−1 = 1G obtained by successive multiplications by x0, x1, x2 and (x0x1x2)−1 (which, being an odd product of elements from X , necessarily belongs to X) contains the twisted edge labeled R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 399 x0 exactly once (if (x0x1x2)−1 were equal to x0, we would get the identity x−12 x −1 1 = x 2 0 which we explicitely prohibited). Hence, all such generalized Cayley maps GCM(G,X, p, ι) are unorientable embeddings of complete bipartite graphs, and there are clearly in- finitely many of them. Note also that all such maps, being embeddings of bipartite graphs, have the property that their Petrie dual is non-orientable. Next, we bring in the Petrie dual operator again and fully describe its action on gener- alized Cayley maps. The mapping −ι used in the statement of the following theorem is the twisting function −ι(x) = (−1) · ι(x), for all x ∈ X . Theorem 6.3. If M = GCM(G,X, p, ι), then P (M) = GCM(G,X, p,−ι). Proof. It is easy to see that −ι is a correctly defined twisting function of a generalized Cay- ley map, i.e., −ι(x) = −ι(x−1), for all x ∈ X . Since the two maps GCM(G,X, p,−ι) and P (GCM(G,X, p, ι)) have the same underlying graphs, one way to prove our the- orem is to show that the Petrie polygons of GCM(G,X, p, ι) and the faces of the map GCM(G,X, p,−ι) are identical. This can be shown by performing a careful calculation of the boundaries of the two oriented Petrie polygons of GCM(G,X, p, ι) that start at (g, i) (which might turn out to be the same polygon visiting (g, i) twice, once followed along p and once followed along p−1, depending on whether we first turn ‘right’ or ‘left’) and the boundaries of the corresponding faces in GCM(G,X, p,−ι). As pointed out by one of our referees, it can also be deduced from the fact that if r0, r1, r2 are the generators of the monodromy group of GCM(G,X, p, ι), then r0r2, r1, and r2 are the generators for the monodromy group of P (GCM(G,X, p, ι)) (Section 4) while they can also be easily seen to be the generators of the monodromy group of GCM(G,X, p,−ι). Based on the above theorem, obtaining the Petrie dual for a Cayley map is very simple. Namely, the Petrie dual of a Cayley map CM(G,X, p) is the generalized Cayley map GCM(G,X, p, ι), where ι twists all the edges of the map, i.e., ι(x) = −1, for every x ∈ X . Since all the edges are twisted, the Petrie dual of a Cayley map M is non- orientable if and only if M contains at least one odd cycle (i.e., the underlying graph is non-bipartite). Hence, Theorem 5.2 is a corollary of Theorem 6.3. Example 6.4. To obtain a specific example of the use of Theorem 6.3, we go back to a classical example of a pair of mutually Petrie dual maps, namely the regular embedding of the tetrahedron on the sphere and its Petrie dual embedded in the projective plane. Since tetrahedron is the Cayley map CM(Z2 × Z2, {(1, 0), (0, 1), (1, 1)}, ((1, 0), (0, 1), (1, 1))) with 4 triangular faces, its Petrie dual is the generalized Cayley map GCM(Z2 × Z2, {(1, 0), (0, 1), (1, 1)}, ((1, 0), (0, 1), (1, 1)), ι), with ι(x) = −1 for all x ∈ {(1, 0), (0, 1), (1, 1)}. Since the tetrahedron contains odd cycles (triangles), the dual map is non-orientable, well-known to be an embedding in the projective plane with 3 faces of length 4 (which can also be easily verified via direct calcu- lations in the generalized Cayley map). This example can be viewed as the first member of an infinite family of examples. For every even n ≥ 4, let Dn = ⟨a, b | an = b2 = 1, bab = a−1⟩ be the dihedral group of order 2n, X = {b, an2 , ba} be a set of three involutions, and p = (b, an2 , ba). Since n is assumed even, the Petrie dual of the trivalent Cayley map CM(Dn, X, p) is non-orientable (as the underlying graph of both maps contains a cycle of lenght n + 1). For even n ≥ 4, 400 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 the Petrie dual of CM(G,X, p) looks ‘like’ the Petrie dual of the tetrahedron: a 2n-cycle with inside chords at every other vertex of the cycle, and an outside edge for the rest. All these Petrie duals are embeddings in the projective plane with one face of length 2n and all the remaining faces of length 4. 1 ba b a a3 ba2 a2 ba3 Figure 2: Representation of the Petrie dual of the map CM(D4, {b, a2, ba}, (b, a2, ba)) indicating local rotations and marking the edges with inconsistent end-vertex rotations of its underlying graph. Next, we characterize non-orientable generalized Cayley maps whose Petrie dual is orientable. As should be expected, those are the ‘obvious’ Petrie duals of non-bipartite orientable Cayley maps, i.e., the non-bipartite non-orientable generalized Cayley maps sat- isfying ι(x) = −1, for all x ∈ X , as well as other maps with a rather special twisting function. The Petrie duals of all other non-orientable generalized Cayley maps are non- orientable. Corollary 6.5. The Petrie dual of a non-orientable generalized Cayley map GCM(G,X, p, ι) is orientable if and only if the underlying Cayley graph C(G,X) contains an odd cycle and every cycle of C(G,X) contains an even number of untwisted edges. Proof. Suppose that M = GCM(G,X, p, ι) is non-orientable while its Petrie dual P (M) = GCM(G,X, p,−ι) is orientable. This means that M cannot be bipartite as that would make the orientable P (M) bipartite and hence by Theorem 5.2 its Petrie dual P (P (M)) = M orientable; which it is not. Thus M contains an odd cycle. If any cycle in M contained an odd number of untwisted edges, that very same cycle would contain an odd number of twisted edges in P (M); making it non-orientable. This proves one implication. Suppose now that M is non-orientable, contains an odd length cycle and every cycle of M contains an even number of untwisted edges. Then M is not bipartite and its Petrie dual contains no cycles with an odd number of twisted edges. As stated prior to Theorem 6.3, non-orientable non-bipartite generalized Cayley maps with all edges twisted trivially satisfy the property that their Petrie dual is also non- orientable. Next, we provide an infinite family of examples that satisfy the conditions of Theorem 6.3 but contain both twisted and untwisted edges. R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 401 Example 6.6. Let Γ be a bipartite d-regular graph, d-odd, let M be any orientable em- bedding of Γ, and ρ and λ be the generators for the monodromy group G of M. Con- sider the Cayley graph C(G, {ρ, ρ−1, λ}). There is an easy way to visualize this graph. It is the underlying graph of the truncation T (M) of M, i.e., the map obtained from M by locally removing its vertices and replacing them by d-cycles attached to the dan- gling edges in the order determined by the local rotation. We claim that every cycle of C(G, {ρ, ρ−1, λ}) contains an even number of edges labeled λ. To see this, take any cycle C of C(G, {ρ, ρ−1, λ}). The edges labeled λ trace the ‘preimage’ of this cycle in Γ, which is necessarily a union of edge disjoint even cycles. Since every edge labeled λ in C corre- sponds to exactly one edge of the preimage, the number of edges labeled λ in C is even. Thus, any choice of p together with the twisting function ι(ρ) = ι(ρ−1) = −1, ι(λ) = 1 makes the non-orientable GCM(G,X, p, ι) satisfy the conditions of Theorem 6.3. The visualization of this example is in fact quite simple. We start with a ‘bipartite’ map whose ‘vertices’ are the d-cycles formed by edges labelled ρ, all of which are twisted, with the edges labelled λ and connecting the two sets of d-cycles untwisted. This map is non-orientable, since d is required to be odd. The orientable dual untwists the d-cycles and twists the connecting edges. To conclude this section, we consider one more classical concept from topological graph theory – the orientable double covering Mo of a non-orientable map M given by its underlying graph, rotation system and twisting function. The underlying graph of the dou- ble covering is the Z2-lift of the underlying graph of the original map with the untwisted edges given the voltage 0 ∈ Z2, and the twisted edges receiving the voltage 1 ∈ Z2. Infor- mally, it is the double cover of the underlying graph of the original map with the untwisted edges each lifted into two edges connecting vertices in the same layer, and the two lifts of the twisted edges crossing from one layer to the other (for more details consult [20]). The- orem 2.4 in [20] asserts that Aut(M) lifts into the orientation preserving automorphism group Aut+Mo, and the full group Aut(M)o is the direct product of Aut+Mo with some orientation reversing involutory automorphism ψ of order 2 (that swaps the two layers of the double covering). It follows that the direct product of the lift of a subgroup of Aut(M) acting regularly on the vertices of M with the group ⟨ψ⟩ acts regularly on the vertices of Mo, and we obtain: Theorem 6.7. The orientable double covering Mo of any non-orientable generalized Cay- ley map M is an orientable generalized Cayley map admitting a group of automorphisms containing an orientation reversing automorphism and acting regularly on its vertices. 7 Regular generalized Cayley maps In conclusion of our paper, we would like to direct the reader toward [12] where the authors of that paper developed a complete theory of regular generalized Cayley maps including necessary and sufficient conditions based on the existence of special permutations of the elements of the underlying group. Their language, however, is different from the one used here, and they do not consider the orientable generalized Cayley maps GCM(G,H,X, p) as a special case. ORCID iDs Robert Jajcay https://orcid.org/0000-0002-2166-2092 402 Ars Math. Contemp. 24 (2024) #P3.01 / 385–403 Jozef Širáň https://orcid.org/0000-0002-5901-7646 Yan Wang https://orcid.org/0000-0002-0148-2932 References [1] M. Abas, personal communication. [2] G. Cunningham, Self-dual, self-Petrie covers of regular polyhedra, Symmetry 4 (2012), 208– 218, doi:10.3390/sym4010208, https://doi.org/10.3390/sym4010208. [3] S.-F. Du, G. Jones, J. H. Kwak, R. Nedela and M. Skoviera, Regular embeddings of Kn,n where n is a power of 2. II: the non-metacyclic case, Eur. J. Comb. 31 (2010), 1946–1956, doi:10.1016/j.ejc.2010.01.009, https://doi.org/10.1016/j.ejc.2010.01.009. [4] J. L. Gross and T. W. Tucker, Topological Graph Theory, Dover Publications, Inc., Mineola, NY, 2001, reprint of the 1987 original (Wiley, New York). [5] R. Jajcay, Automorphism groups of Cayley maps, J. Comb. Theory Ser. B 59 (1993), 297–310, doi:10.1006/jctb.1993.1071, https://doi.org/10.1006/jctb.1993.1071. [6] R. Jajcay, C.-H. Li, J. Širáň and Y. Wang, Regular and orientably-regular maps with quasiprim- itive automorphism groups on vertices, Geom. Dedicata 203 (2019), 389–418, doi:10.1007/ s10711-019-00440-6, https://doi.org/10.1007/s10711-019-00440-6. [7] 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. [8] L. D. James and G. A. Jones, Regular orientable imbeddings of complete graphs, J. Comb. The- ory Ser. B 39 (1985), 353–367, doi:10.1016/0095-8956(85)90060-7, https://doi.org/ 10.1016/0095-8956(85)90060-7. [9] G. A. Jones, Regular embeddings of complete bipartite graphs: classification and enumeration, Proc. Lond. Math. Soc. (3) 101 (2010), 427–453, doi:10.1112/plms/pdp061, https://doi. org/10.1112/plms/pdp061. [10] G. A. Jones, R. Nedela and M. Škoviera, Regular embeddings of Kn,n where n is an odd prime power, Eur. J. Comb. 28 (2007), 1863–1875, doi:10.1016/j.ejc.2005.07.021, https: //doi.org/10.1016/j.ejc.2005.07.021. [11] G. A. Jones and M. Ziv-Av, Petrie duality and the Anstee-Robertson graph, Symmetry 7 (2015), 2206–2223, doi:10.3390/sym7042206, https://doi.org/10.3390/sym7042206. [12] J. H. Kwak and Y. S. Kwon, Unoriented Cayley maps, Studia Sci. Math. Hungarica 43 (2006), 137–157, doi:10.1556/SScMath.43.2006.2.1, https://doi.org/10.1556/SScMath. 43.2006.2.1. [13] J. H. Kwak and Y. S. Kwon, Classification of reflexible regular embeddings and self-Petrie dual regular embeddings of complete bipartite graphs, Discrete Math. 308 (2008), 2156–2166, doi: 10.1016/j.disc.2007.04.062, https://doi.org/10.1016/j.disc.2007.04.062. [14] C. H. Li and J. Širáň, Regular maps whose groups do not act faithfully on vertices, edges, or faces, Eur. J. Comb. 26 (2005), 521–541, doi:10.1016/j.ejc.2003.09.022, https://doi. org/10.1016/j.ejc.2003.09.022. [15] R. Nedela and M. Škoviera, Maps, in: J. Gross, J. Yellen and P. Zhang (eds.), Handbook of Graph Theory, 2nd ed., CRC Press, Boca Raton, FL, pp. 820–859, 2013, doi:10.1201/b16132, https://doi.org/10.1201/b16132. [16] R. B. Richter, J. Širáň, R. Jajcay, T. W. Tucker and M. E. Watkins, Cayley maps, J. Comb. The- ory Ser. B 95 (2005), 189–245, doi:10.1016/j.jctb.2005.04.007, https://doi.org/10. 1016/j.jctb.2005.04.007. R. Jajcay et al.: Generalized Cayley maps and their Petrie duals 403 [17] R. B. Richter, J. Širáň and Y. Wang, Self-dual and self-Petrie-dual regular maps, J. Graph Theory 69 (2012), 152–159, doi:10.1002/jgt.20570, https://doi.org/10.1002/jgt. 20570. [18] G. Sabidussi, Vertex-transitive graphs, Monatsh. Math. 68 (1964), 426–438, doi:10.1007/ BF01304186, https://doi.org/10.1007/BF01304186. [19] T. W. Tucker, Symmetric embeddings of Cayley graphs in nonorientable surfaces, in: Graph Theory, Combinatorics, and Applications. Vol. 2 (Kalamazoo, MI, 1988), Wiley, New York, Wiley-Intersci. Publ., pp. 1105–1120, 1991. [20] J. Širáň and T. W. Tucker, Symmetric maps, in: L. W. Beineke and R. J. Wilson (eds.), Topics in Topological Graph Theory, Cambridge University Press, Cambridge, volume 128 of En- cyclopedia Math. Appl., pp. 199–224, 2009, doi:10.1017/CBO9781139087223.013, https: //doi.org/10.1017/CBO9781139087223.013. [21] M. Škoviera and J. Širáň, Regular maps from Cayley graphs. I: Balanced Cayley maps, Dis- crete Math. 109 (1992), 265–276, doi:10.1016/0012-365X(92)90296-R, https://doi. org/10.1016/0012-365X(92)90296-R. [22] S. E. Wilson, Operators over regular maps, Pacific J. Math. 81 (1979), 559–568, http:// projecteuclid.org/euclid.pjm/1102785296. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.02 / 405–414 https://doi.org/10.26493/1855-3974.2570.5e8 (Also available at http://amc-journal.eu) On p-gonal fields of definition* Ruben A. Hidalgo † Departamento de Matemática y Estadı́stica, Universidad de La Frontera, Temuco, Chile Received 24 February 2021, accepted 28 November 2023, published online 11 June 2024 Abstract Let S be a closed Riemann surface of genus g ≥ 2 and φ be a conformal automorphism of S of prime order p such that S/⟨φ⟩ has genus zero. Let K ≤ C be a field of definition of S. We prove the existence of a field extension F of K, of degree at most 2(p−1), for which S is definable by a curve of the form yp = F (x) ∈ F[x], in which case φ corresponds to (x, y) 7→ (x, e2πi/py). If, moreover, φ is also definable over K, then F can be chosen to be at most a quadratic extension of K. For p = 2, that is when S is hyperelliptic and φ is its hyperelliptic involution, this fact is due to Mestre (for even genus) and Huggins and Lercier-Ritzenthaler-Sijslingit in the case that Aut(S)/⟨φ⟩ is non-trivial. Keywords: Riemann surfaces, p-gonal curves, automorphisms. Math. Subj. Class. (2020): 30F10, 30F20, 14H37, 14H55 1 Introduction In [23], H. A. Schwarz proved that the group Aut(S) of conformal automorphisms of a closed Riemann surface S of genus g ≥ 2 is finite. Later, in [17], A. Hurwitz obtained the upper bound |Aut(S)| ≤ 84(g − 1) (this is known as the Hurwitz’s bound). Let p ≥ 2 be a prime integer. We say that a closed Riemann surface S is cyclic p- gonal if there exists some φ ∈ Aut(S) of order p such that the quotient orbifold S/⟨φ⟩ has genus zero. In this case, φ is called a p-gonal automorphism and the cyclic group ⟨φ⟩ a p-gonal group of S. The case p = 2 corresponds to S being hyperelliptic and φ its (unique) hyperelliptic involution. The case p = 3 was studied by R. D. M. Accola in [1]. In [10], G. González-Diez proved that p-gonal groups are unique up to conjugation in Aut(S). In [13], it was observed that, if p ≥ 5n − 7, where n ≥ 3 is the number of fixed *The author would like to express their gratitude to both referees for their valuable feedback, suggestions, and corrections that have significantly improved the paper. †Partially supported by projects Fondecyt 1230001 and 1220261. E-mail address: ruben.hidalgo@ufrontera.cl (Ruben A. Hidalgo) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 406 Ars Math. Contemp. 24 (2024) #P3.02 / 405–414 points of φ, then ⟨φ⟩ is the unique p-group in Aut(S). Results concerning automorphisms of p-gonal Riemann surfaces can be found, for instance, in [2, 3, 4, 5, 11, 27]. As a consequence of the Riemann-Roch theorem, a closed Riemann surface S can be described by an (either affine or projective) irreducible complex algebraic curve, i.e., after desingularization (if it is non-smooth) and filling at some punctures in the affine case, it carries a Riemann surface structure which is biholomorphic to that of S (see Remark 1.1 for the case of cyclic p-gonal surfaces). A subfield K of the field C of complex numbers is called a field of definition of S (or that S is definable over K) if there is an irreducible al- gebraic curve representing S, which is defined as the common zeroes of some polynomials with coefficients in K. The intersection of all the fields of definition of S is called the field of moduli of S. In general, it is not a field of definition (see Section 3). If we are given a (finite) group G < Aut(S) and the geometrical structure of the quotient orbifold S/G, then it is not a simple task to find an algebraic curve for S reflecting the action of G. A family of surfaces for which algebraic models are well known is the case of cyclic p-gonal Riemann surfaces, which we proceed to recall below. Let S be a p-gonal Riemann surface, φ ∈ Aut(S) be a p-gonal automorphism and π : S → Ĉ be a regular branched cover with ⟨φ⟩ its deck group. Let {a1, . . . , am} ⊂ Ĉ = C ∪ {∞} be the set of branch values of π. If aj ̸= ∞, for every j = 1, . . . ,m, then there exist integers n1, ..., nm ∈ {1, ..., p−1}, n1+ · · ·+nm ≡ 0 mod p, such that S is defined by the affine, irreducible and smooth p-gonal curve with equation E : yp = F (x) = m∏ j=1 (x− aj)nj ∈ C[x]. (1.1) If one of the branch values is equal to ∞, say am = ∞, then in (1.1) we delete the corresponding factor (x − am)nm and assume n1 + · · · + nm−1 ̸≡ 0 mod p. In this affine algebraic model, π(x, y) = x and φ(x, y) = (x, ωpy), where ωp = e2πi/p. In the hyperelliptic case, i.e., p = 2, in the above one has m ∈ {2g + 1, 2g + 2} and nj = 1. Remark 1.1. The affine curve (1.1) is smooth at those points (x, y), where y ̸= 0. At a point (aj , 0), the curve is smooth exactly when nj = 1 (anyway, if nj > 1, it has a neighborhood homeomorphic to a disc). An irreducible projective algebraic curve defining S is obtained from the above affine one as Ê : ypzn1+···+nm−p = m∏ j=1 (x− ajz)nj . (1.2) As in the affine model, the projective curve Ê is smooth at the points [x : y : 1], where y ̸= 0. At the points [aj : 0 : 1] it is smooth if and only if nj = 1 (again, in the other cases there is a neighborhood homeomorphic to a disc). The curve is also non-smooth at the point [0 : 1 : 0]. After normalization of the curve, one obtains a closed Riemann surface which is biholomorphic to S. In this case, π([x : y : z]) = x/z and φ([x : y : z]) = [x : ωpy : z]. If F is a subfield of C such that in (1.1) we have F (x) ∈ F[x], then we say that F is a p-gonal field of definition of S (and that S is cyclically p-gonally defined over F). Note that there are infinitely many different p-gonal fields of definition for S (for instance, if T is a Möbius transformation, then we may replace the values aj by T (aj)). Given a field of definition of a p-gonal Riemann surface S, it is not clear at first sight if it is a p-gonal field of definition. Also, it might be that a minimal p-gonal field of definition R. A. Hidalgo: On p-gonal fields of definition 407 is not a minimal field of definition (see the exceptional case (m, p) = (4, 3) in Section 4.1). This paper aims to provide an argument to show that, given any field of definition K of S, there is a p-gonal field of definition F which is an extension of degree at most 2(p−1) over K. If φ is an automorphism of S, then we say that S and φ are simultaneously defined over K if there is an algebraic curve model of S, defined over K, such that φ is given by a rational map on it with coefficients in K. Theorem 1.2. Let S be a cyclic p-gonal Riemann surface of genus g ≥ 2, with a p-gonal automorphism φ, and let K be a field of definition of S. Then (1) There is p-gonal field of definition of S, this being an extension of degree at most 2(p− 1) of K (which is also a field of definition of φ). (2) If both S and φ are simultaneously defined over K, then there is a p-gonal field of definition of S, this being an extension of degree at most two of K. (3) If in Equation (1.1) n1 = · · · = nm, then there is a p-gonal field of definition of S, this being an extension of degree at most two of K. Remark 1.3. Theorem 1.2 is still valid if we change C to any algebraically closed field, where in positive characteristic we need to assume that p is different from the characteristic. Remark 1.4. For each integer n ≥ 2, not necessarily prime, the definition of cyclic n- gonal Riemann surface S, n-gonal automorphism φ and n-gonal group ⟨φ⟩ is the same as for the prime situation. In the particular case that every fixed point of a non-trivial power φk is also a fixed point of φ, the definition of an n-gonal curve is the same as in (1.1), but replacing p by n and assuming each the exponent nj to be relatively prime to n. In this case, under the assumption that S has a unique n-gonal group ⟨φ⟩ (this is the situation for generalized superelliptic Riemann surfaces [15]), then the arguments of the proof of Theorem 1.2 allows us to obtain that: if K is a field of definition of S, then there is an n- gonal field of definition of S, this being an extension of degree at most 2ϕ(n) of K, where ϕ(n) is the ϕ-Euler function. 2 An application to hyperelliptic Riemann surfaces Let S be a hyperelliptic Riemann surface (i.e., p = 2) with hyperelliptic involution φ and let K be a field of definition of S. As φ is unique, one may consider the group Autred(S) := Aut(S)/⟨φ⟩, called the reduced group of automorphisms of S. For even genus, in [22], J-F. Mestre proved that S is also hyperelliptically definable over K. If the genus is odd, then the previous fact is in general false; as can be seen from examples in [8, 9, 20, 21]. In [16], B. Huggins proved that if Autred(S) is neither trivial nor cyclic, then S is also hyperelliptically definable over K. In [21], R. Lercier, C. Ritzenthaler and J. Sijslingit proved that S can be hyperelliptically defined over a quadratic extension of K if the reduced group is a non-trivial cyclic group. Our theorem asserts that this fact is still valid even if the reduced group is trivial. Corollary 2.1. If K is a field of definition of a hyperelliptic Riemann surface, then it is hyperelliptically definable over an extension of degree at most two of K. 408 Ars Math. Contemp. 24 (2024) #P3.02 / 405–414 3 An application to fields of moduli Let S be a closed Riemann surface and let C be an irreducible algebraic curve representing it. The field of moduli MS of S is the fixed field of the group ΓC = {σ ∈ Aut(C/Q) : Cσ ∼= C}; this field does not depend on the choice of the algebraic model C. In [18], S. Koizumi proved that MS coincides with the intersection of all fields of definition of S, but in general it might not be a field of definition [6, 7, 12, 16, 19]. If Aut(S) is trivial (the generic situation for g ≥ 3), then Weil’s descent theorem [25] asserts that MS is a field of definition of S. In [26], J. Wolfart proved that if S/Aut(S) is the Riemann sphere with exactly 3 cone points (i.e., S is quasiplatonic), then MS is also a field of definition of S. In a more general setting, if S/Aut(S) has genus zero, then it is known that S is definable over an extension of degree at most two of MS (see [14] for a more general statement). Now, let S be a p-gonal Riemann surface of genus g ≥ 2 and let G = ⟨φ⟩ < Aut(S) be a p-gonal group. As previously noted, S is either definable over MS or over a suitable quadratic extension of it (but it might not be cyclically p-gonally definable over such a minimal field of definition). In the case that G is not a unique p-gonal subgroup, in [28], A. Wootton noted that S can be cyclically p-gonally defined over an extension of degree at most 2 of its field of moduli. In the case that G is the unique p-gonal subgroup, the quotient group Aut(S)/G is called the reduced group of S. In [19], A. Kontogeorgis proved that if the reduced group is neither trivial nor a cyclic group, then S can always be defined over its field of moduli. So, a direct consequence of Theorem 1.2 is the following. Corollary 3.1. Let S be a cyclic p-gonal Riemann surface with a p-gonal group G = ⟨φ⟩. (1) If G is not a normal subgroup of Aut(S), then S is cyclically p-gonally definable over an extension of degree at most two of MS . (2) If G is a normal subgroup of Aut(S) and Aut(S)/G is different from the trivial group or a cyclic group, then S is cyclically p-gonally definable over an extension of degree at most 2(p − 1) of MS . Moreover, if φ also is defined over MS , then the extension can be chosen to be of degree at most two. (3) If G = Aut(S), then S is cyclically p-gonally definable over an extension of degree at most 4(p− 1) of its field of moduli. Moreover, if φ also is defined over MS , then the extension can be chosen of degree at most 4. As every hyperelliptic Riemann surface is definable over an extension of degree at most two of its field of moduli, Corollary 2.1 asserts the following. Corollary 3.2. Every hyperelliptic Riemann surface is hyperelliptically definable over an extension of degree at most 4 of its field of moduli. Moreover, if either (i) the genus is even or (ii) the genus is odd and the reduced group is not trivial, then the hyperelliptic Riemann surface is hyperelliptically defined over an extension of degree at most 2 of its field of moduli. Examples of hyperelliptic Riemann surfaces with a trivial reduced group that cannot be defined over their field of moduli were provided by C. J. Earle [6, 7] and G. Shimura [24]. The same type of examples, but with a non-trivial cyclic reduced group, were provided by B. Huggins [16]. R. A. Hidalgo: On p-gonal fields of definition 409 4 Proof of Theorem 1.2 We assume the p-gonal Riemann surface S to be provided by an irreducible curve C, de- fined over a subfield K of C. If K is the algebraic closure of K inside C, then (in this algebraic model) the p-gonal automorphism φ is given by a rational map defined over K. We divide the arguments depending on the uniqueness of the cyclic group ⟨φ⟩. 4.1 The case when ⟨φ⟩ is not unique The following result, due to A. Wootton, describes those cases where the uniqueness fails. Theorem 4.1 ([28, A. Wootton]). Let S be a cyclic p-gonal Riemann surface of genus g ≥ 2 and let m = 2(g+p−1)/(p−1). If (m, p) is different from any the following tuples (i) (3, 7), (ii) (4, 3), (iii) (4, 5), (iv) (5, 3), (v) (p, p), p ≥ 5, (vi) (2p, p), p ≥ 3, then S has a unique p-gonal group. In the same paper, Wootton describes the exceptional cyclic p-gonal Riemann surfaces, ie., where the p-gonal group is non-unique. (i) Case (m, p) = (3, 7) corresponds to Klein’s quartic (a non-hyperelliptic Riemann surface of genus 3) x3y+y3z+z3x = 0, whose group of automorphisms is PGL2(7) (of order 168). This surface is cyclically 7-gonally defined as y7 = x2(x− z)z4. (ii) Case (m, p) = (4, 3) corresponds to the genus 2 Riemann surface defined hyperel- liptically by y2z3 = x(x4−z4), whose group of automorphisms is GL2(3) (of order 48). This surface is cyclically 3-gonally defined as y3z3 = (x2 − z2)(x2 − (15 √ 3− 26)z2)2. (iii) Case (m, p) = (4, 5) corresponds to the genus 4 non-hyperelliptic Riemann surface, called Bring’s curve, which is the complete intersection of the quadric x1x4+x2x3 = 0 and the cubic x21x3 + x 2 2x1 + x 2 3x4 + x 2 4x2 = 0 in the 3-dimensional complex projective space. Its group of automorphisms is S5, the symmetric group in five letters S5. This surface is cyclically 5-gonally defined as y5z5 = (x2−z2)(x2+z2)4. (iv) Case (m, p) = (5, 3) corresponds to the genus 3 non-hyperelliptic closed Riemann surface x4+y4+ z4+2i √ 3z2y2 = 0, whose group of automorphisms has order 48. The quotient of that surface by its group of automorphisms has signature (0; 2, 3, 12). This surface is cyclically 3-gonally defined as y3z3 = x2(x4 − z4). (v) Case (m, p) = (p, p), where p ≥ 5, corresponds to the Fermat curve xp+yp+zp = 0, whose group of automorphisms is Z2p ⋊ S3. This is already in a p-gonal form as yp = −zp − xp. (vi) Case (m, p) = (2p, p), where p ≥ 3. There is a 1-dimensional family with group of automorphisms Z2p ⋊ Z22 (the quotient by that group has signature (0; 2, 2, 2, p)). Also, there is a surface with group of automorphisms Z2p ⋊D4 (the quotient by that group has signature (0; 2, 4, 2p). These surfaces are cyclically p-gonally defined as ypzp = (xp − apzp)(xp − zp/ap) = x2p − (ap + 1/ap)xpzp + z2p. 410 Ars Math. Contemp. 24 (2024) #P3.02 / 405–414 Note that, in all the above exceptional cases, the surface S is cyclically p-gonally de- fined over an extension of degree at most 2 over the field of moduli. In fact, with only the exception of case (ii), S is cyclically p-gonally defined over its field of moduli. So, we are done in this situation. 4.2 The case when ⟨φ⟩ is unique We now assume that ⟨φ⟩ is unique. Set Γ = Gal(K/K). Let us consider a rational map π : C → P1K, defined over K, which is a regular branched covering with ⟨φ⟩ as its deck group and whose branch values are a1, ..., am ∈ C (in fact, these values belong to K). Let the integers n1, ..., nm ∈ {1, ..., p − 1}, n1 + · · · + nm ≡ 0 mod p, be such that C is isomorphic to a p-gonal curve E with Equation (1.1). 4.2.1 Proof of Part (1) Let us recall that φ is already defined over K. In the next, we note that φ is defined over an extension of K of degree at most p− 1. Claim 4.2. The rational map φ is defined over an extension K1 of K of degree at most p− 1. Proof. If σ ∈ Γ, then φσ is an automorphism of order p of Cσ = C. As we are assuming the uniqueness of ⟨φ⟩, we must have that φσ ∈ Ω := {φ,φ2, . . . , φp−1}. In particular, the subgroup A of Γ consisting of those σ such that φσ = φ must have index at most the cardinality of the set Ω, which is p − 1. This asserts that φ is defined over the fixed field K1 of A, which is an extension of degree at most p− 1 of K. Set Γ1 = Gal(K/K1). If τ ∈ Γ1, then (as the identity I : C → C = Cτ conjugates ⟨φ⟩ = ⟨φ⟩τ = ⟨φτ ⟩ to itself), there is a (unique) automorphism gτ of P1K such that π τ = πτ ◦ I = gτ ◦ π (see the following diagram). C I−−−−→ C = Cτ π y πτy P1K gτ−−−−→ P1K As the group of automorphisms of P1K is given by Möbius transformations (i.e., ele- ments of PGL2(K)), we must have gτ ∈ PGL2(K). We may apply each σ ∈ Γ1 to the above diagram to obtain the following one Cσ = C I−−−−→ C = Cστ πσ y πστy P1K gστ−−−−→ P1K The above permits us to obtain the following diagram R. A. Hidalgo: On p-gonal fields of definition 411 C I−−−−→ C = Cσ I=I σ −−−−→ C = Cστ π y πσy πστy P1K gσ−−−−→ P1K gστ−−−−→ P1K As the transformation gρ is uniquely determined by ρ ∈ Γ1, the collection {gρ}ρ∈Γ1 satisfies the co-cycle relation gστ = g σ τ ◦ gσ, σ, τ ∈ Γ1. Weil’s descent theorem [25] ensures the existence of a genus zero irreducible and non- singular algebraic curve B, defined over K1, and an isomorphism R : P1K → B, defined over K, so that gσ ◦Rσ = R, σ ∈ Γ1. Also, for σ ∈ Γ1, we have {σ(a1), ..., σ(am)} = {gσ(a1), ..., gσ(am)}, so it follows that {R(a1), ..., R(am)} is Γ1-invariant. Let us denote by A(nj) the set of those ak’s for which nk = nj . Claim 4.3. Each set R(A(nj)) is Γ1-invariant. Proof. If σ ∈ Γ1, then (as πσ = gσ ◦ π) the set gσ(A(nj)) corresponds to the set of those σ(ak) having the same nl (for some l), that is, gσ(A(nj)) = σ(A(nl)). As φσ = φ, we must have nl = nj , that is, gσ(A(nj)) = σ(A(nj)). This last equality implies the desired claim. Claim 4.4. There is an effective K1-rational divisor U ≥ 0 of degree at most two in B. Proof. We follow similar techniques as used by Huggins in her thesis [16] (and other au- thors). Let us consider any K1-rational meromorphic 1-form ω in B. Since B has genus zero, the canonical divisor K = (ω) is a K1-rational of degree −2. In this way, there is a positive integer d such that the divisor D = R(a1) + · · ·+R(am) + dK is K1-rational of degree 1 or 2. If D ≥ 0, then we set U := D. Let us assume D is not effective. Let us consider the Riemann-Roch space L(D), consisting of those non-constant rational maps ϕ : B → P1K whose divisors satisfy (ϕ) + D ≥ 0 together with the constant ones. As the divisor D is K1-rational, for every σ ∈ Γ1 and every ϕ ∈ L(D), it follows that ϕσ ∈ L(D). This, in particular, permits us to observe that we can find a basis of L(D) consisting of rational maps defined over K1. One of the elements of such a basis must be a non-zero constant map. As, by Riemann-Roch’s theorem, L(D) has dimension 2 (if D has degree one) or 3 (if D has degree two), we may find a non-constant f ∈ L(D) belonging to such a basis (defined over K1). In this case, we may take U = (f) +D ≥ 0. By Claim 4.4, there is an effective K1-rational divisor U of degree 1 or 2 and U ≥ 0. We have three possibilities: (1) U = s, where s ∈ B is K1-rational; or (2) U = 2t, where t ∈ B is K1-rational; or 412 Ars Math. Contemp. 24 (2024) #P3.02 / 405–414 (3) U = r + q, where r, q ∈ B, r ̸= q, and {r, q} is Γ1-invariant. In cases (1) and (2) we have the existence of a K1-rational point in B. In this case, we set K2 = K1. In case (3) we have a point (say r) in B which is rational over a quadratic extension K2 of K1. Let b ∈ B be a K2-rational point (whose existence is provided above). By Riemann- Roch’s theorem, the Riemann-Roch space L(b) (where b is thought of as a divisor of degree one) has dimension 2. Similarly as above, we may choose a basis {1, L} of L(b), with each element defined over K2. In this case, L : B → Ĉ turns out to be an isomorphism defined over K2. We have that Q = L ◦R ◦ π : C → Ĉ is a Galois (branched) covering with deck group ⟨φ⟩ and whose branch values are {L(R(a1)), ..., L(R(am))}. It follows that S is p-gonally defined by yp = F (x) = m∏ j=1 (x− L(R(aj)))nj . As the sets {L(R(a1)), ..., L(R(am))} and L(R(A(nj))) are Gal(K/K2)-invariant (by Claim 4.3 and the fact that K1 is a subfield of K2), it follows that F (x) =∏m j=1 (x− L(R(aj))) nj ∈ K2[x]. As K2 is an extension of degree at most two of K1 and the last one is an extension of degree at most p− 1 of K, we are done. 4.2.2 Proof of Parts (2) and (3) If φ is already defined over K then we assume K1 = K (i.e., we set Γ1 = Γ) in the above arguments. Similarly, if in Equation (1.1) we have that n1 = · · · = nm = n, then there will be only one set A(n). In this case, in the previous arguments, we do not need to use Claim 4.3 (where it was needed for the choice of K1) and we may work as in the proof of Part (1) with K instead of K1. ORCID iDs Ruben A. Hidalgo https://orcid.org/0000-0003-4070-2819 References [1] R. D. M. Accola, On cyclic trigonal Riemann surfaces. I, Trans. Amer. Math. Soc. 283 (1984), 423–449, doi:10.2307/1999139, https://doi.org/10.2307/1999139. [2] G. Bartolini, A. F. Costa and M. Izquierdo, On automorphisms groups of cyclic p-gonal Riemann surfaces, J. Symbolic Comput. 57 (2013), 61–69, doi:10.1016/j.jsc.2013.05.005, https://doi.org/10.1016/j.jsc.2013.05.005. [3] S. A. Broughton and A. Wootton, Topologically unique maximal elementary abelian group actions on compact oriented surfaces, J. Pure Appl. Algebra 213 (2009), 557–572, doi:10. 1016/j.jpaa.2008.08.003, https://doi.org/10.1016/j.jpaa.2008.08.003. [4] E. Bujalance, F. J. Cirre and G. Gromadzki, Groups of automorphisms of cyclic trigonal Riemann surfaces, J. Algebra 322 (2009), 1086–1103, doi:10.1016/j.jalgebra.2009.05.017, https://doi.org/10.1016/j.jalgebra.2009.05.017. [5] A. F. Costa and M. Izquierdo, Maximal order of automorphisms of trigonal Riemann sur- faces, J. Algebra 323 (2010), 27–31, doi:10.1016/j.jalgebra.2009.09.041, https://doi. org/10.1016/j.jalgebra.2009.09.041. R. A. Hidalgo: On p-gonal fields of definition 413 [6] C. J. Earle, On the moduli of closed Riemann surfaces with symmetries, in: L. V. Ahlfors, L. Bers, H. M. Farkas, R. C. Gunning, I. Kra and H. E. Rauch (eds.), Advances in the Theory of Riemann Surfaces, Princeton Univ. Press, Princeton, volume 66 of Ann. of Math. Stud., pp. 119–130, 1971. [7] C. J. Earle, Diffeomorphisms and automorphisms of compact hyperbolic 2-orbifolds, in: Ge- ometry of Riemann Surfaces, Cambridge Univ. Press, Cambridge, volume 368 of London Math. Soc. Lecture Note Ser., pp. 139–155, 2010. [8] Y. Fuertes, Fields of moduli and definition of hyperelliptic curves of odd genus, Arch. Math. 95 (2010), 15–81, doi:10.1007/s00013-010-0138-9, https://doi.org/10. 1007/s00013-010-0138-9. [9] Y. Fuertes and G. González-Diez, Erratum to: Fields of moduli and definition of hyperelliptic covers, Arch. Math. 101 (2013), 599–600, doi:10.1007/s00013-013-0582-4, https://doi. org/10.1007/s00013-013-0582-4. [10] G. González-Diez, On prime Galois coverings of the Riemann sphere, Ann. Mat. Pura Appl. (4) 168 (1995), 1–15, doi:10.1007/BF01759251, https://doi.org/10.1007/ BF01759251. [11] G. Gromadzki, A. Weaver and A. Wootton, On gonality of Riemann surfaces, Geom. Dedi- cata 149 (2010), 1–14, doi:10.1007/s10711-010-9459-x, https://doi.org/10.1007/ s10711-010-9459-x. [12] R. A. Hidalgo, Non-hyperelliptic Riemann surfaces with real field of moduli but not definable over the reals, Arch. Math. 93 (2009), 219–224, doi:10.1007/s00013-009-0025-4, https: //doi.org/10.1007/s00013-009-0025-4. [13] R. A. Hidalgo, p-groups acting on Riemann surfaces, J. Pure Appl. Algebra 222 (2018), 4173– 4188, doi:10.1016/j.jpaa.2018.02.026, https://doi.org/10.1016/j.jpaa.2018. 02.026. [14] R. A. Hidalgo, A remark on the field of moduli of Riemann surfaces, Arch. Math. 114 (2020), 515–526, doi:10.1007/s00013-019-01411-9, https://doi.org/10.1007/ s00013-019-01411-9. [15] R. A. Hidalgo, S. Quispe and T. Shaska, Generalized superelliptic riemann surfaces, 2016, arXiv:1609.09576 [math.AG]. [16] B. Huggins, Fields of Moduli and Fields of Definition of Curves, Ph.D. thesis, UCLA, 2005. [17] A. Hurwitz, Ueber algebraische Gebilde mit eindeutigen Transformationen in sich, Math. Ann. 41 (1892), 403–442, doi:10.1007/BF01443420, https://doi.org/10.1007/ BF01443420. [18] S. Koizumi, The fields of moduli for polarized abelian varieties and for curves, Nagoya Math. J. 48 (1972), 37–55, http://projecteuclid.org/euclid.nmj/1118798785. [19] A. Kontogeorgis, Field of moduli versus field of definition for cyclic covers of the projec- tive line, J. Théor. Nombres Bordeaux 21 (2009), 679–692, http://jtnb.cedram.org/ item?id=JTNB_2009__21_3_679_0. [20] R. Lercier and C. Ritzenthaler, Hyperelliptic curves and their invariants: geometric, arithmetic and algorithmic aspects, J. Algebra 372 (2012), 595–636, doi:10.1016/j.jalgebra.2012.07.054, https://doi.org/10.1016/j.jalgebra.2012.07.054. [21] R. Lercier, C. Ritzenthaler and J. Sijsling, Explicit Galois obstruction and descent for hyperel- liptic curves with tamely cyclic reduced automorphism group, Math. Comp. 85 (2016), 2011– 2045, doi:10.1090/mcom3032, https://doi.org/10.1090/mcom3032. 414 Ars Math. Contemp. 24 (2024) #P3.02 / 405–414 [22] J.-F. Mestre, Construction de courbes de genre 2 à partir de leurs modules. (French) [Construct- ing genus-2 curves from their moduli], in: Effective Methods in Algebraic Geometry (Cas- tiglioncello, 1990), Birkhäuser Boston, Boston, MA, volume 94 of Progr. Math., pp. 313–334, 1991. [23] H. A. Schwarz, Ueber diejenigen algebraischen Gleichungen zwischen zwei veränderlichen Grössen, welche eine Schaar rationaler eindeutig umkehrbarer Transformationen in sich selbst zulassen, J. Reine Angew. Math. 87 (1879), 139–145, doi:10.1515/crll.1879.87.139, https: //doi.org/10.1515/crll.1879.87.139. [24] G. Shimura, On the field of rationality for an abelian variety, Nagoya Math. J. 45 (1972), 167– 178, http://projecteuclid.org/euclid.nmj/1118798531. [25] A. Weil, The field of definition of a variety, Amer. J. Math. 78 (1956), 509–524, doi:10.2307/ 2372670, https://doi.org/10.2307/2372670. [26] J. Wolfart, ABC for polynomials, dessins d’enfants and uniformization—a survey, in: Ele- mentare und Analytische Zahlentheorie, Franz Steiner Verlag Stuttgart, Stuttgart, volume 20 of Schr. Wiss. Ges. Johann Wolfgang Goethe Univ. Frankfurt am Main, pp. 313–345, 2006. [27] A. Wootton, Defining equations for cyclic prime covers of the Riemann sphere, Israel J. Math. 157 (2007), 103–122, doi:10.1007/s11856-006-0004-4, https://doi.org/10.1007/ s11856-006-0004-4. [28] A. Wootton, The full automorphism group of a cyclic p-gonal surface, J. Algebra 312 (2007), 377–396, doi:10.1016/j.jalgebra.2007.01.018, https://doi.org/10.1016/j. jalgebra.2007.01.018. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.03 / 415–425 https://doi.org/10.26493/1855-3974.2701.b7d (Also available at http://amc-journal.eu) An online bin-packing problem with an underlying ternary structure Helmut Prodinger Mathematics Department, Stellenbosch University, 7602 Stellenbosch, South Africa and NITheCS (National Institute for Theoretical and Computational Sciences), South Africa Received 19 September 2021, accepted 30 June 2023, published online 12 June 2024 Abstract Following an orginal idea by Knödel, an online bin-packing problem is considered where the large items arrive in double-packs. The dual problem where the small items arrive in double-packs is also considered. The enumerations have a ternary random walk flavour, and for the enumeration, the kernel method is employed. Keywords: Knödel walks, third-order recursion, kernel method, coefficient extraction, state diagram. Math. Subj. Class. (2020): 05A15, 68R05 1 Introduction Walter Knödel introduced the following online bin-packing problem [3]: There are bins of size 1, and random items of size 23 (large items) and of size 1 3 (small items) appear and are put into the boxes. A typical scenario is that a number j of partially filled boxes exist, and the number j becomes j + 1 resp. j − 1, depending on whether a the new item is of large resp. small type. “At random” means that both types appear with the same probability 12 . In my collection of examples [4], I showed how to deal with the Knödel problem us- ing the kernel method. I was, however, not the only author who was intrigued by such questions; a notable paper is by Michael Drmota [1], which is of a more probabilistic type, whereas I tried to emphasize the combinatorial point of view. The present paper has a certain ‘ternary’ flavour: the next section deals with the instance of large items appearing in double-packs. The handler breaks off the double-packs, and then treats the items as Knödel would have done. Typically, the number of partially filled boxes increases by 2 or decreases by 1. In order to keep the system balanced, we assume that the small items appear twice as often as the double-packs. E-mail address: hproding@sun.ac.za (Helmut Prodinger) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 416 Ars Math. Contemp. 24 (2024) #P3.03 / 415–425 The last section deals with the dual problem, where the small items appear in double- packs and the large items as single units. The kernel method is used to obtain all the relevant enumerations. The recent paper [5] served as an inspiration, but deals with a different issue. It must be said that, when [4] was prepared, such ternary questions would have been outside of my reach. Luckily, now, they are not. We confine ourselves here just to enumerations, deriving explicit generating functions in one or two variables. Questions of a more probabilistic nature are not treated. 2 The first model The following items arrive at random: a double-pack of items, each of size 23 , and an item of size 13 . We could equip the set-up with general probabilities p and q = 1 − p, but we restrict ourselves to the ‘balanced’ case where the single items are twice as likely as the double-packs, so we set p = 13 and q = 2 3 . The following state diagram (we show only a finite part of it) describes the situation. There are states representing ‘i boxes filled to 23 ’; a double-pack pushes the i to i+ 2, and a single item reduces it to i − 1. There is an exceptional state, called β, standing for one box, filled to 13 . The red edges represent an arrival of a double-pack, and will be labelled by pz; the black edges represent an arrival of a single item, and will be labelled by qz. 0 1 2 3 4 5 6 7 8 β From the state diagram, we set off an infinite set of generating functions in the variable z, where the coefficient of zn is the probability that n random steps lead to state i, for i ≥ 0 or i = β. Mostly, we just write fi instead of fi(z). The following system of recursions can be read off immediately: f0 = 1 + qzf1, fβ = qzf0, f1 = zfβ + qzf2 = qz 2f0 + qzf2, fi = pzfi−2 + qzfi+1, i ≥ 2. Our method to solve this system is the kernel method. For that, we introduce a bivariate H. Prodinger: An online bin-packing problem with an underlying ternary structure 417 generating function F (u, z), but we mostly write just F (u): F (u) = ∑ i≥0 uifi(z) = 1 + qzf1 + qz 2uf0 + qzuf2 + ∑ i≥2 ui [ pzfi−2 + qzfi+1 ] = 1 + qzf1 + qz 2uf0 + pzu 2F (u) + ∑ i≥1 uiqzfi+1 = 1 + qzf1 + qz 2uf0 + pzu 2F (u) + qz u ( F (u)− f0 − uf1 ) = 1 + qz2uf0 + pzu 2F (u) + qz u ( F (u)− f0 ) Note that f0 = F (0). It is beneficial to introduce the new variable u = zU ; doing this, powers of z that appear are multiples of 3. Later, it will be convenient to set x = z3. As can be seen, the numbers of steps leading to a state i belong to just one residue class modulo 3. We compute F (u) = −3U − 2z3U2f0 + 2f0 z3U3 − 3U + 2 = −3U − 2xU2f0 + 2f0 xU3 − 3U + 2 . As it is common using the kernel method, setting U = 0 leads to a void equation. How- ever, factorizing the denominator is the method of choice. There is ‘bad’ factor in the denominator, which must also appear in the numerator, which allows us to compute f0 and consequently the whole bivariate generating function. In order to deal with the ternary equation successfully, we further set x = z3 = 274 t(1− t) 2 and we find the 3 roots U1 = 2 3(1− t) , U2 = 1 σ , U3 = 1 τ , with σ = 3 4 (t− √ 4t− 3t2 ), τ = 3 4 (t+ √ 4t− 3t2). A motivation for this substitution is the Lagrange inversion formula and/or the enumeration of ternary trees; see also [6]. Plugging U = 23(1−t) into the numerator (this is the bad factor, as explained a little bit later), leads to f0 = 1 (1− t)(1− 3t) and furthermore to the simplified numerator −3U − 2xU2f0 + 2f0 U − 23(1−t) = 1 1− 3t ( − 3 + 27 2 t(t− 1)U ) . 418 Ars Math. Contemp. 24 (2024) #P3.03 / 415–425 The variable x is given in terms of t. The inverse relation is of interest. It can be obtained by the Lagrange inversion formula or, as here, by contour integration: [xk]t = 1 2πi ∮ dx xk+1 t = 1 2πi 27 4 ( 4 27 )k+1 ∮ dt(1− t)(1− 3t) tk+1(1− t)2k+2 t = 1 2πi ( 4 27 )k ∮ dt(1− 3t) tk(1− t)2k+1 = ( 4 27 )k [tk−1] 1− 3t (1− t)2k+1 = ( 4 27 )k[(3k − 1 k − 1 ) − 3 ( 3k − 2 k − 2 )] , which, after simplification, gives us t = ∑ k≥1 1 k ( 3k − 2 k − 1 ) 22k 33k xk. A similar computation leads to 1 1− t = ∑ k≥0 1 2k + 1 ( 3k k ) 22k+1 33k+1 xk. From this we infer that for z ∼ 0, U ∼ 23 , or u ∼ 2 3z, explaining why we are talking about the bad factor. We continue the computation: F (u) = 1 1− 3t ( − 3 + 27 2 t(t− 1)U ) 1 x(U − 1σ )(U − 1 τ ) = 1 1− 3t ( − 3 + 27 2 t(t− 1)U ) 9 4 t(t− 1) x(1− σU)(1− τU) = 1 (1− 3t)(1− t) ( 1− 9 2 t(t− 1)U ) 1 (1− σU)(1− τU) . Partial fraction decomposition leads to (we use the abbreviation W = √ 4t− 3t2 ) 1 (1− σU)(1− τU) = 1 2 ( 1− t W ) 1 1− σU + 1 2 ( 1 + t W ) 1 1− τU = 1 2 [ 1 1− σU + 1 1− τU ] + t 2W [ 1 1− τU − 1 1− σU ] = 1 2 [ 1 1− σU + 1 1− τU ] + 3t 4(τ − σ) [ 1 1− τU − 1 1− σU ] . For the further simplification we will resort to two identities going by the name of Girard- Waring formula, see e. g. [2]: Xm + Y m = ∑ 0≤k≤m/2 (−1)k ( m− k k ) m m− k (XY )k(X + Y )m−2k; Xm − Y m X − Y = ∑ 0≤k≤(m−1)/2 (−1)k ( m− 1− k k ) (XY )k(X + Y )m−1−2k. H. Prodinger: An online bin-packing problem with an underlying ternary structure 419 Of course, we will apply them with X = τ and Y = σ. Then [Um] 1 2 [ 1 1− σU + 1 1− τU ] = 1 2 ∑ 0≤k≤m/2 (−1)k ( m− k k ) m m− k (9 4 t(t− 1) )k(3 2 t )m−2k = 1 2 (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) m m− k tk(t− 1)ktm−2k = 1 2 (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) m m− k (t− 1)ktm−k and [Um] 3t 4(τ − σ) [ 1 1− τU − 1 1− σU ] = 3t 4 ∑ 0≤k≤(m−1)/2 (−1)k ( m− 1− k k )(9 4 t(t− 1) )k(3 2 t )m−1−2k = (3 2 )m−1 3t 4 ∑ 0≤k≤(m−1)/2 (−1)k ( m− 1− k k ) (t− 1)ktm−1−k = 1 2 (3 2 )m ∑ 0≤k≤(m−1)/2 (−1)k ( m− 1− k k ) (t− 1)ktm−k. Combining the two leads to a pleasant simplification: [Um] 1 2 [ 1 1− σU + 1 1− τU ] + [Um] 3t 4(τ − σ) [ 1 1− τU − 1 1− σU ] = (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) (t− 1)ktm−k, or simpler [Um] 1 (1− σU)(1− τU) = (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) (t− 1)ktm−k. We need a second similar term: [Um] ( − 9 2 t(t− 1)U ) 1 (1− σU)(1− τU) = −9 2 t(t− 1)[Um−1] 1 (1− σU)(1− τU) = −9 2 t(t− 1) (3 2 )m−1 ∑ 0≤k≤m/2 (−1)k ( m− 1− k k ) (t− 1)ktm−1−k = −3 (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− 1− k k ) (t− 1)k+1tm−k. 420 Ars Math. Contemp. 24 (2024) #P3.03 / 415–425 Putting everything together we found F (u) = 1 (1− 3t)(1− t) ∑ m≥0 um zm (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) (t− 1)ktm−k − 3 (1− 3t)(1− t) ∑ m≥0 um zm (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− 1− k k ) (t− 1)k+1tm−k. Reading off the coefficients of zj for j ≥ 1 as well is now done with Cauchy’s integral for- mula; the contours are always small circles (or equivalent) around the origin. The starting point is [znuj ]F (u) = (3 2 )j [zn+j ] 1 (1− 3t)(1− t) ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)ktj−k − (3 2 )j [zn+j ] 3 (1− 3t)(1− t) ∑ 0≤k≤j/2 (−1)k ( j − 1− k k ) (t− 1)k+1tj−k and we will treat the two sums separately. There is only a contribution if n+ j ≡ 0 mod 3. (This is also clear from the combinatorial context.) Assume this and set N := n+j3 . Step 1: [xN ] 1 (1− 3t)(1− t) ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)ktj−k = 1 2πi ∮ dx xN+1 1 (1− 3t)(1− t) ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)ktj−k = 1 2πi ∮ 27 4 dt( 27 4 t(1− t)2 )N+1 ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)ktj−k = 1 2πi ∮ ( 4 27 )N dt ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)k−2N−2tj−k−N−1 = ( 4 27 )N [tN−j+k] ∑ 0≤k≤j/2 (−1)k ( j − k k ) (t− 1)k−2N−2 = ( 4 27 )N (−1)N−j ∑ 0≤k≤j/2 ( j − k k )( k − 2N − 2 N − j + k ) . Step 2: [xN ] 3 (1− 3t)(1− t) ∑ 0≤k≤j/2 (−1)k ( j − 1− k k ) (t− 1)k+1tj−k = 3 ( 4 27 )N [tN−j+k] ∑ 0≤k≤j/2 (−1)k ( j − 1− k k ) (t− 1)k−2N−1 = 3 ( 4 27 )N (−1)N−j ∑ 0≤k≤j/2 ( j − 1− k k )( k − 2N − 1 N − j + k ) . H. Prodinger: An online bin-packing problem with an underlying ternary structure 421 We put all the results of this section together in a theorem. Theorem 2.1. The generating function F (u) = F (u, z) has the following explicit form: F (u) = 1 (1− 3t)(1− t) ( 1− 9 2 t(t− 1)U ) 1 (1− σU)(1− τU) Here, u = zU , z3 = x = 274 t(1− t) 2, and σ = 3 4 (t− √ 4t− 3t2 ), τ = 3 4 (t+ √ 4t− 3t2 ). Note that (1 − σU)(1 − τU) = 1 − 32 tU + 9 4 t(t − 1)U 2. Written in the new variable U , only powers of z that are multiples of 3 appear. Further, we get the representation sorted by powers of u: F (u, z) = 1 (1− 3t)(1− t) ∑ m≥0 um zm (3 2 )m ∑ 0≤k≤m/2 (−1)k ( m− k k ) (t− 1)ktm−k − 3 (1− 3t)(1− t) ∑ m≥0 um zm (3 2 )m ∑ 0≤k≤(m−1)/2 (−1)k ( m− 1− k k ) (t− 1)k+1tm−k. Reading off coefficients of zNuj , where N = n+j3 leads to [zNuj ]F (u, z) = ( 4 27 )N (−1)N−j ∑ 0≤k≤j/2 ( j − k k )( k − 2N − 2 N − j + k ) − 3 ( 4 27 )N (−1)N−j ∑ 0≤k≤(j−1)/2 ( j − 1− k k )( k − 2N − 1 N − j + k ) . For the special state β, the following series representation holds: fβ(z) = ∑ n≥0 22n+1 33n+1 ( 3n+ 1 n ) z3n+1. The computation for the special state was not shown yet: [z3n+1]fβ = 2 3 [xn] 1 (1− t)(1− 3t) = 2 3 1 2πi ∮ dx xn+1 1 (1− t)(1− 3t) = 2 3 27 4 1 2πi ∮ dt ( 274 ) n+1tn+1(1− t)2n+2 = 2 3 ( 4 27 )n 1 2πi ∮ dt tn+1(1− t)2n+2 = 2 3 ( 4 27 )n [tn] 1 (1− t)2n+2 = 22n+1 33n+1 ( 3n+ 1 n ) . 3 The dual model Now, the red edges mean the arrival of the large objects (size 23 ) and the black edges mean a double-pack of the small edges (size 13 each). To keep the system balanced, the large objects should arrive twice as often as the double-packs of small edges. Again, the generating 422 Ars Math. Contemp. 24 (2024) #P3.03 / 415–425 function gi refers to paths of length n leading eventually into state i. After n steps, only a state i can be reached with n ≡ i mod 3. The state diagram and the recursions are immediate: 0 1 2 3 4 5 6 7 8 β We work only with p = 23 , q = 1 3 . Directly from the state diagram, g0 = 1 + zgβ + qzg2 = 1 + qz 2g1 + qzg2, gβ = qzg1, g1 = zg0 + qzg3, gi = pzgi−1 + qzgi+2, i ≥ 2. Summing the recursions, G(u) = ∑ i≥0 uigi(z) = g0 + ug1 + ∑ i≥2 ui ( pzgi−1 + qzgi+2 ) = g0 + uzg0 + qzug3 + pzu ∑ i≥1 uigi + qz u2 ∑ i≥4 uigi = g0 + uzg0 + pzuG(u)− pzug0 + qz u2 ∑ i≥3 uigi = g0 + uzg0 + pzuG(u)− pzug0 + qz u2 (G(u)− g0 − ug1 − u2g2) = g0 + quzg0 + pzuG(u) + qz u2 G(u)− qz u2 g0 − qz u g1 − qzg2 = g0 + quzg0 + pzuG(u) + qz u2 G(u)− qz u2 g0 − qz u g1 + 1 + qz 2g1 − g0. Solving, we find with V = uz: G(u) = −V 3g0 − 3V 2 − g1V 2z2 + z3g0 + g1V z2 2V 3 − 3V 2 + x . Now we factorize the denominator, using the same substitutions x = z3 and x = 274 t(1 − t)2: 2(V − 32 (1− t))(V − σ)(V − τ) = 2V 3 − 3V 2 + x. This time, both, (V −σ) and (V − τ) are bad factors. Plugging into the numerator, we find two equations, and the solutions: g0 = 4 (1− 3t)(4− 3t) , g1 = 27t(1− t) z2(1− 3t)(4− 3t) . It can be noted that Ṽ := V −1, with the denominator of the previous section; thus, the three roots carry over. Dividing out the bad factors, we find −V 3g0 − 3V 2 − g1V 2z2 + z3g0 + g1V z2 (V − σ)(V − τ) = 12(t− 1)− 4V (1− 3t)(4− 3t) . H. Prodinger: An online bin-packing problem with an underlying ternary structure 423 Altogether: G(u) = 6(t− 1)− 2V (1− 3t)(4− 3t) 1 V − 32 (1− t) = 6(1− t) + 2V (1− 3t)(4− 3t) 32 (1− t) 1 1− 23(1−t)V = 2 2 + 23(1−t)V (1− 3t)(4− 3t) 1 1− 23(1−t)V . Furthermore [V j ]G(u) = 2 (1− 3t)(4− 3t) [ 2 (2 3 1 1− t )j + (2 3 1 1− t )j] = 6 (1− 3t)(4− 3t) (2 3 1 1− t )j , j ≥ 1, and [uj ]G(u) = zj 6 (1− 3t)(4− 3t) (2 3 1 1− t )j . Now let us consider j + 3N steps to reach state j, and then [zj+3Nuj ]G(u) = [xN ] 6 (1− 3t)(4− 3t) (2 3 1 1− t )j = 1 2πi ∮ dx xN+1 6 (1− 3t)(4− 3t) (2 3 1 1− t )j = 27 4 ( 4 27 )N+1 1 2πi ∮ dt(1− t)(1− 3t) tN+1(1− t)2N+2 6 (1− 3t)(4− 3t) (2 3 1 1− t )j = ( 4 27 )N(2 3 )j 1 2πi ∮ dt tN+1(1− t)2N+j+1 6 (4− 3t) = ( 4 27 )N(2 3 )j−1 [tN ] 1 (1− t)2N+j+1 1 (1− 34 t) = ( 4 27 )N(2 3 )j−1 N∑ i=0 (3 4 )N−i(2N + j + i i ) = N∑ i=0 22i+j−1 32N+i+j−1 ( 2N + j + i i ) . The coefficients of g0 are different: [z3N ]g0 = N∑ i=0 22i 32N+i ( 2N + i i ) . Furthermore, [z3N+1]gβ = 1 3 [z3N ]g1 = N∑ i=0 22i 32N+i+1 ( 2N + 1 + i i ) . Here are the main results of this section: 424 Ars Math. Contemp. 24 (2024) #P3.03 / 415–425 Theorem 3.1. The generating function G(u) = G(u, z) has the following explicit form: G(u) = 2 2 + 23(1−t)V (1− 3t)(4− 3t) 1 1− 23(1−t)V . Here, u = Vz , z 3 = x = 274 t(1− t) 2. Written in the new variable V , only powers of z that are multiples of 3 appear. Further, we get the representation sorted by powers of u: [V j ]G(u) = 6 (1− 3t)(4− 3t) (2 3 1 1− t )j , j ≥ 1, and [uj ]G(u) = zj 6 (1− 3t)(4− 3t) (2 3 1 1− t )j . Reading off coefficients of zj+3Nuj leads to [zj+3Nuj ]G(u, z) = N∑ i=0 22i+j−1 32N+i+j−1 ( 2N + j + i i ) . For the special cases, the following series representation holds: [z3N ]g0 = N∑ i=0 22i 32N+i ( 2N + i i ) , [z3N+1]gβ = 1 3 [z3N ]g1 = N∑ i=0 22i 32N+i+1 ( 2N + 1 + i i ) . ORCID iDs Helmut Prodinger https://orcid.org/0000-0002-0009-8015 References [1] M. Drmota, Discrete random walks on one-sided “periodic” graphs, in: Discrete Random Walks, DRW’03. Proceedings of the conference, Paris, France, September 1–5, 2003, Maison de l’Informatique et des Mathématiques Discrètes (MIMD), Paris, pp. 83–94, 2003. [2] H. W. Gould, The Girard-Waring power sum formulas for symmetric functions and Fibonacci sequences, Fibonacci Q. 37 (1999), 135–140, https://www.fq.math.ca/37-2.html. [3] W. Knödel, Über das mittlere verhalten von online-packungsalgorithmen, Elektron. Informa- tionsverarb. Kybernetik 19 (1983), 427–433. [4] H. Prodinger, The kernel method: a collection of examples, Sémin. Lothar. Comb. 50 (2003), b50f, 19 pp. [5] H. Prodinger, Enumeration of S-Motzkin paths from left to right and from right to left: a kernel method approach, PU.M.A., Pure Math. Appl. 29 (2020), 28–38, doi:10.1515/puma-2015-0039, https://doi.org/10.1515/puma-2015-0039. H. Prodinger: An online bin-packing problem with an underlying ternary structure 425 [6] H. Prodinger, S. J. Selkirk and S. Wagner, On two subclasses of Motzkin paths and their relation to ternary trees, in: Algorithmic Combinatorics: Enumerative Combinatorics, Special Functions and Computer Algebra, Springer, Cham, pp. 297–316, 2020, doi:10.1007/978-3-030-44559-1 15, https://doi.org/10.1007/978-3-030-44559-1_15. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.04 / 427–452 https://doi.org/10.26493/1855-3974.2710.f3d (Also available at http://amc-journal.eu) Variants of the domination number for flower snarks Ryan Burdett, Michael Haythorpe * , Alex Newcombe Flinders University, 1284 South Road, Tonsley Park, SA, Australia Received 25 October 2021, accepted 14 July 2023, published online 8 July 2024 Abstract We consider the flower snarks, a widely studied infinite family of 3–regular graphs. For the Flower snark Jn on 4n vertices, it is trivial to show that the domination number of Jn is equal to n. However, results are more difficult to determine for variants of domination. The Roman domination, weakly convex domination, and convex domination numbers have been determined for flower snarks in previous works. We add to this literature by determining the independent domination, 2-domination, total domination, connected domination, upper domination, secure domination and weak Roman domination numbers for flower snarks. Keywords: Flower, snarks, domination, variants, secure. Math. Subj. Class. (2020): 05C69 1 Introduction Consider a graph G containing the vertex set V and edge set E. A subset of the vertices S ⊂ V is said to be a dominating set if every vertex in V \S is adjacent to at least one vertex in S. Then, the domination problem is to determine the size of the smallest dominating set in a given graph G, which is known as the domination number of G and is denoted by γ(G). There are obvious real-world applications for the domination problem. For example, suppose that the vertices of a graph correspond to locations in a secure site, and that each location needs to remain under observation by guards. If a guard at one location is able to simultaneously observe another, there is an edge between the corresponding vertices. Then, by placing guards at the sites corresponding to any dominating set, all locations are under observation. Clearly, it is desirable to do so with as few guards as possible. *Corresponding author. E-mail addresses: ryan.burdett@flinders.edu.au (Ryan Burdett), michael.haythorpe@flinders.edu.au (Michael Haythorpe), alex.newcombe@flinders.edu.au (Alex Newcombe) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 428 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 However, in many real-world applications, the definition of domination may be unsuit- able in some way, and so variants of domination have been described to handle such cases. Continuing the example in the previous paragraph, suppose that the guards carry out their observation from the top of large towers. The vantage point from these towers enables them to observe adjacent locations, but does not permit them to observe their own location. Then, their location would need to be observed by a guard at an adjacent location. In such a situation, we are seeking not a dominating set, but a total dominating set, which we define formally now along with several other variants of dominating sets that we will consider in this manuscript. Definition 1.1. Consider a dominating set S ⊆ V . Then: • S is a (weakly) convex dominating set if S is (weakly) convex in G. • S is an independent dominating set if S is an independent set in G. • S is a minimal dominating set if any proper subset of S is not a dominating set. • S is a 2-dominating set if any vertex v ̸∈ S is adjacent to at least two vertices in S. • S is a total dominating set if every vertex in V is adjacent to at least one vertex in S. • S is a connected dominating set if the subgraph of G induced by the vertices in S is connected. • S is a secure dominating set if, for every vertex v ∈ V \ S, there exists a vertex w ∈ S such that vw ∈ E, and (S \ {w}) ∪ {v} is a dominating set. Analogously to the domination number, we define γwcon(G) to be the weakly convex domination number, γcon(G) to be the convex domination number, i(G) to be the inde- pendent domination number, γ2(G) to be the 2-domination number, γt(G) to be the total domination number, γc(G) to be the connected domination number, and γs(G) to be the secure domination number. We also define the upper domination number, Γ(G), as follows. Definition 1.2. The upper domination number, denoted by Γ(G), is equal to the size of the largest minimal dominating set. In addition, we consider two more variants of domination. Suppose that for our graph G, we have a function f : V → {0, 1, 2}. Then, the weight of f , denoted w(f), is equal to∑ v∈V f(v). Definition 1.3. f is a Roman dominating function on G if it satisfies the condition that every vertex v for which f(v) = 0 is adjacent to at least one vertex w for which f(w) = 2. In the following definition, we say that a vertex v is undefended with respect to f if f(v) = 0 and for every vertex w adjacent to v, f(w) = 0. Definition 1.4. f is a weak Roman dominating function on G if, for every vertex v such that f(v) = 0, there is at least one vertex w, adjacent to v and satisfying the following: if we define a new function g : V → {0, 1, 2} defined by g(v) = 1, g(w) = f(w) − 1, and g(u) = f(u) for all u ̸= {v, w}, then there are no undefended vertices with respect to g. R. Burdett et al.: Variants of the domination number for flower snarks 429 Then, the Roman domination number γR(G) is the minimum weight over all Roman dominating functions, and equivalently for the weak Roman domination number γr(G). It is worth noting that if we further demand that f(v) ≤ 1 for all v ∈ V then weak Roman domination is equivalent to secure domination. Hence, it is clear that γr(G) ≤ γs(G). Domination numbers are known for some infinite families of graphs. Other than trivial results such as for paths or cycles, perhaps the most famous result is the sprawling effort over a 27 year period [2, 8, 14, 16, 20, 36] to provide a complete characterisation of domi- nation numbers for grid graphs G(n,m) of all possible sizes, consisting of 23 special cases before settling into a standard formula for n,m ≥ 16. Other results for domination include generalized Petersen graphs [13, 24, 41], Cartesian products involving cycles [1, 9, 28], King graphs [40], Latin square graphs [29], hypercubes [3], Sierpiński graphs [34], Knödel graphs [12], and various graphs from chemistry [26, 32], among others. However, far fewer results are known for variants of domination beyond trivial results such as for paths, cycles, stars, wheels, or complete graphs. We summarise the most note- worthy of these results for the variants of domination considered in this paper. For upper domination, results are known for various graphs based on chessboards [4, 11, 18, 39, 40]. For total domination, results are known for some grid graphs G(n,m) (for n ≤ 6) [15, 21], Knödel graphs [27], and various graphs from chemistry [26]. For connected domination, results are known for trees [33], circulant graphs [35], Centipede graphs [38], and various graphs from chemistry [26]. For weak Roman domination, results are known for various graphs based on chessboards [30], Cartesian products involving complete graphs [37], and Helm graphs and Web graphs [23]. For secure domination, results are known for some grid and torus graphs (for n ≤ 3) [17], various Cartesian products of stars, cycles, paths and complete graphs [37], and middle graphs [31]. In recent works, the following results were shown for flower snarks (which will be defined in the next section). Theorem 1.5 ([25, Maksimovic et al. (2018)], [22, Kratica et al. (2020)]). Consider the flower snark Jn, for n ≥ 3. Then we have γR(Jn) = γwcon(Jn) = 2n, and γcon(Jn) = 4n. We now add to the above literature by proving the following additional results for flower snarks: Theorem 1.6. Consider the flower snark Jn, for n ≥ 3. Then, γ(Jn) = i(Jn) = n, γs(Jn) = γr(Jn) = ⌈ 3n+ 1 2 ⌉ , γ2(Jn) = {⌈ 5n 3 ⌉ , if n ̸= 1 mod 3, 5n+4 3 , if n = 1 mod 3, γt(Jn) = {⌈ 3n 2 ⌉ , if n ̸= 2 mod 4, 3n 2 + 1, if n = 2 mod 4, γc(Jn) = Γ(Jn) = { 2n, if n is even, 2n− 1, if n is odd. In most cases, the proofs will be by induction, and hence it will be necessary to first prove the results for some number of base cases. Rather than provide these proofs here, we will simply use mixed-integer linear programming formulations of each variant of domina- tion from literature to handle the base cases. 430 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 2 Flower snarks The chromatic index of a graph is the minimum required number of colours to color the edges of the graph, such that no two incident edges have the same colour. By Vizing’s theorem, it is known that all 3–regular graphs have chromatic index 3 or 4. The latter case is rare, and simple, connected, bridgeless 3–regular graphs with chromatic index 4 are called snarks. It is common to further add the restriction that the girth should be at least 5, with such graphs known as nontrivial snarks. Flower snarks [19], discovered by Isaacs in 1975, were the first known infinite family of nontrivial snarks, and are denoted by Jn. They are defined as follows. Definition 2.1 (Flower snarks). For n ≥ 3, take the union of n copies of K1,3. Denote the degree 3 vertex in the i-th copy as ai, and the other three vertices in the i-th copy as bi, ci and di. Then construct an n-cycle through vertices b1, b2, . . . , bn, and a 2n-cycle through vertices c1, c2, . . . , cn, d1, d2, . . . , dn. In order to have the properties of a nontrivial snark, n must be odd and n ≥ 5. However, for other values of n ≥ 3 a 3–regular graph is nonetheless obtained by this construction. In this paper we will consider all n ≥ 3, and for convenience we will refer to all of them as flower snarks. For the remainder of this document, we will consider various kinds of dominating sets of flower snarks. As such, it is convenient to go over some brief terms and notation here. Recall that Jn contains n copies of K1,3. We will refer to the i-th copy as J i, and its four vertices as ai, bi, ci and di. Note that this notation does not include n, as typically n will be fixed in our considerations. Also, we will say that a copy J i has weight k in a dominating set S, if S contains k vertices from J i. We will use the term pattern to describe a sequence of weights of consecutive copies of Jn in S. For example, if we say that S contains the pattern 121, it means there is a set of three consecutive copies J i−1, J i, J i+1 which have weights 1, 2 and 1 respectively in S. Further to this, in a set of consecutive copies of Jn, we will use the term configuration to describe the specific allocation of its vertices to S. Finally, we will define wSi to be equal to the number of copies with weight i in S. Flower snarks can be visualised in various ways. A common method is to distribute the copies of K1,3 in a circle, using curved edges for the 2n-cycle and straight edges elsewhere. We display one such drawing of Jn in part (a) of Figure 1, for n = 7. However, for our purposes it will be convenient to focus only on a small section of a flower snark at a time, and so we will use the drawing style displayed in part (b) of Figure 1, with each copy displayed vertically. As indicated in Figure 1 we will assume that in these drawings, the bottom vertex in copy J i is bi, followed by ai, ci and di. When useful to avoid confusion, a label will be given above each copy. It is worth noting that, when viewing only a section of Jn, vertices bi, ci and di are all essentially equivalent, and this will be useful in simplifying many of the upcoming proofs. In a global sense this is not the case, as there is a “twist” in the final copy in which cn links to d1, and dn links to c1. However, due to the symmetry of the flower snark, when viewing any copy locally we can choose to relabel the vertices so that this twist occurs elsewhere in the graph. Hence, in the arguments that follow, whenever we are viewing only a portion of the graph we will always assume that the twist occurs elsewhere in the graph. Various proofs in this paper will go as follows. We will begin with a set of copies of Jn for which the pattern is known, as well as possibly knowing in advance that some vertices are either in S, or not in S. From there, depending on the variant of domination being R. Burdett et al.: Variants of the domination number for flower snarks 431 (a) a b c d J1 J2 J3 J4 J5 J6 (b) Figure 1: In part (a) a common drawing of the flower snark J7. In part (b), a section of a larger flower snark consisting of six copies. considered, and the structure of Jn, we will go on to prove that certain vertices either must be in S, or must not be in S. For these proofs, figures will be provided, using the following convention. The pattern known in advance will be indicated by listing the corresponding weight underneath each copy. Vertices which are known in advance to be in S will be marked with , and vertices which are known in advance not to be in S will be marked with . Then, vertices which are subsequently shown (up to equivalence) to be in S will be marked with , while vertices which are subsequently shown not to be in S will be marked with . We demonstrate this convention with a simple example. Example 2.2. Suppose we have four copies J1, J2, J3, J4 which meet the pattern 1111, and that we know that d2 ∈ S and c1 ̸∈ S. This situation is displayed in Figure 2. Clearly, since d2 ∈ S and J2 has weight 1, we know that a2 ̸∈ S, b2 ̸∈ S, and c2 ̸∈ S. These three are marked with in Figure 2 as no argument was needed to establish they are not in S. Then, the only remaining vertex which can dominate c2 is c3, and hence c3 ∈ S. Since this was argued in the proof, c3 is marked with a in Figure 2. Also, since J3 has weight 1 the vertices a3, b3, and d3 cannot be in S, and so they are marked with . Then, the only remaining vertex which can dominate b2 is b1, which we similarly mark in Figure 2. Finally, the only remaining vertex which can dominate b3 is b4, which we again mark in Figure 2. Hence, we now know exactly which vertices in J1, J2, J3, J4 are contained in S. a b c d J1 J2 J3 J4 1 1 1 1 Figure 2: The situation described in Example 2.2. To conclude this section, we note that it is trivial to determine the domination and independent domination numbers for flower snarks. Lemma 2.3. For n ≥ 3, we have γ(Jn) = i(Jn) = n. Proof. The graph Jn contains n copies of K1,3. For each copy J i, the vertex ai is ad- jacent only to other vertices in J i, and so any dominating set must contain at least one 432 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 vertex from each copy. Hence, n ≤ γ(Jn) ≤ i(Jn). Then, it suffices to note that the set {a1, a2, . . . , an} is an independent dominating set, and hence i(Jn) ≤ n, leading to the result. Corollary 2.4. For any variant of dominating set considered in this paper, each copy must have weight at least 1. 3 Upper domination In this section, we will determine the upper domination number for flower snarks. We begin by considering what weights are possible for copies in a minimal dominating set. Lemma 3.1. Consider the graph Jn for n ≥ 3. If S is a minimal dominating set for Jn then the weight of each copy is either 1, 2 or 3. Proof. We know from Corollary 2.4 that each copy must have positive weight. Then sup- pose that a copy J i has weight 4. It can be easily checked that S \{ai} is also a dominating set, contradicting the assumption that S is minimal. In the following lemma, we use the term i depends on j to imply that S ∩N [i] = {j}. Note that since S is minimal, for every vertex in S there must be at least one other vertex which depends on it. Lemma 3.2. Consider the graph Jn for n ≥ 4, and a minimal dominating set S. If a copy J i has weight 3 in S, then both adjacent copies J i−1 and J i+1 have weight 1 in S. Proof. Suppose it is not the case, that is, J i has weight 3 and at least one of J i−1 and J i+1 has weight greater than 1. Since bi, ci and di are all equivalent in this framing, there are only two cases to consider; if ai ∈ S and if ai ̸∈ S. First, consider the case when ai ∈ S. Then there are two other vertices from J i also in S. Without loss of generality, suppose bi ∈ S and ci ∈ S. This situation is shown in Figure 3 part (a). Then, since S is minimal, the removal of ai does not result in a dominating set. This is only possible if di depends on ai. Hence, di−1 ̸∈ S and di+1 ̸∈ S. Then, consider bi. Since S is minimal, at least one of bi−1 and bi+1 must depend on bi. Without loss of generality, suppose it is the former. This implies that both bi−1 ̸∈ S and ai−1 ̸∈ S. Then, ci−1 ∈ S, or else copy Ji−1 has weight 0 which contradicts Lemma 3.1. Then, since S is minimal, ci+1 must depend on ci, which implies that ci+1 ̸∈ S and ai+1 ̸∈ S. Hence, from Lemma 3.1 copy J i+1 also has weight 1, contradicting the initial assumption. Second, consider the case when ai ̸∈ S. Then, S contains bi, ci, and di. Since S is minimal and ai does not depend on bi, at least one of bi−1 or bi+1 must depend on bi. Without loss of generality, suppose it is the former. This implies that both bi−1 ̸∈ S and ai−1 ̸∈ S. We can make analogous arguments for ci and di. Clearly, the dependent vertices can not all be from the same copy, or else that copy has weight 0. Hence, two of the dependent vertices are in one copy and one is in the other; without loss of generality we will assume that bi−1 is dependent on bi, ci−1 is dependent on ci, and di+1 is dependent on di. Hence, J i−1 has weight 1, and so by assumption, J i+1 has weight 2, and bi+1 ∈ S and ci+1 ∈ S. But then, by similar arguments, it must be the case that bi+2 depends on bi+1, and ci+2 depends on ci+1, implying that J i+2 has weight 1, and di+2 ∈ S. This situation is displayed in Figure 3 part (b). Finally, it can be seen from this figure that S \ {di} is a dominating set, contradicting the initial assumption that S is minimal. R. Burdett et al.: Variants of the domination number for flower snarks 433 a b c d (a) a b c d (b) Figure 3: The two situations described in the proof of Lemma 3.2. Recalling that wSi is the number of copies to have weight i in S, Lemma 3.2 implies the following. Corollary 3.3. Consider the graph Jn for n ≥ 4. If S is a minimal dominating set for Jn then wS1 ≥ wS3 , with equality occurring if and only if wS1 = wS3 = n2 . We are now ready to prove the main result of this section. Theorem 3.4. Consider a graph Jn for n ≥ 3. Then, Γ(Jn) = { 2n, if n is even, 2n− 1, if n is odd. Proof. We use the first formulation for upper domination from [6] to confirm that Γ(J3) = 5. Then, suppose that S is a minimal dominating set for Jn, for n ≥ 4. From Lemma 3.1 we have wS0 = w S 4 = 0. Hence, we have w S 1 +w S 2 +w S 3 = n, and w S 1 +2w S 2 +3w S 3 = |S|. Combining these, we obtain |S| = 2n + wS3 − wS1 . From Corollary 3.3 we know that wS1 ≥ wS3 , and the inequality is strict if n is odd. Hence, |S| ≤ 2n if n is even, and |S| ≤ 2n− 1 if n is odd. Then we just need to obtain the corresponding lower bounds. Suppose that n is even. It is easy to check that if we repeat the configuration displayed in Figure 4 part (a) n/2 times, what results is a minimal dominating set with weight 2n. Hence, if n is even, we have Γ(Jn) ≥ 2n. Then suppose that n is odd. Again, it is easy to check that if we repeat the configuration displayed in Figure 4 part (a) (n− 1)/2 times, and then use the configuration displayed in Figure 4 part (b) for the final copy, what results is a minimal dominating set with weight 2n− 1. Hence, if n is odd, Γ(Jn) ≥ 2n− 1. a b c d (a) a b c d (b) Figure 4: The configuration for upper domination which gives the desired lower bound for Γ(Jn) for n ≥ 3. Part (a) can be repeated as many times as necessary. Then, if n is odd, use part (b) to finish. The result is a minimal dominating set with weight 2n if n is even, or weight 2n− 1 if n is odd. 434 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 4 Upper bounds by construction In the upcoming sections, we will determine lower bounds for the 2-domination, total dom- ination, connected domination, secure domination, and weak Roman domination numbers of flower snarks. To obtain equality, we will require corresponding upper bounds, which we provide here. We leave it as an exercise to the reader to verify that the configurations given here satisfy the requirements of the various kinds of domination, and that they imply the appropriate upper bounds for Theorem 1.6. For 2-domination, the configuration shown in Figure 5 can be repeated as often as necessary, truncating the final time to get the desired size. a b c d Figure 5: The configuration for 2-domination which gives the desired upper bound for γ2(Jn). The result is a 2-dominating set of weight ⌈ 5n3 ⌉ if n ̸= 1 mod 3, or weight ⌈ 5n 3 ⌉+1 if n = 1 mod 3. For total domination, the configuration shown in Figure 6 part (a) can be repeated as often as necessary. Then the configurations in parts (b), (c) and (d) can be used to complete the remaining copies. a b c d (a) a b c d (b) a b c d (c) a b c d (d) Figure 6: The configuration for total domination which gives the desired upper bound for γt(Jn) for n ≥ 3. Part (a) can be repeated as many times as necessary. If n = 0 mod 4 this is sufficient. If n = 1 mod 4, use (b) to finish. If n = 2 mod 4, use part (c) to finish. If n = 3 mod 4, use part (d) to finish. The result is a total dominating set of weight ⌈ 3n2 ⌉ if n ̸= 2 mod 4, or weight 3n2 + 1 if n = 2 mod 4. For connected domination, the configuration shown in Figure 7 can be repeated as often as necessary, truncating the final time to get the desired size. If n ̸= 1 mod 4 the result is connected dominating. If n = 1 mod 4, then removing d1 and adding b1 gives the desired result. For secure domination, the configuration shown in Figure 8 part (a) can be repeated as often as necessary. Then the configurations in parts (b), (c), (d), and (e) can be used to com- plete the remaining copies. Since secure domination is an upper bound for weak Roman domination, these configurations also give an upper bound for weak Roman domination. R. Burdett et al.: Variants of the domination number for flower snarks 435 a b c d Figure 7: The configuration for connected domination which gives the desired upper bound for γc(Jn) for n ≥ 3. It may be repeated as many times as necessary, truncating the final time to get the final size. If n = 1 mod 4, then in the first copy, remove d1 and add b1. The result is a connected dominating set of weight 2n if n is even, or 2n− 1 if n is odd. a b c d (a) a b c d (c) a b c d (d) a b c d (b) a b c d (e) Figure 8: The configuration for secure domination which gives the desired upper bound for γs(Jn) for n ≥ 4. Part (a) can be repeated as many times as necessary. If n = 0 mod 4, use part (b) to finish. If n = 1 mod 4, use part (c) to finish. If n = 2 mod 4, use part (d) to finish. If n = 3 mod 4, use part (e) to finish. In all cases, the result is a secure dominating set of weight ⌈ 3n+12 ⌉. 5 2-domination In this section, we will determine the 2-domination numbers for flower snarks. Lemma 5.1. Consider the graph Jn for n ≥ 3, and a 2-dominating set S. If the copy J i has weight 1 in S, then ai ∈ S. Proof. Recall that ai is adjacent only to other vertices in J i. From the definition of 2- domination, if ai ̸∈ S then it must have at least two neighbours in S. Since J i has weight 1, this is impossible, and so ai ∈ S. Theorem 5.2. Consider the graph Jn for n ≥ 3, and a 2-dominating set S. Then any copy with weight 1 in S has a neighbouring copy with weight 3 or 4 in S. Proof. Without loss of generality, suppose that J2 has weight 1. Then, from Lemma 5.1 we have a2 ∈ S. Then, suppose that its neighbours J1 and J3 both have weight less than 3. At this stage, vertices b2, c2 and d2 have only one neighbour in S, so they each need at least one more. Without loss of generality, suppose that b1 ∈ S. Then, consider the case when c1 ∈ S. Since J1 has weight less than 3, this implies that a1 ̸∈ S and d1 ̸∈ S. However, it is then impossible for d1 to have two neighbours in S. This situation is displayed in part (a) of Figure 9. 436 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 a b c d J1 J2 (a) a b c d J1 J2 J3 (b) Figure 9: The two situations described in the proof of Theorem 5.2. In part (a), d1 cannot have two neighbours in S. In part (b), copy b3 cannot have two neighbours in S. Hence, we must have c1 ̸∈ S. Then, c1 must have two neighbours in S, which implies that a1 ∈ S. Since J1 has weight less than 3, this implies that d1 ̸∈ S. Then, in order for c2 and d2 to have two neighbours in S, we must have c3 ∈ S and d3 ∈ S, respectively. Since J3 has weight less than 3, this implies that a3 ̸∈ S and b3 ̸∈ S. However, it is then impossible for b3 to have two neighbours in S, completing the proof. This situation is displayed in part (b) of Figure 9. We are now ready to prove the main result of this section. Theorem 5.3. Consider the graph Jn for n ≥ 3. Then, γ2(Jn) = {⌈ 5n 3 ⌉ , if n ̸= 1 mod 3, 5n+4 3 , if n = 1 mod 3. Proof. The upper bound was established in Section 4. Then, from Corollary 2.4 we know that each copy has weight at least 1. This, combined with Theorem 5.2, implies that any three set of consecutive copies have weight at least 5. Hence we have γ2(S) ≥ ⌈ 5n3 ⌉. Hence, Theorem 5.3 is true for n ̸= 1 mod 3. Suppose that n = 1 mod 3, and we have a 2-dominating set S such that |S| < 5n+43 . Hence, we must have |S| = 5n+13 . For each copy J i, denote by w3(i) the combined weights of J i, J i+1, and J i+2, where the superscripts are taken modulo n. Recall that w3(i) ≥ 5. Hence there must be a single value k such that w3(k) = 6 and w3(j) = 5 for all j ̸= k. Consider the copies Jk, Jk+1, Jk+2. We will consider the possible patterns they could have, and show that each is impossible. Since w3(k) has weight 6, the only possible patterns (up to symmetry) for copies Jk, Jk+1 and Jk+2 are 123, 132, 222, and 312. Suppose that Jk, Jk+1 and Jk+2 meet either the pattern 123 or the pattern 132. Since w3(k + 1) = 5, this implies that Jk+3 has weight 0, which contradicts Corollary 2.4. Suppose next that Jk, Jk+1 and Jk+2 meet the pattern 222. Since w3(k + 1) = w3(k+2) = 5, this implies that Jk+3 has weight 1, and Jk+4 has weight 2. However, this contradicts Theorem 5.2 since Jk+3 has no neighbour with weight 3 or 4. Hence, this is impossible. Finally, suppose that Jk, Jk+1 and Jk+2 meet pattern 312. Since w3(k+1) = w3(k+ 2) = w3(k+3) = 5, this implies that Jk+3 has weight 2, Jk+4 has weight 1, and Jk+5 has weight 2. Again, this contradicts Theorem 5.2 since Jk+4 has no neighbour with weight 3 or 4. Hence, all cases are impossible, and so |S| ≥ 5n+43 , completing the proof. R. Burdett et al.: Variants of the domination number for flower snarks 437 6 Total domination In this section, we will determine the total domination numbers for flower snarks. We begin by identifying three patterns which cannot occur in total dominating sets of Jn. Theorem 6.1. Consider the graph Jn for n ≥ 3, and a total dominating set S. Then S does not contain the patterns 111, 1121, or 12121. Proof. From Corollary 2.4, each copy of Jn has weight at least 1 in S. Also, if a copy J i has weight 1, then ai ̸∈ S because otherwise ai itself is not dominated. Suppose that S has the pattern 111. Without loss of generality, suppose the three copies meeting this pattern are J1, J2, J3. As indicated above, a2 ̸∈ S. Regardless of which vertex from J2 is in S, it does not dominate any of b2, c2, d2. Also, any vertex from J1 dominates at most one vertex from J2, and likewise for any vertex from J3. Hence there is at least one vertex in J2 which is not dominated, contradicting the assumption that S is a total dominating set. Then, suppose that S has the pattern 1121. Without loss of generality, suppose the four copies meeting this pattern are J1, J2, J3, J4. Using an equivalent argument to above, it must be the case that b2, c2, d2 are dominated by one vertex from J1 and two vertices from J3. Hence, a3 ̸∈ S. This means none of the vertices from J3 dominate any of b3, c3, d3. Also, any vertex from J2 dominates at most one vertex from J3, and likewise for any vertex from J4. Hence there is at least one vertex in J3 which is not dominated, contradicting the assumption that S is a total dominating set. Finally, suppose that S has the pattern 12121. Without loss of generality, suppose the five copies meeting this pattern are J1, . . . , J5. Since J1 and J3 are both weight 1, they can collectively dominate at most two vertices from J2. Hence, a2 ∈ S, and by an equivalent argument, a4 ∈ S. Note that none of the vertices from J3 can dominate any of b3, c3 or d3. Also, the vertices from J2 dominate at most one vertex from J3, and likewise for the vertices from J4. Hence there is at least one vertex in J3 which is not dominated, contradicting the assumption that S is a total dominating set. An immediate observation arising from Corollary 2.4 and Theorem 6.1 is that any four consecutive copies must collectively have weight at least 6 in S, leading to the following corollary. Corollary 6.2. For n ≥ 4, γt(Jn) ≥ ⌈ 3n 2 ⌉ . We are now ready to prove the main result of this section. Theorem 6.3. Consider the graph Jn for n ≥ 3. Then, γt(Jn) = {⌈ 3n 2 ⌉ , if n ̸= 2 mod 4, 3n 2 + 1, if n = 2 mod 4. Proof. The upper bound was established in Section 4. We use the formulation for total domination from [7] to confirm that γt(J3) = 5. Then, suppose there is some value k ≥ 4 such that Theorem 6.3 is false. If k ̸= 2 mod 4 then we have γt(Jk) ≤ ⌈ 3k 2 ⌉ − 1, contradicting Corollary 6.2. Hence, we must have k = 2 mod 4. Hence, γt(Jk) ≤ 3k2 , and from Corollary 6.2 this implies that γt(Jk) = 3k2 . 438 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 Suppose that we have a total dominating set S with weight 3k2 . Note that this implies that every set of four consecutive copies of Jn has weight 6 in S; call this property 1. It is clear that property 1 implies that no copy has weight 4 in S. If a copy has weight 3 in S, then in order to satisfy property 1, the next three copies must have weight 1, which contradicts Theorem 6.1. Hence, every copy has weight 1 or 2 in S; call this property 2. It can be easily checked that there are only two ways to satisfy properties 1 and 2 without contradicting Theorem 6.1; either S contains the repeated pattern 1212 . . . 12, or S contains the repeated pattern 11221122 . . . 1122. In the former case, S contains the pattern 12121 which by Theorem 6.1 is impossible. Hence, we must have the latter case. Also, the latter case is impossible because k = 2 mod 4. Hence, all cases are impossible, completing the proof. 7 Connected domination In this section, we will determine the connected domination numbers for flower snarks. The following three remarks are obvious, but we list them here to aid the readability of two upcoming proofs. Remark 7.1. Suppose that S is a connected dominating set for a graph G. If a new graph H is created by either (a) adding an edge between two non-adjacent vertices in G, or (b) deleting a vertex v ̸∈ S from G, then S is also a connected dominating set for H . Remark 7.2. Suppose that S is a connected dominating set for a graph G, and that G contains a degree 1 vertex v ∈ S whose neighbour has degree larger than 1. If a new graph H is created by deleting v from G, then S \ {v} is a connected dominating set for H . Remark 7.3. Suppose that S is a connected dominating set for a graph G, and that G contains a triangle uvw such that v is degree 2, and v ∈ S. If a new graph H is created by deleting v from G, then S \ {v} is a connected dominating set for H . In the following, we define V S(J i) := S ∩ (ai, bi, ci, di), that is, V S(J i) is the set of vertices from J i that are contained in S. Theorem 7.4. Consider the graph Jn for n ≥ 4, and let S be a connected dominating set for Jn. Suppose that there is a copy J i such that either ai ̸∈ S, or J i has weight 2 in S. Then S := S \ V S(J i) is a connected dominating set for Jn−1. Proof. Consider first the situation when ai ̸∈ S, and consider the graph Jn−1. One may think of Jn−1 as being constructed by starting with Jn and then “smoothing out” copy J i, in the following sense. First, we add edges connecting bi−1 to bi+1, ci−1 to ci+1 and di−1 to di+1, and then we delete the vertex set V (J i). Suppose that we are midway through this process, having added all three edges, and having deleted ai. Call this intermediate graph G, and note from Remark 7.1 that S is a connected dominating set for G. Then, note that in G, bi is a degree 2 vertex and is part of a triangle bi−1bibi+1. Now, suppose that we delete bi from G, to obtain a new intermediate graph G2. If bi ̸∈ S then from Remark 7.1 we can see that S is a connected dominating set for G2. If bi ∈ S then from Remark 7.3 we can see that S \ {bi} is a connected dominating set for G2. Applying analogous arguments for ci and di, we obtain the result. The graphs G and G2 are displayed in Figure 10. We next consider the situation when J i has weight 2 in S. If ai ̸∈ S then the previous paragraph applies. If ai ∈ S then there must be one more vertex from J i also in S; without R. Burdett et al.: Variants of the domination number for flower snarks 439 loss of generality, suppose that bi ∈ S. Again, we construct Jn−1 by adding edges to and deleting vertices from Jn. Suppose that we are midway through this process, having added all three edges, and having deleted ci and di. Call this intermediate graph G3. From Remark 7.1 we know that S is a connected dominating set for G3. Note that in G3, ai is a degree 1 vertex. Remark 7.2 implies that we can obtain a second intermediate graph, G4, by deleting ai, and that S \ {ai} is a connected dominating set for G4. Finally, note that in G4, bi is degree 2 vertex and is part of a triangle bi−1bibi+1. Hence, Remark 7.3 implies the result. The graphs G3 and G4 are displayed in Figure 10. a b c d G a b c d G2 a b c d G3 a b c d G4 Figure 10: Sections of the four intermediate graphs G, G2, G3, G4 from the proof of Theorem 7.4. We also make use of the concept of “smoothing out” a copy in the following lemma. Lemma 7.5. Consider the graph Jn for n ≥ 4, and let S be a connected dominating set for Jn. Suppose that there is a copy J i with weight 4 in S. Then for any positive integer k ≤ n−3, we have that S := S \(V S(J i+1)∪ . . .∪V S(J i+k)) is a connected dominating set for Jn−k. Proof. Suppose that we delete from Jn all vertices from each of the copies J i+1, . . . , J i+k. Call this intermediate graph G. Then, we add edges connecting bi to bi+k+1, ci to ci+k+1 and di to di+k+1, and call the resulting graph G2. Since we can always relabel the vertices so that the twist occurs elsewhere in the graph, it can then be seen that G2 is isomorphic to Jn−k. Both G and G2 are displayed in Figure 11. Since bi ∈ S, ci ∈ S, and di ∈ S, it is clear that vertices bi+k+1, ci+k+1 and di+k+1 are dominated in G2. Also, from Corol- lary 2.4 we know that J i+k+1 has weight at least 1, and so ai+k+1 is dominated. Hence, S is a dominating set for G2. Next, we need to show that S is a connected dominating set for G2. Recall the intermediate graph G. It is clear that the subgraph of G induced by S is either connected, or has exactly two connected components. In the former case, the subgraph of G2 induced by S is also connected. In the latter case, it is clear that there must be an edge vw in G2 such that v ∈ J i, w ∈ J i+k+1 and both v, w are contained in S, and hence the subgraph of G2 induced by S is connected. Either way, we obtain the desired result. Lemma 7.6. Suppose that there is an odd value n ≥ 5 such that γc(Jn−1) = 2n− 2. If S is a connected dominating set for Jn such that |S| = 2n− 1, then wS4 is even. Proof. Suppose that wS2 > 0. Then there is a copy J i with weight 2. From Theorem 7.4, we can obtain a connected dominating set for Jn−1 of cardinality 2n− 3, contradicting the initial assumption. Hence, wS2 = 0. Then, we have w S 1 +w S 3 +w S 4 = n, and w S 1 +3w S 3 + 4wS4 = 2n− 1. Combining these two, we obtain 3wS4 = n− 1− 2wS3 . Since n is odd, this implies that wS4 is even. 440 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 a b c d J i . . . J i+k+1 G a b c d J i . . . J i+k+1 G2 Figure 11: Sections of the two intermediate graphs G and G2 from the proof of Lemma 7.5. Theorem 7.7. Consider the graph Jn for n ≥ 3, and suppose that S is a connected dominating set. Then S does not contain the patterns 111 or 1312131. Proof. Any connected dominating set with |S| ≥ 2 is also a total dominating set, and hence the result for pattern 111 follows immediately from Theorem 6.1. Next, consider the case when S contains the pattern 1312131. Without loss of gener- ality, suppose the copies meeting this pattern are J1, . . . , J7. Suppose that a4 ̸∈ S. Then, without loss of generality, suppose that b4 ∈ S, c4 ∈ S, and d4 ̸∈ S. This situation is displayed in part (a) of Figure 12. Since S is dominating, it must contain at least one of d3 and d5. Since both J3 and J5 have weight 1, both options are equivalent, so without loss of generality we will assume that d5 ∈ S. Then a5 ̸∈ S, b5 ̸∈ S, and c5 ̸∈ S. In order for S to be connected, we must have b3 ∈ S and c3 ∈ S, but this is impossible since J3 has weight 1. Hence, it must be the case that a4 ∈ S. Again, without loss of generality, suppose that b4 ∈ S, c4 ̸∈ S and d4 ̸∈ S. This situation is displayed in part (b) of Figure 12. Since S is connected, it must contain at least one b3 or b5. As in the previous paragraph, without loss of generality we will assume that b5 ∈ S. Then, in order for S to be dominating, we must have c6 ∈ S and d6 ∈ S. Now, suppose that b6 ∈ S. Then, since S is connected, it must contain both c7 and d7, but this is impossible since J7 has weight 1. Hence, it must be the case that b6 ̸∈ S, and hence a6 ∈ S. Then, since S is connected, we must have b3 ∈ S and b2 ∈ S. In order for S to be dominating, we must have c2 ∈ S and d2 ∈ S, and since J2 has weight 3 this implies that a2 ̸∈ S. Finally, to ensure S is connected, it must contain each of b1, c1 and d1, but this is impossible since J1 has weight 1. a b c d J3 J4 J5 1 2 1 (a) a b c d J1 J2 J3 J4 J5 J6 J7 1 3 1 2 1 3 1 (b) Figure 12: The two situations described in the proof of Theorem 7.7. In part (a), J3 has too many vertices in S. In part (b), J1 has too many vertices in S. R. Burdett et al.: Variants of the domination number for flower snarks 441 Theorem 7.8. Consider the graph Jn for even n ≥ 6, and a connected dominating set S for Jn such that |S| = 2n−1. If γc(Jn−1) = 2n−3, and wS4 = 0, then S does not contain the pattern 112. Proof. Without loss of generality, suppose that one set of copies meeting the pattern 112 is J2, J3, J4, and consider also J1 and J5. Since J2 and J3 have weight 1, we know that a2 ̸∈ S and a3 ̸∈ S. Without loss of generality, suppose that b2 ∈ S. Then, suppose that b3 ∈ S as well. This situation is displayed in part (a) of Figure 13. Since S is dominating, it must also contain each of c1, d1, c4, and d4. Since J4 has weight 2, we have a4 ̸∈ S and b4 ̸∈ S. Since S is connected, it implies that b1 ∈ S. Since there are no copies of weight 4 in S, this implies that a1 ̸∈ S. However, by Theorem 7.4, we can then obtain a connected dominating set for Jn−1 of cardinality 2n−4, contradicting the assumption that γc(Jn−1) = 2n− 3. Hence, the assumption that b3 ∈ S must be false. We instead have b2 ∈ S and b3 ̸∈ S. Without loss of generality, suppose that c3 ∈ S. This situation is displayed in part (b) of Figure 13. Since S is dominating, it must also contain d4. Since S is connected, we have c4 ∈ S, and since J4 has weight 2, we have a4 ̸∈ S and b4 ̸∈ S. Then, since S is dominating, we have b5 ∈ S, and since S is connected we have c5 ∈ S and d5 ∈ S. Since there are no copies of weight 4 in S, it implies that a5 ̸∈ S. However, by Theorem 7.4, we can the obtain a connected dominating set for Jn−1 of cardinality 2n − 4, which again contradicts the assumption that γc(Jn−1) = 2n − 3, completing the proof. a b c d J1 J2 J3 J4 J5 1 1 2 (a) a b c d J1 J2 J3 J4 J5 1 1 2 (b) Figure 13: The two situations described in the proof of Theorem 7.8. In both parts, there is a copy with weight 3, but with the a vertex not contained in S. In the next theorem, we will require the following concept. Suppose that we have a connected dominating set S for Jn. Denote by Jn(S) the subgraph of Jn induced by S. By definition, Jn(S) is connected. Then, consider a set of consecutive copies J i, . . . , J i+k, and define S = S ∩ (V (J i) ∪ . . . ∪ V (J i+k)). If Jn(S) is a disconnected graph, then we say that the copies J i, . . . , J i+k are locally disconnected in S. Since S itself is connected, it is clear that if S contain two sets of locally disconnected consecutive copies, they must overlap by two or more copies. Theorem 7.9. Consider the graph Jn for even n ≥ 6, and a connected dominating set S for Jn such that |S| = 2n − 1. If γc(Jn−1) = 2n − 3, then any set of consecutive copies meeting the patterns 113 or 3123 are locally disconnected. Proof. Consider first the pattern 113. Without loss of generality, suppose that one set of copies meeting this pattern is J1, J2, J3, and that this set of copies is not locally discon- nected. Suppose that a3 ̸∈ S. Then, from Theorem 7.4 there exists a connected dom- inating set for Jn−1 with cardinality 2n − 4, which is impossible since by assumption 442 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 γc(Jn−1) = 2n−3. Hence, a3 ∈ S. Then, without loss of generality, suppose that b3 ∈ S, c3 ∈ S, and d3 ̸∈ S. This situation is displayed in part (a) of Figure 14. Since J2 has weight 1, it is clear that a2 ̸∈ S. Then, in order for S to be dominating, it must contain at least one of d1 and d2. However, if d2 ∈ S, then S does not contain b2 or c2, and hence J1, J2, J3 are locally disconnected, contradicting the initial assumption. Hence, d2 ̸∈ S, and so d1 ∈ S. But then, because J1 has weight 1, S does not contain a1, b1, or c1, which again implies that J1, J2, J3 are locally disconnected, contradicting the initial assumption. Hence, the initial assumption must be false, and J1, J2, J3 are locally disconnected. Next, consider the pattern 3123. Without loss of generality, suppose that one set of copies meeting this pattern is J1, J2, J3, J4, and that this set of copies if not locally dis- connected. Using an identical argument as from the previous paragraph, we must have a1 ∈ S and a4 ∈ S. Then, without loss of generality, suppose that b1 ∈ S and c1 ∈ S. This situation is displayed in part (b) of Figure 14. Since J2 has weight 1, and the set of copies is not locally disconnected, we have a2 ̸∈ S and d2 ̸∈ S. The remaining two choices are equivalent; without loss of generality, suppose that b2 ∈ S. Then, since S is dominating, and the set of copies is not locally disconnected, we must have b3 ∈ S and d3 ∈ S. Finally, since S is dominating, and the set of copies is not locally discon- nected, we must have b4 ∈ S, c4 ∈ S and d4 ∈ S. However, this is impossible since J4 has weight 3. Hence, the initial assumption must be false, and J1, J2, J3, J4 are locally disconnected. a b c d J1 J2 J3 1 1 3 (a) a b c d J1 J2 J3 J4 3 1 2 3 (b) Figure 14: The two situations described in the proof of Theorem 7.9. In part (a), the copies are locally disconnected. In part (b), copy J4 has too many vertices in S. Theorem 7.7, along with the fact that the patterns 113 and 3123 overlap by at most one copy, leads to the following corollary. Corollary 7.10. Consider the graph Jn for even n ≥ 6, and a connected dominating set S for Jn such that |S| = 2n− 1. If γc(Jn−1) = 2n− 3, then S contains at most one instance the patterns 113 or 3123. We are finally ready to prove the main result of this section. Theorem 7.11. Consider the graph Jn for n ≥ 3. Then, γc(Jn) = { 2n, if n is even, 2n− 1, if n is odd. Proof. The upper bounds are provided in Section 4. We use the MTZ formulation for connected domination from [10] to confirm that γc(J3) = 5, γc(J4) = 8, γc(J5) = 9, R. Burdett et al.: Variants of the domination number for flower snarks 443 and γc(J6) = 12. Suppose there is some value k ≥ 7 such that Theorem 7.11 is true for n = 3, . . . , k − 1, but false for n = k. We will first consider the case when k is odd. Then there is a connected dominating set S for Jk such that |S| = 2k − 2. Clearly, S contains a copy, say J i, with weight 1, and ai ̸∈ S. Then from Theorem 7.4, it implies that there is a connected dominating set for Jk−1 with cardinality 2k− 3, which contradicts the assumption that Theorem 7.11 is true for Jk−1. Hence, k must be even. Then, there is a connected dominating set S for Jk such that |S| = 2k − 1. Suppose that there are two copies of Jk with weight 2 in S. Then, from Theorem 7.4 we can obtain a connected dominating set for Jk−2 with cardinality 2k − 5, which contradicts the initial assumption. Hence, there must be at most one copy of Jk with weight 2 in S. That is, either wS2 = 0 or w S 2 = 1. Suppose that wS2 = 0. Then we have w S 1 +w S 3 +w S 4 = k, and w S 1 +3w S 3 +4w S 4 = 2k−1. From the latter equation, it is clear that wS1 and w S 3 must have different parity. Hence, from the former equation, wS4 must be odd. Combining the two equations, we obtain 2wS1 = k + 1 + w S 4 . Note that this implies that more than half of the copies have weight 1 in S. From Theorem 7.7, S cannot contain the pattern 111. Hence, it can be seen that S contains the pattern 11 at least wS4 +1 times. From Corollary 7.10 we know there is at most one instance of the pattern 113 in S. Hence, there is at least wS4 instances of the pattern 11 where both adjacent copies have weight 4, which is impossible. Hence, we must have wS2 = 1. Then, we have w S 1 +w S 3 +w S 4 = k−1, and wS1 +3wS3 + 4wS4 = 2k − 3. From the latter equation, it is clear that wS1 and wS3 must have different parity. Hence, from the former equation, wS4 must be even. Combining the two equations, we obtain 2wS1 = k + w S 4 . Similar to before, this implies that at least half of the copies have weight 1 in S. From Theorem 7.7, S cannot contain the pattern 111. Hence, it can be seen that S contains the pattern 11 at least wS4 times. Now, suppose that S contains the pattern 4114. From Lemma 7.5, we can then obtain a connected dominating set for Jk−3 with cardinality 2k−7 which has one fewer copy of weight 4 than S, which in turn violates Lemma 7.6. Hence, every instance of the pattern 11 has a copy of weight 2 or 3 next to it, and from Corollary 7.10 this means there is at most one instance of the pattern 11. This, in turn, implies that wS4 ≤ 1, and since wS4 is even, we have wS4 = 0. Hence, wS1 = k2 . If there are no instances of the pattern 11, then S must contain the pattern 1312131, violating Theorem 7.7. Hence there is exactly one instance of the pattern 11. This means that there must be one other instance in S of two consecutive copies having non-unit weight. The only options are that S contains the pattern 23, or the pattern 33. Suppose S contains the pattern 23. Without loss of generality, suppose that the copies J3, J4 meet this pattern. Since this is the only instance of S having two consecutive copies of weight other than 1, we know that J2 has weight 1. From Theorem 7.8 we know that J1 cannot have weight 1, or else S would contain the pattern 112. The only remaining option is that J1 has weight 3, and so copies J1, J2, J3, J4 meet the pattern 3123. Then, since J3 has weight 2 and wS2 = 1, there are no other copies with weight 2. Hence, the one instance of the pattern 11 which occurs elsewhere in the graph must be followed by a copy of weight 3. That is, S contains both the patterns 113 and 3123, which from Corollary 7.10 is impossible. Therefore, S must not contain the pattern 23, and instead contains the pattern 33. Finally, without loss of generality, suppose that the copies J1, J2 meet the pattern 33. There is one vertex from J1 which is not contained in S. Suppose we define S2 which is equal to the union of S and this one vertex. Hence, S2 is a connected dominating set for Jk 444 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 with cardinality 2k. Then, from Lemma 7.5, we can obtain a connected dominating set for Jk−1 with cardinality 2k−3, which contains exactly one copy with weight 4, contradicting Lemma 7.6. Hence, in all cases a contradiction is reached, completing the proof. 8 Weak Roman domination and secure domination In this section, we determine (simultaneously) the weak Roman domination numbers and the secure domination numbers for flower snarks. Recall that for the weak Roman domina- tion number, rather than seeking to construct a set S, we instead look to define a function f : V → {0, 1, 2}. In order to use language consistent with the rest of this paper, in this section we define the weight of a copy J i to be equal to f(ai) + f(bi) + f(ci) + f(di). Note that this is somewhat more ambiguous than in previous sections; for instance, if a copy has weight 2, it may have two vertices with weight 1, or a single vertex with weight 2. We will deal with these ambiguities when they arise. Similarly to in previous sections, we will say that f contains a pattern if there is a set of consecutive copies whose weights meet that pattern. Although there is a technical definition for weak Roman domination (e.g. see Defini- tion 1.4), it is useful to provide an intuitive interpretation. Recall that for standard domina- tion, one interpretation is that the set of vertices needs to be protected. If a guard is placed at a vertex, then it protects that vertex, and also all adjacent vertices. A configuration of guards which protects all vertices corresponds to a dominating set. A similar interpreta- tion applies for weak Roman domination. Suppose that at each vertex we can place up to two guards. As before, a vertex is protected if there is at least one guard there, or there is at least one guard among its adjacent vertices. Then we further consider the notion of a vertex being defended. We say a vertex v is defended if there is at least one guard at v, or alternatively, if it possible to relocate one guard from an adjacent vertex w to v, in such a way that all vertices are still protected by the resulting configuration of guards. In the latter case, we will say that w can defend v. Then a weak Roman dominating function is one in which every vertex is defended. Throughout this section, we will often use an argument which goes as follows; suppose there is exactly one guard at v, and that v defends another adjacent vertex w. In order to do so, the guard from v would move to w. However, doing so may mean another vertex u, also adjacent to v, becomes unprotected. In such a case, we will say that v cannot defend w without leaving u unprotected. Another alternative is as follows; suppose again that v defends w. In order to do so, the guard from v would move to w. However, doing so may mean that u will be unprotected unless there is a guard at another vertex, say x. In such a case, we will say in order for v to defend w, there must be a guard at x to avoid leaving u unprotected. Although the following result is simple, we include it here as it will be used many times in this section. Lemma 8.1. Consider the graph Jn, for n ≥ 3, and a weak Roman dominating function f . If a copy J i has weight 1 in f , then none of the vertices from J i can defend any vertices from J i−1 or J i+1. Proof. Since J i has weight 1, there is exactly one vertex with positive weight. If f(ai) = 1, then the result follows immediately since ai is not adjacent to any vertices from J i−1 R. Burdett et al.: Variants of the domination number for flower snarks 445 or J i+1. Then, suppose instead that f(ai) = 0 and one of the other vertices in J i has weight 1. Then that vertex cannot defend any vertex from J i−1 or J i+1 without leaving ai unprotected. Theorem 8.2. Consider the graph Jn, for n ≥ 3, and a weak Roman dominating function f . Then f does not contain the patterns 111, 1121 or 1212121. Proof. Suppose that f contains the pattern 111. Without loss of generality, suppose that copies J1, J2, J3 meet the pattern 111 in f . This situation is displayed in part (a) of Figure 15. Each vertex in J2 must be defended, but since J1 and J3 have weight 1, from Lemma 8.1 none of their vertices can defend any vertices from J2. Hence, each vertex in J2 must be defended solely by vertices in J2. It is clear that this implies that f(a2) = 1. Then, consider b2. In order for a2 to defend b2, there must be a guard at either c1 or c3 to avoid leaving c2 unprotected. Likewise, in order for a2 to defend b2, there must be a guard at either d1 or d3 to avoid leaving d2 unprotected. Since J1 and J3 have weight 1, at most one vertex from each can have a guard. Without loss of generality, suppose that f(c1) = 1 and f(d3) = 1. Then a2 cannot defend c2 without leaving b2 unprotected. Hence, this is impossible, and f does not contain the pattern 111. Next, suppose that f contains the pattern 1121. Without loss of generality, suppose that copies J1, J2, J3, J4 meet the pattern 1121 in f . Since J2 and J4 have weight 1, from Lemma 8.1 none of their vertices can defend any vertices from J3. Hence, each vertex from J3 must be defended solely by vertices in J3. Clearly, this implies that f(a3) ≥ 1. There are two possibilities, either f(a3) = 2, or f(a3) = 1 and there is a guard at another vertex in J3. Suppose first that f(a3) = 2. Then there are no vertices from J1 or J3 that can defend any vertices from J2. Hence, J2 is in an equivalent situation to J2 in the previous paragraph and so this situation is impossible. Instead, suppose that f(a3) = 1 and, without loss of generality, that f(b3) = 1. This situation is displayed in part (b) of Figure 15. Clearly, vertices c2 and d2 can only be defended by vertices from J2, and hence we must have f(a2) = 1. Then, in order for a2 to defend c2, there must be a guard at d1 to avoid leaving d2 unprotected. Similarly, in order for a2 to defend d2, there must be a guard at c1 to avoid leaving c2 unprotected. This is impossible since J1 has weight 1, and so f does not contain the pattern 1121. a b c d J1 J2 J3 1 1 1 (a) a b c d J1 J2 J3 J4 1 1 2 1 (b) Figure 15: The first two situations described in the proof of Theorem 8.2. In part (a), a2 cannot defend c2 without leaving b2 unprotected. In part (b), there are too many guards in J1. Finally, suppose that f contains the pattern 1212121. Without loss of generality, sup- pose that copies J1, . . . , J7 meet the pattern 1212121 in f . Since J1 and J3 have weight 1, from Lemma 8.1 none of their vertices can defend any vertices from J2. Hence, each vertex from J2 must be defended solely by vertices in J2, which implies that f(a2) ≥ 1. 446 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 Analogous arguments imply that f(a4) ≥ 1 and f(a6) ≥ 1. Now, consider J2. There are two possibilities, either f(a2) = 2, or f(a2) = 1 and there is a guard at another vertex in J2. Suppose first that f(a2) = 2. Then there are no vertices from J2 that can defend any vertices from J3, and at most one vertex from J4 can defend a vertex from J3. Hence, J3 is in an equivalent situation to J2 in the previous paragraph, and so this situation is impossible. Instead, suppose that f(a2) = 1 and, without loss of generality, that f(b2) = 1. This situation is displayed in Figure 16. Then, in order for a2 to defend c2, there must be a guard at either d1 or d3 to avoid leaving d2 unprotected. Similarly, in order for a2 to defend d2, there must be a guard at either c1 or c3 to avoid leaving c2 unprotected. Since J1 and J3 have weight 1, at most one vertex from each may have a guard. Without loss of generality, suppose that f(c1) = 1 and f(d3) = 1. Then, since f is dominating, we must have f(c4) = 1 in order to protect c3. Now, in order for a4 to defend d4, there must be a guard at b5 to avoid leaving b4 unprotected. Then, since f is dominating, we must have f(d6) = 1 in order to protect d5. Finally, consider vertex c5. There is only one adjacent vertex with a guard that can defend it, c4. However, c4 cannot defend c5 without leaving c3 unprotected. Thus, c5 is not defended. This is impossible, and hence f does not contain the pattern 1212121. a b c d J1 J2 J3 J4 J5 J6 J7 1 2 1 2 1 2 1 Figure 16: The third situation described in the proof of Theorem 8.2. Here, c4 cannot defend c5 without leaving c3 unprotected. Lemma 8.3. Consider the graph Jn, for n ≥ 4, and a weak Roman dominating function f with weight w(f) ≤ 3n2 . Then any set of four consecutive copies have a combined weight of 6 in f , every copy has weight either 1 or 2 in f , and w(f) = 3n2 . Proof. Denote by w4(i) the combined weights of J i, J i+1, J i+2, J i+3 where the super- scripts are taken modulo n. Now, by Corollary 2.4 we know that no copies have weight 0, and so w4(i) ≥ 4. Suppose w4(i) = 4. This implies that J i, J i+1, J i+2, J i+3 all have weight 1. However, this means f contains the pattern 111, which by Theorem 8.2 is impossible. Then, suppose w4(i) = 5. Up to symmetry, there are only two possibilities; copies J i, J i+1, J i+2, J i+3 either have the pattern 1112 or 1121. Both of these options are impossible by Theorem 8.2. Hence, we have w4(i) ≥ 6. Then, by assumption, we have∑n i=1 w 4(i) ≤ 6n. This is only possible if all w4(i) = 6, and also implies that w(f) = 3n2 . Since each copy has weight at least one, and each set of four consecutive copies has a combined weight of six, it is clear that any individual copy must have weight at most three. Suppose there is a copy J i with weight 3 in f , then the next three copies must each have weight 1. However, by Theorem 8.2 this is impossible. Hence, each copy J i must have weight either 1 or 2 in f . R. Burdett et al.: Variants of the domination number for flower snarks 447 Lemma 8.4. Consider the graph Jn, for n ≥ 3, and a weak Roman dominating func- tion f with weight w(f). Suppose there is an integer k ≥ 3 such that in f , the config- uration of guards in copies J i+1, . . . , J i+k is identical to the configuration of guards in J i+k+1, . . . , J i+2k, and suppose that J i+1, . . . , J i+k collectively contain g guards. Then there is a weak Roman dominating function for Jn−k with weight equal to w(f)− g. Proof. Suppose that we “smooth out” copies J i+1, . . . , J i+k in the way displayed in Fig- ure 11. That is, we add edges connecting bi to bi+k+1, ci to ci+k+1, and di to di+k+1, and then we delete the copies J i+1, . . . , J i+k. Call the resulting graph G2. It is easy to check that G2 is isomorphic to Jn−k. Then, define a new function f2 such that f2(v) = f(v) if v ∈ V \ {V (J i+1) ∪ . . . ∪ V (J i+k)}, and f2(v) is undefined otherwise. Since J i+1, . . . , J i+k collectively contain g guards in f , it is clear that w(f2) = w(f)− g. Then if we can show that f2 is a weak Roman dominating function for G2, the result follows immediately. For a given function, one can check to see if a vertex v is defended by observing the configuration of guards in vertices no more than distance three away from v. Then, for any vertex v in G2, consider the configuration of graphs (according to f2) in the subgraph induced by the vertices at most distance 3 from v. By assumption, this is identical to the corresponding configuration of guards (according to f ) in the subgraph induced by the vertices at most distance 3 from v in Jn. Hence, all vertices of G are defended in f2, and so f2 is a weak Roman dominating function for G, leading to the desired result. In the proof of the following theorem, whenever there are two guards at a vertex, we will mark that vertex with a . Theorem 8.5. Consider the graph Jn, for n ≥ 9, and a weak Roman dominating function f . If f contains the pattern 21122112, then there exists a weak Roman dominating function for Jn−4 with weight equal to w(f)− 6. Proof. Without loss of generality, suppose that copies J1, . . . , J8 meet the pattern 21122112 in f . Consider copy J4, and suppose that f(a4) = 2. This situation is dis- played in part (a) of Figure 17. Then, no vertices from J4 can defend any vertices from J3, and by Lemma 8.1 the same is also true for J2. Hence, all four vertices in J3 must be defended by vertices in J3. This implies that f(a3) = 1. Then, in order for a3 to defend b3, there must be a guard at c2 to avoid leaving c3 unprotected, and there must also be a guard at d2 to avoid leaving d3 unprotected. Since J2 has weight 1, this is impossible, and so f(a4) ̸= 2. Then, suppose that f(a4) = 1. Then there is another vertex in J4 with a guard. With- out loss of generality, suppose that f(b4) = 1. This situation is displayed in part (b) of Figure 17. Using a similar argument as in the previous paragraph, vertices c3 and d3 must be defended by vertices from J3, and hence f(a3) = 1. Then, in order for a3 to defend c3, there must be a guard at d2 to avoid leaving d3 unprotected. Likewise, in order for a3 to defend d3, there must be a guard at c2 to avoid leaving c3 unprotected. Since J2 has weight 1, this is impossible. Hence we can conclude that f(a4) = 0. It is clear that identical arguments can be made to conclude that f(a1) = f(a5) = f(a8) = 0 as well. Then, suppose that there is a vertex in J4 with two guards. Without loss of generality, suppose that f(b4) = 2. This situation is displayed in part (c) of Figure 17. An identical argument to that in the previous paragraph can be used to show that f(a3) = 1, and this again implies that f(c2) = f(d2) = 1. Since J2 has weight 1, this is impossible. Hence, 448 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 exactly two vertices in J4 have weight 1. Again, identical arguments can be made to reach the same conclusion for each of J1, J5 and J8. Now, without loss of generality, suppose that f(b4) = f(c4) = 1. Then, d4 must be defended by either d3 or d5, and by Lemma 8.1 it cannot be d3. Hence, we must have f(d5) = 1, and there must be one more vertex in J5 with a guard, either b5 or c5. Due to symmetry, either choice may be made without loss of generality; we will choose f(b5) = 1. Now, consider vertex d3. From Lemma 8.1 it cannot be defended by d2. Hence, we must have either f(a3) = 1 or f(d3) = 1. Suppose that f(a3) = 1. Then, c3 must be defended by at least one of a3 or c4 (from Lemma 8.1 it cannot be defended by c2). Suppose first that a3 defends c3. This situation is displayed in part (d) of Figure 17. In order for a3 to defend c3, there must be a guard at d2 to avoid leaving d3 unprotected. Then, in order to defend b2 and c2 we must have f(b1) = 1 and f(c1) = 1 respectively. But then, b1 cannot defend a1 without leaving b2 unprotected, and c1 cannot defend a1 without leaving c2 unprotected. Hence, a1 is not defended, which is impossible, and so we conclude that a3 cannot defend c3. Hence, c4 must defend c3. This situation is displayed in part (e) of Figure 17. In order for c4 to defend c3, there must be a guard at c6 to avoid leaving c5 unprotected. Also, the only vertex which can defend d4 is d5, but in order to do so, there must be a guard at d7 to avoid leaving d6 unprotected. Finally, consider a5, which can only be defended by either b5 or d5. However, b5 cannot defend a5 without leaving b6 unprotected, and d5 cannot defend a5 without leaving d4 unprotected. Hence, all of these cases are impossible, and so conclude that f(a3) = 0, and accordingly, f(d3) = 1. The current situation is displayed in part (f) of Figure 17. Now consider c3. From Lemma 8.1 it is clear that c2 cannot defend c3, so c4 must defend c3. In order for c4 to defend c3, there must be a guard at c6 to avoid leaving c5 unprotected. Likewise, from Lemma 8.1, d3 cannot defend d4, so d5 must defend d4. In order for d5 to defend d4, there must be a guard at d7 to avoid leaving d6 unprotected. Furthermore, from Lemma 8.1, c6 cannot defend c5, so c4 must defend c5. In order for c4 to defend c5, there must be a guard at c2 to avoid leaving c3 unprotected. Finally, from Lemma 8.1, the only vertices which can defend b2 and d2 are b1 and d1 respectively, and so f(b1) = f(d1) = 1. Likewise, from Lemma 8.1, the only vertices which can defend b7 and c7 are b8 and c8 respectively, so f(b8) = f(c8) = 1. At this point, it can be seen that the configuration of guards in J1, J2, J3, J4 is identical to the configuration of guards in J5, J6, J7, J8. Then, the result follows immediately from Lemma 8.4. We are finally ready to prove the main result of this section. Theorem 8.6. Consider the graph Jn, for n ≥ 3. Then, γs(Jn) = γr(Jn) = ⌈ 3n+ 1 2 ⌉ . Proof. The (coincident) upper bounds for both weak Roman domination and secure dom- ination are provided in Section 4. Note that γr(Jn) ≤ γs(Jn), so a corresponding lower bound for weak Roman domination will also serve as a lower bound for secure domination. We use the formulation for weak Roman domination from [7] to confirm that γr(J3) = 5, γr(J4) = 7, γr(J5) = 8, γr(J6) = 10, γr(J7) = 11, and γr(J8) = 13, and the formu- lation from [5] to confirm the corresponding results for secure domination. Then, suppose there is a value k ≥ 9 such that Theorem 8.6 is true for n = 3, . . . , k − 1, but is not true for n = k. That is, there exists a weak Roman dominating function f for Jk with R. Burdett et al.: Variants of the domination number for flower snarks 449 a b c d J2 J3 J4 1 1 2 (a) a b c d J1 J2 J3 J4 J5 2 1 1 2 2 (d) a b c d J2 J3 J4 1 1 2 (b) a b c d J3 J4 J5 J6 J7 1 2 2 1 1 (e) a b c d J2 J3 J4 1 1 2 (c) a b c d J1 J2 J3 J4 J5 J6 J7 J8 2 1 1 2 2 1 1 2 (f) Figure 17: The six situations described in the proof of Theorem 8.5. In parts (a), (b) and (c), there are too many guards in copy J2. In part (d), a1 is undefended. In part (e), a5 is undefended. In part (f), the configuration of guards in J1, J2, J3, J4 is identical to that in J5, J6, J7, J8. weight w(f) < ⌈ 3k+1 2 ⌉ . This implies that w(f) ≤ 3k2 . Hence, from Lemma 8.3 we have w(f) = 3k2 , each set of four consecutive copies has weight 6, and each copy has weight 1 or 2. It is easy to check that there are only two possibilities. Either f has the repeating pattern 121212 . . . 12, or f has the repeating pattern 21122112 . . . 2112. Note that it is impossible to combine these two patterns without there being a set of four consecutive copies with weight not equal to 6. Then, since k ≥ 9, by Theorem 8.2 the pattern 121212 . . . 12 is impossible. Hence, f must have the repeating pattern 21122112 . . . 2112. Then, by Theorem 8.5 there is weak Roman dominating function for Jk−4 with weight equal to 3k 2 − 6 < ⌈ 3(k−4)+1 2 ⌉ . This contradicts our initial assumption, completing the proof. ORCID iDs Michael Haythorpe https://orcid.org/0000-0001-8143-6583 Alex Newcombe https://orcid.org/0000-0002-4401-0951 References [1] I. Agustin and D. Dafik, On the domination number of some families of special graphs, Prosid- ing Seminar Matematika dan Pendidikan Matematik 1 (2014), https://jurnal.unej. ac.id/index.php/psmp/article/view/921. 450 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 [2] S. Alanko, S. Crevals, A. Isopoussu, P. Östergård and V. Pettersson, Computing the domination number of grid graphs, Electron. J. Comb. 18 (2011), research paper p141, 18 pp., doi:10. 37236/628, https://doi.org/10.37236/628. [3] S. Arumugam and R. Kala, Domination parameters of hypercubes, J. Indian Math. Soc., New Ser. 65 (1998), 31–38. [4] A. Burcroff, Tightness of domination inequalities for direct product graphs, Discrete Math. 344 (2021), 10, doi:10.1016/j.disc.2021.112495, id/No 112495, https://doi.org/10. 1016/j.disc.2021.112495. [5] R. Burdett and M. Haythorpe, An improved binary programming formulation for the secure domination problem, Ann. Oper. Res. 295 (2020), 561–573, doi:10.1007/s10479-020-03810-6, https://doi.org/10.1007/s10479-020-03810-6. [6] R. Burdett, M. Haythorpe and A. Newcombe, Binary programming formulations for the upper domination problem, Ann. Oper. Res. (2021), submitted. [7] A. Burger, A. De Villiers and J. Van Vuuren, A binary programming approach towards achiev- ing effective graph protection, in: Proc. 2013 ORSSA Annual Conf., ORRSA, pp. 19–30, 2013. [8] T. Y. Chang and W. E. Clark, The domination numbers of the 5×n and 6×n grid graphs, J. Graph Theory 17 (1993), 81–107, doi:10.1002/jgt.3190170110, https://doi.org/10. 1002/jgt.3190170110. [9] S. Crevals and P. R. J. Östergard, On the domination number of 2-dimensional torus graphs, Util. Math. 106 (2018), 289–300. [10] N. Fan and J.-P. Watson, Solving the connected dominating set problem and power dominating set problem by integer programming, in: Combinatorial optimization and applications. 6th international conference, COCOA 2012, Springer, Berlin, pp. 371–383, 2012, doi:10.1007/ 978-3-642-31770-5 33, https://doi.org/10.1007/978-3-642-31770-5_33. [11] G. H. Fricke, S. M. Hedetniemi, S. T. Hedetniemi, A. A. McRae, C. K. Wallis, M. S. Jacobson, H. W. Martin and W. D. Weakley, Combinatorial problems on chessboards: A brief survey, in: Graph Theory, Combinatorics, Algorithms and Applications. Vol. 1. Proceedings of the seventh quadrennial international conference on the theory and applications of graphs, Kalamazoo, MI, USA, June 1-5, 1992, Wiley, New York, NY, pp. 507–528, 1995. [12] X. Fu, X. Xu, Y. Yang and X. Feng, On the domination number of Knödel graph W3,n, Int. J. Pure Appl. Math. 50 (2009), 553–558. [13] X. Fu, Y. Yang and B. Jiang, On the domination number of generalized petersen graphs P (n, 2), Discrete Math. 309 (2009), 2445–2451, doi:10.1016/j.disc.2008.05.026, https: //doi.org/10.1016/j.disc.2008.05.026. [14] D. Gonçalves, A. Pinlou, M. Rao and S. Thomassé, The domination number of grids, SIAM J. Discrete Math. 25 (2011), 1443–1453, doi:10.1137/11082574, https://doi.org/10. 1137/11082574. [15] S. Gravier, Total domination number of grid graphs, Discrete Appl. Math. 121 (2002), 119–128, doi:10.1016/S0166-218X(01)00297-9, https://doi.org/10.1016/ S0166-218X(01)00297-9. [16] E. Hare, Algorithms for Grids and Grid-Like Graphs, Ph.D. thesis, Clemson University, 1989. [17] M. Haythorpe and A. Newcombe, The secure domination number of Cartesian products of small graphs with paths and cycles, Discrete Appl. Math. 309 (2022), 32–45, doi:10.1016/j. dam.2021.11.008, https://doi.org/10.1016/j.dam.2021.11.008. [18] J. T. Hedetniemi and S. T. Hedetniemi, Domination in chessboards, in: Structures of Domi- nation in Graphs, Springer, Cham, pp. 341–386, 2021, doi:10.1007/978-3-030-58892-2 12, https://doi.org/10.1007/978-3-030-58892-2_12. R. Burdett et al.: Variants of the domination number for flower snarks 451 [19] R. Isaacs, Infinite families of nontrivial trivalent graphs which are not Tait colorable, Am. Math. Mon. 82 (1975), 221–239, doi:10.2307/2319844, https://doi.org/10.2307/ 2319844. [20] M. S. Jacobson and L. F. Kinch, On the domination number of products of graphs. I, Ars Comb. 18 (1984), 33–44. [21] A. Klobučar, Total domination numbers of Cartesian products, Math. Commun. 9 (2004), 35– 44. [22] J. Kratica, D. Matić and V. Filipović, Weakly convex and convex domination numbers for generalized Petersen and flower snark graphs, Rev. Unión Mat. Argent. 61 (2020), 441–455, doi:10.33044/revuma.v61n2a16, https://doi.org/10.33044/revuma.v61n2a16. [23] Y. Lai, C. Lin and H. Ho, Weak roman domination on graphs, in: Proceedings of the 28th workshop on combinatorial mathematics and computation theory, Penghu University of Science and Technology, Penghu, Taiwan, May 27–28, 2011, 2011. [24] J. Liu and X. Zhang, The exact domination number of generalized Petersen graphs P (n, k) with n = 2k and n = 2k + 2∗, Comput. Appl. Math. 33 (2014), 497–506, doi:10.1007/ s40314-013-0077-8, https://doi.org/10.1007/s40314-013-0077-8. [25] Z. Maksimovic, J. Kratica, A. Savić and M. Bogdanović, Some static roman domination num- bers for flower snarks, in: XIII Balkan Conference on Operational Research Proceedings, 2018. [26] D. Mojdeh, M. Habibi and L. Badakhshian, Total and connected domination in chemical graphs, Ital. J. Pure Appl. Math. 39 (2018), 393–401. [27] D. A. Mojdeh, R. S. Musawi, E. Nazari Kiashi and N. Jafari Rad, Total domination in cubic Knödel graphs, Commun. Comb. Optim. 6 (2021), 221–230, doi:10.22049/cco.2020.26793. 1143, https://doi.org/10.22049/cco.2020.26793.1143. [28] P. Pavlič and J. Žerovnik, A note on the domination number of the Cartesian products of paths and cycles, Kragujevac J. Math. 37 (2013), 275–285. [29] S. Pouyandeh, M. Golriz, M. Sorouhesh and M. Khademi, On domination number of Latin square graphs of finite cyclic groups, Discrete Math. Algorithms Appl. 13 (2021), 8, doi:10.1142/s1793830920500901, id/No 2050090, https://doi.org/10.1142/ s1793830920500901. [30] P. R. L. Pushpam and M. Kamalam, Weak Roman domination in chess graphs, Malaya J. Mat. 9 (2021), 486–493, doi:10.26637/MJM0901/0082, https://doi.org/10.26637/ MJM0901/0082. [31] P. R. L. Pushpam and C. Suseendran, Secure domination in middle graphs, Bull. Int. Math. Virtual Inst. 9 (2019), 25–38. [32] J. Quadras, A. Sajiya Merlin Mahizl, I. Rajasingh and S. R. Rajan, Domination in cer- tain chemical graphs, J. Math. Chem. 53 (2015), 207–219, doi:10.1007/s10910-014-0422-1, https://doi.org/10.1007/s10910-014-0422-1. [33] E. Sampathkumar and H. B. Walikar, The connected domination number of a graph, J. Math. Phys. Sci. 13 (1979), 607–613. [34] L. Shan, H. Li and Z. Zhang, Domination number and minimum dominating sets in pseud- ofractal scale-free web and Sierpiński graph, Theor. Comput. Sci. 677 (2017), 12–30, doi: 10.1016/j.tcs.2017.03.009, https://doi.org/10.1016/j.tcs.2017.03.009. [35] A. Shobana and P. Senniappan, Connected domination number in circulant graphs, J. Math. Anal. 9 (2015), 2485–2488, doi:10.12988/ijma.2015.59228, https://doi.org/ 10.12988/ijma.2015.59228. 452 Ars Math. Contemp. 24 (2024) #P3.04 / 427–452 [36] A. Spalding, Min-Plus Algebra and Graph Domination, Ph.D. thesis, University of Colorado, 1998. [37] M. Valveny and J. A. Rodrı́guez-Velázquez, Protection of graphs with emphasis on Cartesian product graphs, Filomat 33 (2019), 319–333, doi:10.2298/FIL1901319V, https://doi. org/10.2298/FIL1901319V. [38] A. Vijayan and M. F. N. Mabel, Connected dominating sets and connected domination polynomials of square of centipedes, Journal of Information and Optimization Sciences 37 (2016), 155–165, doi:10.1080/02522667.2015.1103031, https://doi.org/10.1080/ 02522667.2015.1103031. [39] C. Wallis, Domination parameters of line graphs of designs and variations of chessboard graphs, Ph.D. thesis, Clemson University, 1996. [40] A. M. Yaglom and I. M. Yaglom, Challenging mathematical problems with elementary solu- tions. Vol. I, 1967. [41] H. Yan, L. Kang and G. Xu, The exact domination number of the generalized Petersen graphs, Discrete Math. 309 (2009), 2596–2607, doi:10.1016/j.disc.2008.04.026, https: //doi.org/10.1016/j.disc.2008.04.026. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.05 / 453–471 https://doi.org/10.26493/1855-3974.2771.4df (Also available at http://amc-journal.eu) Optimal strategies in fractional games: vertex cover and domination Csilla Bujtás * Faculty of Mathematics and Physics, University of Ljubljana, Ljubljana, Slovenia and University of Pannonia, Veszprém, Hungary and Institute of Mathematics, Physics and Mechanics, Jadranska 19, Ljubljana, Slovenia Günter Rote Institut für Informatik, Freie Universität Berlin, Takustr. 9, 14195 Berlin, Germany Zsolt Tuza † HUN-REN Alfréd Rényi Institute of Mathematics, Reáltanoda u. 13–15, 1053 Budapest, Hungary and Department of Computer Science and Systems Technology, University of Pannonia, Egyetem u. 10, 8200 Veszprém, Hungary Received 14 December 2021, accepted 27 February 2024, published online 18 July 2024 Abstract In a hypergraph H = (V, E) with vertex set V and edge set E , a real-valued function f : V → [0, 1] is a fractional transversal if ∑ v∈E f(v) ≥ 1 holds for every E ∈ E . Its size is |f | := ∑ v∈V f(v), and the fractional transversal number τ ∗(H) is the smallest possible |f |. We consider a game scenario where two players have opposite goals, one of them trying to minimize and the other to maximize the size of a fractional transversal constructed in- crementally. We prove that both players have strategies to achieve their common optimum, and they can reach their goals using rational weights. Keywords: Fractional vertex cover, fractional transversal game, fractional domination game. Math. Subj. Class. (2020): 05C69, 05C65, 05C57 *Corresponding author. The author acknowledges the financial support from the Slovenian Research and Innovation Agency under the projects N1-0108 and P1-0297. †The author’s research was supported in part by the National Research, Development and Innovation Office – NKFIH under the grant SNN 129364. E-mail addresses: bujtasc@gmail.com (Csilla Bujtás), rote@inf.fu-berlin.de (Günter Rote), tuza@dcs.uni-pannon.hu (Zsolt Tuza) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 454 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 1 Introduction Let H = (V, E) be a finite hypergraph, where V is the finite vertex set and E is the edge set, a set system over the underlying set V . We assume that every edge contains at least one vertex; that is, E ⊆ 2V \ {∅}. A hypergraph is k-uniform if |E| = k holds for all E ∈ E . A set T ⊆ V is a transversal 1 of H if every edge is covered by a vertex of T , which formally means that T ∩ E ̸= ∅ holds for all E ∈ E . Its real relaxation, called fractional transversal, is a function f : V → [0, 1] such that ∑ v∈E f(v) ≥ 1 holds for every E ∈ E . The size of f is defined as |f | := ∑ v∈V f(v). The transversal number τ(H) and the fractional transversal number τ∗(H) of H are the minimum cardinality |T | of a transversal and minimum value |f | of a fractional transversal, respectively. The transversal game is a competitive optimization version of hypergraph transversals, which was introduced in [9] and studied further in [10]. It is played on a hypergraph H by two players called Edge-hitter and Staller. They take turns choosing a vertex. The game is over when all edges are covered, and the length of the game is the number of vertices chosen by the players until the end of the game. Edge-hitter wants to finish the game as soon as possible, while Staller wants to delay the end. To prevent Staller from making completely useless moves, we stipulate that the chosen vertex must be contained in at least one previously uncovered edge. Assuming that both players play optimally2 and Edge-hitter starts, the length of the game on H is uniquely determined. It is called the game transversal number of H and is denoted by τg(H). Analogously, the Staller-start game transversal number of H, denoted by τ ′g(H), is the length of the game under the same rules when Staller makes the first move. Among other results, it was proved in [9] that |τg(H) − τ ′g(H)| ≤ 1 always holds. We further recall that, denoting by n(H) and m(H) the number of vertices and edges in H respectively, 411 (n(H) +m(H)) is a (sharp) upper bound on τg(H) if H does not contain one-element edges and it is not isomorphic to the cycle C4. Below we shall refer to this game as the integer game, as opposed to its fractional version which we will introduce in the next section. The important motivation of this approach are the domination game [7] and the to- tal domination game [17], where in fact the transversal game is played on the ‘closed neighborhood hypergraph’ and on the ‘open neighborhood hypergraph’ of a graph, re- spectively.3 Further variants studied so far include the disjoint domination [14], con- nected domination [2], and fractional domination [15] games on graphs, and the dom- ination games on hypergraphs [13]. Some of the most recent results can be found in [3, 4, 8, 11, 12, 16, 18, 19, 20, 21, 22, 23]. For a thorough survey and list of further 1In various areas of discrete mathematics and computer science, a transversal is called vertex cover, or hitting set, or blocking set. It is also equivalent to the set cover in the dual hypergraph. 2A strategy of a player means that every possible state of the game is associated with a move he/she will play if that situation arises. From Edge-hitter’s point of view, the value vE(S) of a strategy S is the smallest integer k such that, if Edge-hitter plays according to S, the transversal game always finishes in at most k moves (no matter which strategy is applied by Staller). We say that S is an optimal strategy for Edge-hitter, if vE(S) is the possible smallest value over the family of all strategies. Similarly, from Staller’s point of view, a strategy S can be associated with the value vS(S) that is the largest integer k such that, if Staller follows strategy S, the length of the game is always at least k; further S is an optimal strategy for Staller, if vS(S) is the largest value over the family of all strategies. The reader may find more about optimal strategies and the uniqueness of the corresponding parameters in [5, Section 1.2]. 3Recall from the literature that the closed and open neighborhood hypergraphs of a graph G are defined on the same vertex set V as G, and the closed (resp. open) neighborhood hypergraph consists of edges corresponding to the closed (resp. open) neighborhoods of vertices in G. Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 455 references see the book [5]. Our results In Section 2 we introduce the rules of the game and prove that its value is well-defined. We present some examples, showing that it makes a difference whether Edge-hitter or Staller starts. Moreover, edges that are supersets of other edges of the hypergraph may influence the game value, in contrast to the standard non-game version of the transversal number. In Section 3 we compare the game transversal number with other related parameters, and prove a monotonicity property, implying that changing the starting player can affect the value of the game by at most 1. The rules of the game allow the players to split their moves into infinitely many sub- moves. In Section 4 and 5 we give some structural results showing that the full generality of the moves allowed by our rules is not needed. Namely, any infinite move is equivalent to a finite move, and Edge-hitter can restrict his strategy to moves in which every permutation of submoves is equally good. In Section 6 we prove that the game can be modeled in a way that leads to an opti- mization problem solvable via the theory of piecewise linear continuous rational functions. From this, we derive that the game value is rational for every finite hypergraph; moreover both players can achieve their goals using rational submoves. Consequences concerning domination games and several conjectures are given in the concluding Section 7. 2 Fractional transversal game Let H = (V, E) be a hypergraph. In the context of the fractional transversal game, we will consider a cover function t : V → [0, 1] that is updated after each move during the game. We denote by |t| the sum ∑ v∈V t(v). Given a cover function t, the corresponding load function is ℓ : E → [0, 1] defined by the rule ℓ(E) = ℓ(E, t) = min { 1, ∑ v∈E t(v) } for every E ∈ E . If ℓ ≡ 1, we say that H is fully covered. We shall write ti and ℓi for the cover and load functions after the ith move. The game begins with t0 ≡ 0 and therefore with ℓ0 ≡ 0. It is finished when the hypergraph becomes fully covered. Edge-hitter and Staller take turns making moves under the following rules. As long as ℓ ̸≡ 1, the next player performs a move, which is a sequence (vi1 , w1), (vi2 , w2), . . . of arbitrary length (possibly infinite). It consists of the submoves (vik , wk), k = 1, 2, . . . , where vi1 , vi2 , . . . are vertices of H with any number of repetitions allowed, and the weights w1, w2, . . . are real numbers from [0, 1]. We say that a submove (vik , wk) is legal if it increases the load of some edge by wk. In a legal move, a player makes a series of legal submoves such that the sum of the weights equals 1 or the move completes the game, whichever comes first. Formally, the ith move (vi1 , w1), (vi2 , w2), . . . is legal, if the following conditions hold: (∗) For every k ≥ 1 there exists an edge E ∈ E such that vik ∈ E and ℓi−1(E) + ( ∑ vis ∈E 1≤s≤k−1 ws ) + wk ≤ 1 . 456 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 (∗∗) The total weight constraint: ∑ k≥1 wk ≤ 1, and if the move does not end the game, then ∑ k≥1 wk = 1. The cover function ti can gradually be reached from ti−1 by adding the weight wk to t(vik) after each submove; this process converts also the corresponding load function from ℓi−1 to ℓi. Suppose that a fractional transversal game G finishes with the qth move. The value |G| of the game is defined as the value |tq| of the cover function obtained at the end, that is the sum of the weights that have been spent during the game. The goal of Edge-hitter is to achieve a value |G| as small as possible, while Staller wants a large |G|. Assuming that Edge-hitter starts the fractional transversal game on H, we consider the set of upper bounds, UH = {a ∈ R : Edge-hitter has a strategy that ensures |G| ≤ a} and the set of lower bounds, LH = {b ∈ R : Staller has a strategy that ensures |G| ≥ b}. Formally the game fractional transversal number τ∗g (H) is defined as τ∗g (H) = inf(UH). The Staller-start game fractional transversal number τ∗g ′(H) is defined similarly, under the condition that the first move is made by Staller. The following assertion shows that τ∗g (H) is also equal to sup(LH), and the situation is similar if Staller starts the game. The proof is essentially the same as the one for the game fractional domination number in [15]. Proposition 2.1. For every hypergraph H we have inf(UH) = sup(LH), and the analo- gous equality holds for the Staller-start game, too. Proof. First, assume that inf(UH) < sup(LH) and consequently, there exist two reals x and y satisfying inf(UH) < x < y < sup(LH). By definition, x ∈ UH and, therefore, Edge-hitter can ensure that, under every strategy of Staller, the value of the game is at most x. Similarly, y ∈ LH and Staller has a strategy that ensures |G| ≥ y whatever strategy is followed by Edge-hitter. This is a contradiction that establishes inf(UH) ≥ sup(LH). Now, we prove the reverse inequality. By definition, z < inf(UH) implies that Edge- hitter does not have a strategy to achieve |G| ≤ z. That is, against each strategy of Edge- hitter there is a strategy of Staller which results in |G| > z. We may infer that z ∈ LH and therefore z ≤ sup(LH). Since it holds for every z < inf(UH), we conclude inf(UH) ≤ sup(LH). This completes the proof of the proposition. Later, in Section 6, we will show that inf(UH) = min(UH) and sup(LH) = max(LH). Therefore, Edge-hitter and Staller have optimal strategies under which, respectively, |G| ≤ τ∗g (H) and |G| ≥ τ∗g (H) are achieved. Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 457 2.1 Examples for the fractional transversal game (1) Our first example is the 4-cycle C4 = v1v2v3v4v1, which can also be considered as a 2-uniform hypergraph. It is easy to check that τ∗(C4) = 2, while τg(C4) = 3 and τ ′g(C4) = 2 were proved for the integer games [9]. Now we prove that τ ∗ g (C4) = 5/2. • For the upper bound, the following strategy of Edge-hitter ensures that the sum of the weights spent during the game is at most 5/2. His first move is (v1, 14 ), (v2, 1 4 ), (v3, 1 4 ), (v4, 1 4 ); it results in ℓ1 ≡ 1 2 and ∑ E∈E ℓ1(E) = 2. Then, for the first 1 2 of the weight spent by Staller, each of her submoves necessarily increases the load of both incident edges; and each of her remaining submoves increases the load of at least one edge. Therefore, the total load increases by at least 12 × 2 + 1 2 = 3 2 to∑ E∈E ℓ2(E) ≥ 7 2 , and Edge-hitter can achieve ∑ E∈E ℓ3(E) = 4 by spending at most 12 in the final move. This proves τ ∗ g (C4) ≤ 5/2. • Let us show the reverse inequality. We note that the first move of the game has the same effect as a sequence (v1, w1), (v2, w2), (v3, w3), (v4, w4) of submoves with∑4 i=1 wi = 1. 4 After this move of Edge-hitter, ∑ E∈E ℓ1(E) = 2 and hence, there is an edge E with ℓ1(E) ≤ 12 . By symmetry, we may assume that ℓ1(v1v2) ≤ 1 2 . Let Staller play the move (v3, w1+w4), (v4, w2+w3). The move is legal as ℓ1(v2v3) = w2 + w3 = 1 − (w1 + w4) and ℓ1(v4v1) = w1 + w4 = 1 − (w2 + w3). We then have ℓ2(v1v2) = ℓ1(v1v2) ≤ 12 and Edge-hitter needs to spend at least 1 2 to finish the game. This strategy of Staller shows τ∗g (C4) ≥ 5/2. If Staller starts the fractional transversal game on C4 with the move (v1, w1), (v2, w2), (v3, w3), (v4, w4), then Edge-hitter can ensure |G| = 2 by playing (v1, w3), (v2, w4), (v3, w1), (v4, w2). Indeed, ℓ2 assigns w1 +w2 +w3 +w4 = 1 to every edge of the graph. Therefore, τ∗g ′(C4) ≤ 2. Since τ∗g ′(C4) ≥ τ∗(C4) = 2 also holds5, we get τ∗g ′(C4) = 2. v4 u4 v1 u1 v2 u2 v3 u3 Figure 1: A hypergraph H with nested edges. 4Theorem 4.1 and Lemma 5.1 will give conditions for this replacement property in general. For the first move of the game, it is much easier to see as, for each edge E, its load ℓ1(E) equals the sum of the weights assigned to the vertices of E. 5See the proof of Proposition 3.1(i) for a simple explanation. 458 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 (2) Now we modify the previous example C4 by adding four new vertices u1, . . . , u4 and four new edges {v1, v2, u1}, . . . , {v4, v1, u4} to get the hypergraph H shown in Figure 1. When the fractional (or integer) transversal number is considered, each edge that is a super- set of another edge can be deleted, which implies τ∗(H) = τ∗(C4) = 2. We show that the situation is different for the fractional transversal game on H, that is τ∗g (H) = 3 ̸= τ∗g (C4). Suppose that Edge-hitter starts the fractional transversal game G on H. • We first show that Edge-hitter can ensure that the value |G| of the game is at most 3. An optimal first move for him is (v1, 14 ), (v2, 1 4 ), (v3, 1 4 ), (v4, 1 4 ). Then, inde- pendently of Staller’s reply, Edge-hitter plays (v1, w1), (v2, w2), (v3, w3), (v4, w4) as his second move, where wi equals 14 or, if (vi, 1 4 ) is not a legal submove, wi is the maximum legal weight for vi. After this move, ℓ3 ≡ 1 and we may infer that |G| ≤ 3. This proves τ∗g (H) ≤ 3. • Our second claim is that Staller has a strategy that always results in |G| ≥ 3. As each vertex of H belongs to at most two 3-element edges, after Edge-hitter’s first move the sum of the loads of the 3-element edges is at most 2. Thus, Staller can play a legal move that does not assign weights to v1, v2, v3, v4. After this move, the sum of the loads of the 2-element edges remains at most 2, and Edge-hitter has to spend a weight of at least 1 to finish the game. This shows τ∗g (H) ≥ 3. We therefore conclude τ∗g (H) = 3 > τ∗g (C4). (3) The removal of the edges which are subsets of other edges in a hypergraph F may also change the values of the parameters. For instance, let F be the hypergraph obtained from C4 by adding the 4-element edge E = {v1, v2, v3, v4}. As the load of E equals 1 after the first move in the game, it is easy to see that τ∗g (F) = τ∗g (C4) = 5/2, while the removal of all 2-element edges results in a one-edge hypergraph F ′ with τ∗g (F ′) = 1. 3 Some basic facts and the Continuation Principle In this section we first observe some simple inequalities which are analogous to the ones in other games concerning graph domination and hypergraph transversal, most notably to the fractional domination game [15]. Proposition 3.1. (i) For every hypergraph H, it holds that τ∗(H) ≤ τ∗g (H) < 2τ∗(H) and τ∗(H) ≤ τ∗g ′(H) < 2τ∗(H) + 1. (ii) There is no universal constant C with τg(H) ≤ C · τ∗g (H), and not even with τ(H) ≤ C · τ∗g (H). The same holds true for τ∗g ′(H), too. Proof. No matter which player starts the game, at the end the cover function tq is a frac- tional transversal. This implies the lower bounds τ∗g (H) ≥ τ∗(H) and τ∗g ′(H) ≥ τ∗(H). Concerning a fractional transversal game G on H and the upper bounds in (i), we can write the value of the game in the form |G| = W + W ′, where W and W ′ denote the total sum of weights assigned by Edge-hitter and Staller, respectively. To keep the claimed bounds, first Edge-hitter can fix an optimal fractional transversal f , i.e. one with |f | = Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 459 τ∗(H). After that, in his moves he can apply the strategy to play submoves (vij , wj) with the largest possible weights wj which are not only allowed by (∗) but also respect the inequalities ti−1(vij ) + wj ≤ f(vij ). If such a legal submove with a positive weight does not exist anymore, then H is fully covered and the game is finished. This strategy yields W ≤ τ∗(H), with strict inequality if the game is finished by Staller. We also have W ′ ≤ W or W ′ ≤ W + 1, depending on whether the first move is made by Edge-hitter or Staller, both with strict inequalities if the game is finished by Edge-hitter. Since only one of the players can make the last move, the claimed strict upper bounds follow. For the proof of (ii) we apply the following result of Alon [1]: For every ϵ > 0 and for any sufficiently large k, there is a k-uniform hypergraph H = (V, E) such that τ(H) ≥ (1 − ϵ) ln kk (|V | + |E|). On the other hand, a very simple fractional transversal f with |f | = |V |/k may be constructed by assigning f(v) = 1/k to each vertex v ∈ V . Therefore, τ∗(H) ≤ |V |k and we obtain (1/2− ϵ) ln k < sup H τ(H) 2τ∗(H) < sup H τ(H) τ∗g (H) ≤ sup H τg(H) τ∗g (H) due to the obvious fact τ ≤ τg and the inequality τ∗g < 2τ∗ from (i). For τ∗g ′(H) the proof is similar, by the second part of (i). Proposition 3.2. The upper bounds in Proposition 3.1(i) are tight apart from an additive constant at most 2. Proof. Consider the complete bipartite graph G = Kk,k2 on k+k2 vertices as a 2-uniform hypergraph. Clearly, τ∗(G) = k. In any submove of a fractional transversal game, while G is not fully covered, Staller can always select a vertex from the bigger partite class. Following this strategy, during k − 1 moves, Staller increases the sum of the loads by at most (k − 1)k. As G has maximum degree k2, k − 1 moves of Edge-hitter increase the loads by at most (k−1)k2. Hence, no matter whether Edge-hitter or Staller starts the game, after 2k − 2 moves we have∑ E∈E ℓ2k−2(E) ≤ (k − 1)k + (k − 1)k2 = k3 − k < |E| , therefore the game is not over yet. This shows τ∗g (G) > 2k − 2 = 2τ∗(G) − 2 and, similarly, τ∗g ′(H) ≥ 2τ∗(G)− 1 follows if Staller starts the game. A monotone property of the game fractional transversal number is expressed in the following idea, which provides a useful tool in simplifying several arguments. Let a hy- pergraph H with a pre-defined load function ℓ be given, which we consider as a non-zero starting configuration. We ask about the value |G| of the game started by Edge-hitter, where the game is finished when ℓ is completed to a load function under which H is fully covered. The rules are the same as they were in the case of ℓ0 ≡ 0, but here we have ℓ0 = ℓ, while the value of the game is still computed by starting with the cover function t0 ≡ 0. Under these conditions and assuming that Edge-hitter starts the fractional transversal game G on hypergraph H with the pre-defined ℓ, we consider the sets UH|ℓ = {a ∈ R : Edge-hitter has a strategy that ensures |G| ≤ a} , LH|ℓ = {b ∈ R : Staller has a strategy that ensures |G| ≥ b}. 460 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 Then the game fractional transversal number with predefined load function ℓ is defined as τ∗g (H|ℓ) = inf(UH|ℓ). It can be shown with a proof analogous to that of Proposition 2.1 that inf(UH|ℓ) = sup(LH|ℓ) always holds. The corresponding value τ∗g ′(H|ℓ) is defined analogously for the Staller-start game on H|ℓ. The imagination strategy is a useful technique applied in many proofs when domination and transversal games are considered (see e.g. [3, 6, 9, 15, 17, 20]). It was introduced and first used in [7]; for a detailed explanation and examples we refer the reader to [5, Chapter 2.2]. Theorem 3.3 (Continuation Principle). If ℓ and ℓ′ are load functions on the hypergraph H = (V, E) such that ℓ(E) ≤ ℓ′(E) holds for every E ∈ E , then τ∗g (H|ℓ) ≥ τ∗g (H|ℓ′), and similarly τ∗g ′(H|ℓ) ≥ τ∗g ′(H|ℓ′). Proof. Assume for a contradiction that τ∗g (H|ℓ) < τ∗g (H|ℓ′), and choose two reals t1, t2 with τ∗(H|ℓ) < t1 < t2 < τ∗g (H|ℓ′). We use the imagination strategy between the following two games: Game 1: Edge-hitter plays on H|ℓ applying a strategy which ensures that the value of the game is at most t1. Staller’s moves in Game 1 are defined according to her moves in Game 2. Game 2: Staller plays on H|ℓ′ applying a strategy which ensures that the value of the game is at least t2. Edge-hitter’s moves in Game 2 are defined according to his moves in Game 1. Assume that Edge-hitter starts the game. He chooses his first move in Game 1 according to the prescribed strategy. We copy this move into Game 2 if it is a legal move there, or choose an appropriate replacement move for Edge-hitter in Game 2. In the next turn, Staller replies with a move according to her prescribed strategy in Game 2 and we copy the same move into Game 1. (We will see that it is always legal.) Then, Edge-hitter replies in Game 1 and we copy it or make the according move for Edge-hitter in Game 2. The two parallel games continue this way until at least one of them is finished. The moves essentially are copied (or interpreted) between Game 1 and Game 2 such that ℓ(E) ≤ ℓ′(E) remains true for all E ∈ E after every move. If this inequality is valid before Staller’s move in Game 2, then her next move in H|ℓ′ is legal in H|ℓ as well, so that we can simply copy it into Game 1. The condition ℓ(E) ≤ ℓ′(E) for all E ∈ E clearly remains valid for the new load functions. Suppose now that the inequality ℓ(E) ≤ ℓ′(E) is true for all E ∈ E before Edge- hitter’s move in Game 1. If it is legal, we simply copy it into Game 2 and the inequality remains valid. In the other case one or more submoves (vik , wk) made in Game 1 are not legal in Game 2. We then choose the maximum w′k such that (vik , w ′ k) is a legal submove in Game 2. The remaining weight wk − w′k can be distributed between arbitrary vertices such that the submoves are legal. Observe, however, that if this happens, all loads on the edges incident with vik reach 1 after the submove (vik , w ′ k) in Game 2. We infer that ℓ(E) ≤ ℓ′(E) remains true for all E ∈ E after Edge-hitter’s move. It follows that the loads will never become smaller in Game 2 than the corresponding ones in Game 1. Thus, the values g1 and g2 of Games 1 and 2 satisfy g1 ≥ g2. By the strategies of the players, it is true that t1 ≥ g1 and g2 ≥ t2. We therefore obtain t1 ≥ g1 ≥ g2 ≥ t2 > t1 Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 461 and this contradiction proves τ∗g (H|ℓ) ≥ τ∗g (H|ℓ′). The analogous conclusion can be reached in the Staller-start game as well, literally by the same argument, deriving a contradiction from the assumption τ∗g ′(H|ℓ) < τ∗g ′(H|ℓ′). We obtain the following immediate consequence. Theorem 3.4. The game fractional transversal numbers for the Staller-start and for the Edge-hitter-start games on H may differ by at most 1. Proof. Consider the Staller-start game. Whatever Staller moves first, she assigns total weight 1, and creates a situation which is at least as favorable for Edge-hitter as the all- zero load at the beginning of the original transversal game. Then, due to Theorem 3.3, Edge-hitter can ensure that the game ends using at most τ∗g (H) further weight. This proves τ∗g ′(H) ≤ τ∗g (H) + 1. Similarly, if Edge-hitter starts, after his first move he is in at least as favorable position as with the all-zero load at the beginning of the Staller-start game. This proves the reverse inequality τ∗g (H) ≤ τ∗g ′(H) + 1. 4 Infinite moves are not necessary The definition of a legal move in the transversal game admits the option that a player splits the value 1 into an infinite number of pieces; e.g., wk = 2−k. It turns out, however, that each legal move on H = (V, E) is equivalent to a move which consists of at most |V | submoves. Theorem 4.1. Every legal move in a fractional transversal game can be replaced with a legal move such that each vertex occurs in at most one submove of it and the two moves result in the same load function. Proof. First, consider a vertex v which occurs in two different submoves (vij , wj) and (vik , wk) of a move. That is, v = vij = vik and we may assume j < k. By the condi- tion (∗), there exists an edge E ∈ E such that v ∈ E and the second submove (vik , wk) increases the load of E by exactly wk. If the submove (vij , wj) is deleted from the se- quence and the weight wk is replaced by wj +wk in the kth submove, the submove and the whole move remain legal and result in the same load function as before. Performing this modification repeatedly we can achieve that every vertex occurs in either zero or exactly one or infinitely many submoves of the move in question. This already proves the statement if the move contains only a finite number of submoves. Now, assume that the move is infinite. Then, the sequence of submoves can be split into two, such that the first part is finite, and in the second infinite part every vertex (which is present there) is repeated infinitely many times. Consider this infinite subse- quence S = (vis , ws), . . . . By renaming the vertices of H if necessary, we may assume that {v1, . . . , vℓ} is the set of the vertices which are present in S. We prove that the finite sequence S′ = (v1, ∑ j: ij=1 wj), . . . , (vℓ, ∑ j: ij=ℓ wj) is equivalent to S. Clearly, S and S′ yield the same load function after the move. So, it is enough to prove that S′ is legal. Assume for a contradiction that (vk, ∑ j: ij=k wj) is not a legal submove in S′, and let k be the smallest such index. Then, after the (k − 1)st submove of S′, every edge E which contains vk has a load ℓ(E) > 1− ∑ j: ij=k wj and, moreover, there is a positive constant 462 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 ϵ such that minvk∈E ℓ(E) + ∑ j: ij=k wj = 1 + ϵ. Now, consider S again. There is an index p = p(ϵ) such that ∑ j≥p wj < ϵ and hence, before the p th submove of S, each edge containing vk is fully covered. As vk occurs infinitely often in S, and also the occurrences after the pth submove are legal, this is a contradiction. 5 Edge-hitter’s moves are transposable We say that a finite move (vi1 , w1), . . . , (vik , wk) is transposable if for any permutation β(1), . . . , β(k) of 1, . . . , k, the move (viβ(1) , wβ(1)), . . . , (viβ(k), wβ(k)) is legal. We will show that from the point of view of Edge-hitter, we can restrict our attention to transposable moves. Note that every transposable move is legal, but not conversely. We first give a characterization of transposable moves: Lemma 5.1. A move (vi1 , w1), . . . , (vik , wk), where wj > 0 for all 1 ≤ j ≤ k and no vertices are repeated, is transposable if and only if the total weight constraint (∗∗) is satisfied, and after performing the entire move, for every vertex vij , min E∋vij ∑ vi∈E t(vi) ≤ 1. (5.1) Proof. A legal move must satisfy the two constraints (∗) and (∗∗). The total weight constraint (∗∗) for a legal move is explicitly required in the lemma, and it is insensitive to permuting the submoves. Let us turn to (∗). If ∑ vi∈E t(vi) ≤ 1 for an edge E containing vij , then omitting the submove (vij , wj) from the move we obtain ∑ vi∈E\{vij } t(vi) ≤ 1−wj . Hence (vij , wj) is a legal submove no matter when it is performed during the move. This means that the move is transposable whenever the condition (5.1) is satisfied for all j. In the other direction, assume that for some vij , the left-hand side of (5.1) is bigger than 1. Consider a permutation in which (vij , wj) is the last submove. Then (vij , wj) is not legal because (∗) is violated. Consequently the move is not transposable. Theorem 5.2. If a finite legal move m = (vi1 , w1), . . . , (vik , wk) is not transposable in the fractional transversal game, then it can be replaced by a transposable (and legal) move after which no edge gets smaller load than after m. Proof. First, consider the legal move m = (vi1 , w1), . . . , (vik , wk) and the move m ′ = (vi2 , w2), . . . , (vik , wk), (vi1 , w1) that is obtained by the cyclic permutation β = 2, . . . , k, 1. It is clear that condition (∗) in the definition remains true for the first k − 1 submoves of m′ and consequently, these submoves are legal. For the last (and not nec- essarily legal) submove, determine w∗1 as the maximum weight which results in a legal submove with vi1 . If w ∗ 1 ≥ w1, then m′ is legal and gives exactly the same load function as m. If w∗1 < w1, then the same load function is obtained after the submove (vi1 , w ∗ 1) as after m, because in both cases every edge incident with vi1 is fully covered and the loads of the other edges are unchanged. In the latter case, the sum of the weights is decreased by w1−w∗1 . After this change, the submove (vi1 , w ∗ 1) will be legal in any permutation of (vi1 , w ∗ 1), (vi2 , w2), . . . , (vik , wk). That is, if a permutation is not legal after this replacement, this is due to another vertex. The same is true if some weights ws are replaced by smaller weights. Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 463 We repeat this step for the modified sequence with permutation β = 3, . . . , k, 1, 2, then with β = 4, . . . , k, 1, 2, 3, and so on, finally with β = k, 1, 2, . . . , k − 1, keeping all modifications incrementally. Decreasing the weight of the last submove in each step if necessary, at the end a legal transposable move m∗ is obtained, which yields the same load function as m and satisfies ∑k j=1 w ∗ j ≤ ∑k j=1 wj . If ∆ = ∑k j=1 wj − ∑k j=1 w ∗ j is positive, we use the weight ∆ to increase the loads of some non-fully covered edges in an arbitrary way. The total absolute change of weights is 2∆, or possibly less if the game is over. Finally, we normalize the move by eliminating any multiple occurrences of vertices, using Theorem 4.1. The redistribution of ∆ and the normalization may lead to a move that is not transpos- able. If so, we repeat the modification described above. If the process terminates after a finite number of iterations, then the last version of the move is transposable by construction, and the proof is complete. Otherwise we obtain an infinite sequence ∆(1),∆(2), . . . of re-distributions from the total unit weight of the move. The total load of edges increases by at least ∑ i≥1 ∆ (i); hence this sum converges because altogether the total load is at most |E|. On the other hand, the weight of a vertex changes by at most ∆(i) in the ith iteration, hence the local changes in weight (at least one of which is negative in each iteration) in absolute value sum up to at most 2 ∑ i≥1 ∆ (i), and therefore their sum also converges. Let m(i) denote the move constructed in the ith iteration, before the re-distribution of the weight ∆(i). We know that this move is transposable. Let w(i)v denote the weight used for vertex v in the corresponding submove of m(i), and set w(i)v = 0 if v does not appear in this move. We have shown that the limit of these weights w∗v = limi→∞ w (i) v = 0 exists. Let p denote the number of vertices with a positive weight in w∗. Relabeling these vertices in an arbitrary order, we define a move m∗ = (v1, w∗1), . . . , (vp, w ∗ p) with p submoves. By continuity, the loads achieved by the moves m(i) converge to the corresponding loads after the move m∗. We claim that m∗ is transposable. This will complete the proof because the loads never decrease, hence under w∗ no edge gets smaller load than by move m. For showing that m∗ is transposable, we use the characterization of Lemma 5.1. The inequality (5.1) follows from the fact that its left-hand side is a continuous function of the weights. Let us finally check that the total weight constraint (∗∗) is satisfied: Since ∑ v w (i) v ≤ 1 for all i, this inequality is satisfied in the limit. If in some iteration, the current move terminates the game (i.e., each edge gets load at least 1 while the total weight in the move is at most 1), this will hold in all successive moves, since the loads never decrease, and hence it holds also for m∗, by continuity. Therefore, we need to show ∑ v w ∗ v = 1 only when none of the moves terminate the game. In this case, the difference 1 − ∑ v w (i) v is bounded by ∆i, which goes to 0, and hence ∑ v w ∗ v = 1 in the limit. Remark 5.3. Based on Theorem 5.2, Edge-hitter may restrict his strategy to transposable moves. On the other hand, the result suggests that Staller is advised to perform moves, if possible, which are ‘very non-transposable’ in a sense. 464 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 6 Algorithm for computing the value of the game We consider an equivalent version of the game, the structured game, which is easier to analyze. Each move consists of n + 1 rounds. Each round consists of n submoves, which are dedicated to the vertices v1, . . . , vn in succession. In each submove, the player whose turn it is can decide the amount w, the weight spent in the submove, by which the cover value of t(vi) is increased, subject to the usual rules: The increase must be useful, i.e. each submove must satisfy the condition (∗), and it must be within the budget constraint of total weight 1 to be spent per move. It is possible to skip a submove by simply choosing w to be zero. The first n rounds are identical, but the last round is special: In each submove, the weight is greedily chosen as the largest possible legal weight, hence not allowing any free- dom in choosing w for the player in those submoves. This ensures that the whole move spends a total weight of 1 unless a cover is obtained. There are n moves, alternating between the two players. This is enough to ensure that a cover is constructed when the game terminates. Every move consists of n2 + n submoves, and in total, the game consists of N = n3 + n2 submoves. We illustrate this for a small example with n = 4 vertices, where the Edge-hitter starts. The sequence of submoves is H1, H2, H3, H4; H1, H2, H3, H4; H1, H2, H3, H4; H1, H2, H3, H4; G1, G2, G3, G4; S1, S2, S3, S4; S1, S2, S3, S4; S1, S2, S3, S4; S1, S2, S3, S4; G1, G2, G3, G4; H1, H2, H3, H4; H1, H2, H3, H4; H1, H2, H3, H4; H1, H2, H3, H4; G1, G2, G3, G4; S1, S2, S3, S4; S1, S2, S3, S4; S1, S2, S3, S4; S1, S2, S3, S4; G1, G2, G3, G4. Here Hi denotes a move of Edge-hitter for vertex i, and Si denotes a move of Staller for vertex i. The greedy moves are denoted by Gi. We do not stipulate as part of the rules that the whole budget of 1 unit must be spent during a move. This capacity is only an upper bound. It is still true that the whole budget is spent in each move if the game is played from the beginning. However, this arises as a consequence of the new setup, due to the greedy moves. As soon as a cover is found, the rules imply that no more weight can be spent, and thus the game is effectively over. Lemma 6.1. The structured game has the same value as the original game. Proof. According to Theorem 4.1, we can assume that every vertex occurs at most once in a move. We can realize this in the structured game by selecting one vertex per round and leaving the weight at 0 otherwise. Thus, the structured game does not restrict the players’ strategies, when compared to the original game. On the other hand, the structured game does not give the players more power: The greedy moves ensure that the total weight of 1 is used as long as it is possible. Example for the structured game. Consider an Edge-hitter-start game on C4 with strate- gies of the players as described in Section 2.1. A corresponding structured game, where we follow the idea in the proof of Lemma 6.1, can be given as follows. Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 465 Edge-hitter’s first move: (v1, 14 ), (v2, 0), (v3, 0), (v4, 0); (v1, 0), (v2, 14 ), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 14 ), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 14 ); (v1, 0), (v2, 0), (v3, 0), (v4, 0). Staller’s first move: (v1, 0), (v2, 0), (v3, 12 ), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 12 ); (v1, 0), (v2, 0), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 0). Edge-hitter’s second move: (v1, 0), (v2, 16 ), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 0); (v1, 0), (v2, 0), (v3, 0), (v4, 0); (v1, 1 3 ), (v2, 0), (v3, 0), (v4, 0). For demonstration, the last move makes use of a greedy submove. By Theorem 5.2, it is not a disadvantage for Edge-hitter restricting himself to transposable moves. Consequently, his submoves can always be performed in the order (v1, w1), . . . , (vk, wk), and a single round H1, H2, . . . ,Hn followed by n greedy sub- moves would be a sufficient model for Edge-hitter’s moves. For simplicity, we have how- ever chosen to treat the two players uniformly. Note that the greedy submoves are necessary also in case of Edge-hitter. Otherwise, for example, he might pass on the first move and transform the game to the Staller-start version, which sometimes admits a smaller game value (as in the example of C4). Consider the situation after the jth submove, 0 ≤ j ≤ N . Let x⃗ ∈ [0, 1]E be an arbitrary load vector, and let r ∈ [0, 1] be the budget for the current move that is still available. If j is written in the form j = k(n2 + n) + i, i.e. k = ⌊ jn2+n⌋ and 0 ≤ i < n 2 + n, then 1− r is the total weight spent in the last i submoves. We define Tj(x⃗, r) as the sum of weights spent during the remaining part of the game, if both players play optimally, starting from the current situation. If j is large and the entries of x⃗ are small, it may happen that a complete fractional cover is not reached, because the game neces- sarily ends after the nth round. Nevertheless, we have chosen our definition because it makes Tj well-defined for arbitrary x⃗ and r. (The definition of Tj(x⃗, r) is related to the game fractional transversal number with predefined load function used in the proof of the Continuation Principle (Theorem 3.3)6.) The value of the original game is T0(⃗0, 1). We will derive a backward recursion for the functions Tj , and thus show that they are piecewise linear and continuous. 6In particular, if the jth submove is the last submove in a move of Staller and ℓ is the load function corre- sponding to x⃗, then we would expect Tj(x⃗, 1) to be τ∗g (H|ℓ). However this does not hold in general because, as just discussed, the structured game may terminate too early. 466 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 Given j, we know the type of the jth submove (H , S, or G) and the vertex vi to which it applies. We denote the maximum permitted weight by wmaxi (x⃗, r) = min{r,max{ 1− xE | E ∋ vi }} , (6.1) where xE denotes the entry of x⃗ that corresponds to the edge E ∈ E . For the result of increasing the cover value of vi by w we write updatei(x⃗, w) = x⃗ ′ with x′E = { xE , if vi /∈ E, min{1, xE + w}, if vi ∈ E. (6.2) With these definitions, the recursion for a submove Hi for Edge-hitter can be written easily: Tj−1(x⃗, r) = min{w + Tj(updatei(x⃗, w), r − w) | 0 ≤ w ≤ wmaxi (x⃗, r) } (6.3) If the submove is for Staller (Si), the recursion is the same as (6.3), except that min is replaced by max. In the greedy submoves Gi, we always choose w = wmaxi (x⃗, r): Tj−1(x⃗, r) = w max i (x⃗, r) + Tj(updatei(x⃗, w max i (x⃗, r)), r − wmaxi (x⃗, r)) (6.4) The last greedy submove Gn of each move is an exception: Since a different move is about to start, the budget r is reset to 1. Thus, when j is a multiple of n2 + n, then Tj−1(x⃗, r) = w max n (x⃗, r) + Tj(updaten(x⃗, w max n (x⃗, r)), 1). (6.5) As the recursion anchor, we use the value after the final move, which is simply TN (x⃗, r) = 0. (6.6) Theorem 6.2. Each function Tj(x⃗, r) for 0 ≤ j ≤ N is a piecewise linear continuous function with finitely many linear pieces defined on [0, 1]E × [0, 1]. Moreover, all Tj are rational in the sense that each linear piece has rational coefficients and rational constant part. As a consequence, the boundaries between regions of the domain with different linear functions can be described by linear equations with rational coefficients. Proof. We will call a function with all the desired properties — piecewise linear, continu- ous, and rational, with finitely many linear pieces — a PLCR function. The proof proceeds by backward recursion from TN down to T0. The function TN from (6.6) is obviously PLCR. The sum, difference, maximum, or minimum of two PLCR functions is again PLCR, and the same holds true when substituting one PLCR function into another. It follows directly that the functions wmaxi and updatei are PLCR functions on the domain [0, 1] E × [0, 1]. This allows us to perform the induction step in the recursions (6.4)–(6.5) for Gi. In the recursion (6.3) we additionally have a minimization (or, in the analogous recur- sion for Staller, a maximization) over some range of values w. It has the form min{F (x⃗, r, w) | 0 ≤ w ≤ wmaxi (x⃗, r) } with the PLCR function F (x⃗, r, w) := w + Tj(updatei(x⃗, w), r − w) Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 467 To get rid of the varying upper bound on w, we rewrite the recursion in terms of another PLCR function F̂ (x⃗, r, w) = F (x⃗, r,min{w,wmaxi (x⃗, r)} as Tj−1(x⃗, r) = min{ F̂ (x⃗, r, w) | 0 ≤ w ≤ 1 }. Lemma 6.3 below establishes that Tj−1 is a PLCR function. The same argument applies to the recursion for Staller (Si), where min is replaced by max. Lemma 6.3. Suppose that F̂ (y, w) : [0, 1]m × [0, 1] → R is a PLCR function. Then the function T (y) : [0, 1]m → R defined by minimizing over w: T (y) := min{ F̂ (y, w) | 0 ≤ w ≤ 1 } (6.7) is also a PLCR function. Proof. We first show that T is continuous. Since F̂ is PLCR, it is Lipschitz-continuous. Let L be its Lipschitz constant with respect to the ∞-norm. (We can compute L as the maximum L1-norm of all coefficient vectors of the linear pieces of F̂ .) It follows that the function T in (6.7) is also Lipschitz-continuous with Lipschitz-constant L. To see this, let ∥y0 − y1∥ ≤ ε, and let T (y0) = F̂ (y0, w0) for some w0. Then T (y1) ≤ F̂ (y1, w0) ≤ F̂ (y0, w0) + Lε = T (y0) + Lε. The converse bound T (y0) ≤ T (y1) + Lε follows in the same way. We still need to show that T is piecewise linear. For an intuitive way to see this, one can interpret the minimization over w geometrically. The graph of F̂ : [0, 1]m× [0, 1] → R is a subset of Rm+2. Taking the minimum over all w amounts to projecting away the coordinate corresponding to w and taking the lower envelope (with respect to the last coordinate) in the projection in Rm+1. Figure 2 shows a two-dimensional illustration. This picture can also be interpreted as a three-dimensional view of the graph of a bivariate function F̂ (y, w) when the viewing direction is parallel to the w-axis. (In this hypothetical example, the resulting minimum is discontinuous; this cannot happen when F̂ is continuous and its domain is the box [0, 1]m × [0, 1].) Formally, we conduct the proof as follows. We know that the domain [0, 1]m+1 of F̂ splits into finitely many rational convex (m + 1)-dimensional polytopes P on which F̂ is linear: F̂ (y, w) = aP y + bPw + cP , for (y, w) ∈ P for some rational coefficient vector aP and rational coefficients bP and cP . We can thus write T (y) as the minimum of finitely many functions TP (y) of the form TP (y) := min{ aP y + bPw + cP | 0 ≤ w ≤ 1, (y, w) ∈ P }, (6.8) where the minimum of an empty set is taken as ∞. For fixed y, the minimum in (6.8) depends on the sign of bP . If bP > 0, the minimum is achieved on a boundary point that lies on some facet P ′ of P whose outer normal has negative w-coordinate. On such a facet, w can be expressed as a linear function of y, and thus, TP can be written as a linear function TP ′(y) = aP ′y + cP ′ , for y ∈ P̄ ′, (6.9) 468 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 y { (y, F̂ (y, w)) | 0 ≤ w ≤ 1 } F̂ 0 1 Figure 2: The lower envelope of a polyhedral set in 2 dimensions (m = 1). where P̄ ′ is the projection of the facet P ′ to [0, 1]m. Thus, TP (y) is the minimum of finitely many functions TP ′(y), with the understanding that TP ′(y) is taken as ∞ when y is outside its domain P̄ ′. The situation is similar for bP < 0. When bP = 0, then F̂ does not depend on w and we can simply write TP (y) = aP y + cP , for y ∈ P̄ , (6.10) where P̄ is the projection of P . In summary, the function T (y) can be written as the minimum of finitely many pieces TP (y), each of which can in turn be written as the minimum of finitely many linear pieces (6.9) or (6.10). All these pieces have rational coefficients and rational domain boundaries, and since continuity of T has already been established, the PLCR property of T follows. The proof of Theorem 6.2 is constructive and, in principle, it provides an algorithm for computing the value T0(⃗0, 1) of the game. From this, we obtain the following important corollary. Theorem 6.4. For every finite hypergraph H = (V, E), the game fractional transversal number τ∗g (H) and its Staller-start version τ∗g ′(H) are rational. Moreover, each player in every step has an optimal move with only rational weights, provided that the weights in all previous submoves were rational. Remark 6.5. It is not true in general that every optimal strategy uses only rational weights. A simple counterexample is the graph C4 (Example 1 from Section 2.1), where Staller can start by placing x and 1 − x on two vertices with any x ∈ [0, 1], no matter if x is rational or irrational. 7 Concluding remarks and open problems Putting the fractional domination game [15] into a more general context, in this paper we introduced the fractional transversal game on hypergraphs. Among other results, we Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 469 proved that the game value is rational, and both players have optimal strategies using ratio- nal weights and with a finite number of submoves. Since a dominating set of a graph is a transversal of the closed neighborhood hypergraph, and a total dominating set is a transver- sal of the open neighborhood hypergraph, the following consequence is immediate. Theorem 7.1. The fractional versions of both the domination game and the total domi- nation game have rational game values (game fractional domination number and game fractional total domination number) on every graph. We conclude this paper with some conjectures and open questions. Conjecture 7.2. If each of the first 2k − 1 (k ≥ 1) moves was an integer move in the fractional transversal game, i.e. of the form (vi1 , 1), then Staller has an integer move in the (2k)th turn, which is optimal in the fractional transversal game. This means that fractional moves would be advantageous for Edge-hitter only. If true then this conjecture implies the following weaker one. Conjecture 7.3. For every hypergraph H , τ∗g (H) ≤ τg(H). Perhaps the following stronger version of Conjecture 7.2 is also true. Conjecture 7.4. Starting from any cover function, there is an optimal strategy for Staller where, in every submove, she always spends the largest legal weight. These conjectures could be approached by implementing the algorithm that is implicit in the proof of Theorem 6.2 by computer. We have not derived an estimate for the com- plexity (number of pieces) of the piecewise linear continuous functions Tj(x⃗, r) that are involved in the construction. If the growth of the complexity is not too steep, there is hope to solve some examples of moderate size, beyond the range of small examples that we considered in Section 2.1, and this could shed some light on the conjectures. One would naturally expect that T0(x⃗, r) is monotonically decreasing in x⃗, for fixed r, and moreover, that it is Lipschitz-continuous with Lipschitz constant 1. In other words, in every linear piece, the coefficient of each variable xi is between 0 and −1. We have not explored these properties. Acknowledgement We are grateful to Gábor Tardos for helpful discussions. ORCID iDs Csilla Bujtás https://orcid.org/0000-0002-0511-5291 Günter Rote https://orcid.org/0000-0002-0351-5945 Zsolt Tuza https://orcid.org/0000-0003-3235-9221 References [1] N. Alon, Transversal numbers of uniform hypergraphs, Graphs Comb. 6 (1990), 1–4, doi: 10.1007/BF01787474, https://doi.org/10.1007/BF01787474. 470 Ars Math. Contemp. 24 (2024) #P3.05 / 453–471 [2] M. Borowiecki, A. Fiedorowicz and E. Sidorowicz, Connected domination game, Appl. Anal. Discrete Math. 13 (2019), 261–289, doi:10.2298/AADM171126020B, https://doi.org/ 10.2298/AADM171126020B. [3] B. Brešar, Cs. Bujtás, T. Gologranc, S. Klavžar, G. Košmrlj, T. Marc, B. Patkós, Zs. Tuza and M. Vizer, The variety of domination games, Aequationes Math. 93 (2019), 1085–1109, doi:10. 1007/s00010-019-00661-w, https://doi.org/10.1007/s00010-019-00661-w. [4] B. Brešar, P. Dorbec, S. Klavžar and G. Košmrlj, Domination game: effect of edge- and vertex- removal, Discrete Math. 330 (2014), 1–10, doi:10.1016/j.disc.2014.04.015, https://doi. org/10.1016/j.disc.2014.04.015. [5] B. Brešar, M. A. Henning, S. Klavžar and D. F. Rall, Domination Games Played on Graphs, SpringerBriefs in Mathematics, Springer, Cham, 2021, doi:10.1007/978-3-030-69087-8, https://doi.org/10.1007/978-3-030-69087-8. [6] B. Brešar, S. Klavžar, G. Košmrlj and D. F. Rall, Guarded subgraphs and the domination game, Discrete Math. Theor. Comput. Sci. 17 (2015), 161–168, doi:10.46298/dmtcs.2123, https: //doi.org/10.46298/dmtcs.2123. [7] B. Brešar, S. Klavžar and D. F. Rall, Domination game and an imagination strategy, SIAM J. Discrete Math. 24 (2010), 979–991, doi:10.1137/100786800, https://doi.org/10. 1137/100786800. [8] Cs. Bujtás, General upper bound on the game domination number, Discrete Appl. Math. 285 (2020), 530–538, doi:10.1016/j.dam.2020.06.018, https://doi.org/10.1016/j. dam.2020.06.018. [9] Cs. Bujtás, M. A. Henning and Zs. Tuza, Transversal game on hypergraphs and the 3 4 - conjecture on the total domination game, SIAM J. Discrete Math. 30 (2016), 1830–1847, doi: 10.1137/15M1049361, https://doi.org/10.1137/15M1049361. [10] Cs. Bujtás, M. A. Henning and Zs. Tuza, Bounds on the game transversal number in hy- pergraphs, European J. Comb. 59 (2017), 34–50, doi:10.1016/j.ejc.2016.07.003, https: //doi.org/10.1016/j.ejc.2016.07.003. [11] Cs. Bujtás, V. Iršič and S. Klavžar, Perfect graphs for domination games, Ann. Comb. 25 (2021), 133–152, doi:10.1007/s00026-021-00523-w, https://doi.org/10.1007/ s00026-021-00523-w. [12] Cs. Bujtás, V. Iršič and S. Klavžar, 1/2-conjectures on the domination game and claw-free graphs, European J. Comb. 101 (2022), Paper No. 103467, 17 pp., doi:10.1016/j.ejc.2021. 103467, https://doi.org/10.1016/j.ejc.2021.103467. [13] Cs. Bujtás, B. Patkós, Zs. Tuza and M. Vizer, Domination game on uniform hypergraphs, Dis- crete Appl. Math. 258 (2019), 65–75, doi:10.1016/j.dam.2018.11.013, https://doi.org/ 10.1016/j.dam.2018.11.013. [14] Cs. Bujtás and Zs. Tuza, The disjoint domination game, Discrete Math. 339 (2016), 1985– 1992, doi:10.1016/j.disc.2015.04.028, https://doi.org/10.1016/j.disc.2015. 04.028. [15] Cs. Bujtás and Zs. Tuza, Fractional domination game, Electron. J. Comb. 26 (2019), Paper No. 4.3, 17 pp., doi:10.37236/8730, https://doi.org/10.37236/8730. [16] M. A. Henning and W. B. Kinnersley, Domination game: a proof of the 3/5-conjecture for graphs with minimum degree at least two, SIAM J. Discrete Math. 30 (2016), 20–35, doi: 10.1137/140976935, https://doi.org/10.1137/140976935. [17] M. A. Henning, S. Klavžar and D. F. Rall, Total version of the domination game, Graphs Comb. 31 (2015), 1453–1462, doi:10.1007/s00373-014-1470-9, https://doi.org/10.1007/ s00373-014-1470-9. Cs. Bujtás et al.: Optimal strategies in fractional games: vertex cover and domination 471 [18] M. A. Henning and D. F. Rall, The enclaveless competition game, Ars Math. Contemp. 20 (2021), 129–142, doi:10.26493/1855-3974.2227.e1a, https://doi.org/10.26493/ 1855-3974.2227.e1a. [19] V. Iršič, Effect of predomination and vertex removal on the game total domination number of a graph, Discrete Appl. Math. 257 (2019), 216–225, doi:10.1016/j.dam.2018.09.011, https: //doi.org/10.1016/j.dam.2018.09.011. [20] W. B. Kinnersley, D. B. West and R. Zamani, Extremal problems for game domination number, SIAM J. Discrete Math. 27 (2013), 2090–2107, doi:10.1137/120884742, https: //doi.org/10.1137/120884742. [21] S. Klavžar and D. F. Rall, Domination game and minimal edge cuts, Discrete Math. 342 (2019), 951–958, doi:10.1016/j.disc.2018.12.001, https://doi.org/10.1016/j. disc.2018.12.001. [22] G. Košmrlj, Domination game on paths and cycles, Ars Math. Contemp. 13 (2017), 125–136, doi:10.26493/1855-3974.891.e93, https://doi.org/10.26493/1855-3974.891. e93. [23] K. Xu, X. Li and S. Klavžar, On graphs with largest possible game domination number, Discrete Math. 341 (2018), 1768–1777, doi:10.1016/j.disc.2017.10.024, https://doi.org/10. 1016/j.disc.2017.10.024. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.06 / 473–488 https://doi.org/10.26493/1855-3974.2892.f07 (Also available at http://amc-journal.eu) The Sierpiński domination number* Michael A. Henning † Department of Mathematics and Applied Mathematics, University of Johannesburg, Auckland Park, 2006 South Africa Sandi Klavžar ‡ Faculty of Mathematics and Physics, University of Ljubljana, Slovenia and Institute of Mathematics, Physics and Mechanics, Ljubljana, Slovenia and Faculty of Natural Sciences and Mathematics, University of Maribor, Slovenia Elżbieta Kleszcz Department of Mathematics and Applied Mathematics, University of Johannesburg, Auckland Park, 2006 South Africa and AGH University, Department of Discrete Mathematics, al. Mickiewicza 30, 30-059 Krakow, Poland Monika Pilśniak § AGH University, Department of Discrete Mathematics, al. Mickiewicza 30, 30-059 Krakow, Poland Received 31 May 2022, accepted 8 November 2023, published online 18 July 2024 Abstract Let G and H be graphs and let f : V (G) → V (H) be a function. The Sierpiński prod- uct of G and H with respect to f , denoted by G⊗f H , is defined as the graph on the vertex set V (G) × V (H), consisting of |V (G)| copies of H; for every edge gg′ of G there is an edge between copies gH and g′H of H associated with the vertices g and g′ of G, respec- tively, of the form (g, f(g′))(g′, f(g)). In this paper, we define the Sierpiński domination number as the minimum of γ(G ⊗f H) over all functions f : V (G) → V (H). The up- per Sierpiński domination number is defined analogously as the corresponding maximum. *We would like to thank the reviewer very much for careful reading of the paper and in particular for pointing to a subtle technical detail that led to the reformulation of one of the main theorems. †This research was obtained during the sabbatical visit of the first author at the University of Ljubljana and he thanks them for their kindness in hosting him. The first author also acknowledges that his research visit was supported in part by the University of Johannesburg and the South African National Research Foundation (grants 132588, 129265). ‡S.K. acknowledges the financial support from the Slovenian Research Agency (research core funding No. P1-0297 and projects J1-2452, N1-0285). §Corresponding author. cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 474 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 After establishing general upper and lower bounds, we determine the upper Sierpiński dom- ination number of the Sierpiński product of two cycles, and determine the lower Sierpiński domination number of the Sierpiński product of two cycles in half of the cases and in the other half cases restrict it to two values. Keywords: Sierpiński graph, Sierpiński product, domination number, Sierpiński domination number. Math. Subj. Class. (2020): 05C69, 05C76 1 Introduction Sierpiński graphs represent a very interesting and widely studied family of graphs. They were introduced in 1997 in the paper [15], where the primary motivation for their intro- duction was the intrinsic link to the Tower of Hanoi problem, for the latter problem see the book [11]. Intensive research of Sierpiński graphs led to a review article [12] in which state of the art up to 2017 is summarized and unified approach to Sierpiński-type graph families is also proposed. Later research on Sierpiński graphs includes [2, 3, 6, 19, 23]. Sierpiński graphs have a fractal structure, the basic graphs of which are complete graphs. In 2011, Gravier, Kovše, and Parreau [7] introduced a generalization in such a way that any graph can act as a fundamental graph, and called the resulting graphs gener- alized Sierpiński graphs. We refer to the papers [1, 4, 5, 13, 14, 16, 17, 20, 21, 22, 24] for investigations of generalized Sierpiński graphs in the last few years. An interesting generalization of Sierpiński graphs in the other direction has recently been proposed by Kovič, Pisanski, Zemljič, and Žitnik in [18]. Namely, in the spirit of classical graph products, where the vertex set of a product graph is the Cartesian product of the vertex sets of the factors, they introduced the Sierpiński product of graphs as follows. Let G and H be graphs and let f : V (G) → V (H) be an arbitrary function. The Sierpiński product of graphs G and H with respect to f , denoted by G⊗f H , is defined as the graph on the vertex set V (G)× V (H) with edges of two types: • type-1 edge: (g, h)(g, h′) is an edge of G⊗f H for every vertex g ∈ V (G) and every edge hh′ ∈ E(H), • type-2 edge: (g, f(g′))(g′, f(g)) is an edge of G⊗f H for every edge gg′ ∈ E(G). We observe that the edges of type-1 induce n(G) = |V (G)| copies of the graph H in the Sierpiński product G ⊗f H . For each vertex g ∈ V (G), we let gH be the copy of H corresponding to the vertex g. A type-2 edge joins vertices from different copies of H in G⊗f H , and is called a connecting edges of G⊗f H . A vertex incident with a connecting edge is called a connecting vertex. We observe that two different copies of H in G ⊗f H are joined by at most one edge. A copy of the graph H corresponding to a vertex of the graph G in the Sierpiński product G⊗f H is called an H-layer. Let G and H be graphs and HG be the family of functions from V (G) to V (H). We introduce new types of domination, the Sierpiński domination number, denoted by γS(G,H), as the minimum over all functions f from HG of the domination number of the Sierpiński product with respect to f , and upper Sierpiński domination number, denoted E-mail addresses: mahenning@uj.ac.za (Michael A. Henning), sandi.klavzar@fmf.uni-lj.si (Sandi Klavžar), elzbieta.kleszcz@agh.edu.pl (Elżbieta Kleszcz), pilsniak@agh.edu.pl (Monika Pilśniak) M. A. Henning et al.: The Sierpiński domination number 475 by ΓS(G,H), as the maximum over all functions f ∈ HG of domination number of the Sierpiński product with respect to f . That is, γS(G,H) := min f∈HG {γ(G⊗f H)} and ΓS(G,H) := max f∈HG {γ(G⊗f H)} . In this paper, we initiate the study of Sierpiński domination in graphs. In Section 1.1 we present the graph theory notation and terminology we follow. In Section 2 we discuss general lower and upper bounds on the (upper) Sierpiński domination number. Our main contribution in this introductory paper is to determine the upper Sierpiński domination number of the Sierpiński product of two cycles, and to determine the lower Sierpiński domination number of the Sierpiński product of two cycles in half of the cases and in the other half cases restrict it to two values. 1.1 Notation and terminology We generally follow the graph theory notation and terminology in the books [8, 9, 10] on domination in graphs. Specifically, let G be a graph with vertex set V (G) and edge set E(G), and of order n(G) = |V (G)| and size m(G) = |E(G)|. For a subset S of vertices of a graph G, we denote by G− S the graph obtained from G by deleting the vertices in S and all edges incident with vertices in S. If S = {v}, then we simply write G − v rather than G − {v}. The subgraph induced by the set S is denoted by G[S]. We denote the path, cycle and complete graph on n vertices by Pn, Cn, and Kn, respectively. For k ≥ 1 an integer, we use the notation [k] = {1, . . . , k} and [k]0 = {0, 1, . . . , k}. We generally label vertices of the considered graphs by elements of [n]. In this case, the mod function over the set [n] is to be understood in a natural way, more formally, we apply the following operation for t ≥ 1: t mod∗ n = (t− 1) mod n+ 1. A vertex dominates itself and its neighbors, where two vertices are neighbors in a graph if they are adjacent. A dominating set of a graph G is a set S of vertices of G such that every vertex in G is dominated by a vertex in S. The domination number, γ(G), of G is the minimum cardinality of a dominating set of G. A dominating set of cardinality γ(G) is called a γ-set of G. A thorough treatise on dominating sets can be found in [8, 9]. If S is a set of vertices in a graph G, then we will use the notation G|S to denote that the vertices in the set S are assumed to be dominated and hence γ(G|S) is the minimum number of vertices in the graph G needed to dominate V (G) \ S. We note that it could be that a vertex in S is still a member of a such a minimum dominating set no matter that we do not need to dominate the vertices in S themselves. If S = {x}, then we simply denote G|S by G|x rather than G|{x}. 2 General lower and upper bounds We present in this section general lower and upper bounds on the (upper) Sierpiński domi- nation number. Theorem 2.1. If G and H are graphs, then n(G)γ(H)−m(G) ≤ γS(G,H) ≤ ΓS(G,H) ≤ n(G)γ(H) . 476 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 Proof. Let G ⊗f H be an arbitrary Sierpiński product of graphs G and H and let X be a γ-set of G ⊗f H . Assuming for a moment that all the connecting edges are removed from G⊗f H , we obtain n(G) disjoint copies of H for which we clearly need n(G)γ(H) vertices in a minimum dominating set. Consider now an arbitrary connecting edge e = (g, f(g′))(g′, f(g)) of G⊗fH . If no end-vertex of e lies in X , then clearly γ(G⊗fH−e) = γ(G⊗fH). Similarly, if both end-vertices of e lie in X , then γ(G⊗fH−e) = γ(G⊗fH). Hence the only situation in which e has an effect on γ(G⊗fH) is when (g, f(g′)) ∈ X and (g′, f(g)) /∈ X (or the other way around). But in this case, the effect of the presence of the edge e is that because (g, f(g′)) dominates one vertex of g′H , the edge e might reduce the domination number by 1. That is, each connecting edge can drop the domination number of G ⊗f H by at most 1, which proves the left inequality. The other two inequalities are clear. To show that the lower bound of Theorem 2.1 is achieved, we show later in Theo- rem 3.10 that for n ≥ 3 and k ≥ 1, if we take G = Cn and H = C3k+1 where n ≡ 0 (mod 4), then γS(G,H) = kn = n(G)γ(H) −m(G). The upper bound of Theo- rem 2.1 is obtained, for example, for the Sierpiński product of two complete graphs. More generally, to achieve equality in the upper bound of Theorem 2.1 we require the graph H to have the following property. Theorem 2.2. The equality in ΓS(G,H) ≤ n(G)γ(H) is achieved if and only if there exists a vertex x ∈ V (H) such that γ(H|x) = γ(H). Proof. Suppose that H has a vertex x that satisfies γ(H|x) = γ(H). In this case, we consider the Sierpiński product G ⊗f H with the function f : V (G) → V (H) defined by f(v) = x for every vertex v ∈ V (G). Consequently each connecting edge in the product is of the form (g, x)(g′, x). Thus, if X is a γ-set of G ⊗f H , then |X ∩ V (gH)| = γ(H) because the only vertex of gH that can be dominated from outside gH is (g, x), but we have assume that γ(H|x) = γ(H). Therefore, ΓS(G,H) = n(G)γ(H). For the other implication suppose that γ(H|x) < γ(H) for every vertex x ∈ V (H). For an arbitrary edge g1g2 ∈ E(G), if D corresponds to a γ-set of the product G ⊗f H , then |D ∩ V ((G⊗f H)[V (g1H) ∪ V (g2H)])| ≤ 2γ(H)− 1. Consequently, ΓS(G,H) < n(G)γ(H). To conclude this section we describe large classes of graphs for which the second and the third inequality of Theorem 2.1 are both equality. Proposition 2.3. If G and H are graphs such that ∆(G) < n(H) and γ(H) = 1, then ΓS(G,H) = γS(G,H) = n(G). Proof. Let G and H be graphs such that ∆(G) < n(H) and γ(H) = 1. Thus, by The- orem 2.1, γS(G,H) ≤ ΓS(G,H) ≤ n(G) and it is straightforward that the Sierpiński product of graphs G and H can be dominated by taking one dominating vertex from each H-layer to the dominating set. It remains to show that the inequality γS(G,H) ≥ n(G) also holds. Suppose that γS(G,H) ≤ n(G) − 1. Let D be a dominating set of G ⊗f H , where f is such that it minimizes the domination number. Therefore there is an H-layer, denote it by H ′, of G ⊗f H such that D ∩ V (H ′) = ∅. Since there are at most ∆(G) connecting edges incident with vertices from each H-layer and ∆(G) < n(H),then all the vertices from an H-layer cannot be dominated by the vertices from the neighboring layers. Therefore we have D∩V (H ′′) ̸= ∅ for each H ′′-layer of G⊗f H . Thus γS(G,H) ≥ n(G) and the result follows. M. A. Henning et al.: The Sierpiński domination number 477 3 The Sierpiński domination number of cycles Let us recall firstly the domination number of a path and a cycle. Fact 3.1. For n ≥ 3, γ(Pn) = γ(Cn) = ⌈ n 3 ⌉ . In this section, we shall prove the following results. Theorem 3.2. For n ≥ 3, k ≥ 1, and p ∈ [2]0, γS(Cn, C3k+p) ∈  {kn}; p = 0, {kn, kn+ 1}; p = 1, {kn+ ⌊n 2 ⌋ , kn+ ⌊n 2 ⌋ + 1}; p = 2. Moreover, if n ≡ 0 mod 4, then γS(Cn, C3k+1) = kn and γS(Cn, C3k+2) = kn+ ⌊n 2 ⌋ . Theorem 3.3. For n ≥ 3, k ≥ 1, and p ∈ [2]0, ΓS(Cn, C3k+p) =  kn; p = 0, kn+ ⌈n 3 ⌉ ; p = 1, (k + 1)n; p = 2. In order to prove Theorems 3.2 and 3.3, we consider three cases, depending on the value of p. 3.1 The cycle Cn and cycles C3k+1 To determine ΓS(Cn, C3k+1), we prove a slightly more general result. For this purpose, we define a class of graphs Hk as follows. Definition 3.4. For k ≥ 1, let Hk be the class of all graphs H that have the following properties. (a) γ(H) = k + 1 and γ(H − v) = k for every vertex v ∈ V (H). (b) If x, y ∈ V (H), then there exists a γ-set of H that contains x and y, where x = y is allowed. We show, for example, that for every k ≥ 1, the cycle C3k+1 belongs to the class Hk. Proposition 3.5. For k ≥ 1, the class Hk of graphs contains the cycle C3k+1. Proof. For k ≥ 1, let H ∼= C3k+1. Since γ(Cn) = γ(Pn) = ⌈n/3⌉, property (a) in Defi- nition 3.4 holds. To prove that property (b) in Definition 3.4 holds, let x, y ∈ V (H). Since H is vertex-transitive, every specified vertex belongs to some γ-set of H . In particular, if x = y, then property (b) is immediate. Hence, we may assume that x ̸= y. Let H be the cycle v1v2 . . . v3k+1v1, where renaming vertices if necessary, we may assume that x = v1. Let y = vi, and so i ∈ [3k + 1] \ {1}. Let H ′ = H − N [{x, y}], that is, H ′ is obtained from H by removing x and y, and removing all neighbors of x and y. If H ′ is connected, then H ′ is a path P3(k−2)+j for 478 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 some j where j ∈ [3]. In this case, γ(H ′) = k − 1. If H ′ is disconnected, then H ′ is the disjoint union of two paths Pk1 and Pk2 , where k1+ k2 = 3(k− 2)+1. Thus renaming k1 and k2 if necessary, we may assume that either k1 = 3j1 and k2 = 3j2 + 1 where j1 ≥ 1, j2 ≥ 0, and j1 + j2 = k − 2 or k1 = 3j1 + 2 and k2 = 3j2 + 2 where j1, j2 ≥ 0 and j1 + j2 = k − 3. In both cases, γ(H ′) = ⌈k1/3⌉+ ⌈k2/3⌉ = k − 1. Letting D′ be a γ-set of H ′, the set D = D′ ∪ {x, y} is a dominating set of H of cardinality k + 1 = γ(H), implying that D is a γ-set of H that contains both x and y. Hence, property (b) holds. For n ≥ 3 an integer, a circulant graph Cn⟨L⟩ with a given list L ⊆ {1, . . . , ⌊ 12n⌋} is a graph on n vertices in which the ith vertex is adjacent to the (i + j)th and (i − j)th vertices for each j in the list L and where addition is taken modulo n. For example, for n = 3k + 1 where k ≥ 1 and L = {1}, the circulant graph Cn⟨L⟩ is the cycle C3k+1, which, by Proposition 3.5, belongs to the class Hk. More generally, for n = k(2p+1)+ 1 where k ≥ 1, p ≥ 1, and L = [p], the circulant graph Cn⟨L⟩ belongs to the class Hk. We omit the relatively straightforward proof. These examples of circulant graphs serve to illustrate that for each k ≥ 1, one can construct infinitely many graphs in the class Hk. We determine next the upper Sierpiński domination number ΓS(Cn, H) of a cycle Cn and a graph H in the family Hk. Theorem 3.6. For n ≥ 3 and k ≥ 1, if H ∈ Hk, then ΓS(Cn, H) = kn+ ⌈n 3 ⌉ . Proof. For n ≥ 3 and k ≥ 1, let G ∼= Cn and let H ∈ Hk. Let G be the cycle given by g1g2 . . . gng1. In what follows, we adopt the following notation. For each i ∈ [n], we denote the copy giH of H corresponding to the vertex gi simply by Hi. We proceed further with two claims. The first claim establishes a lower bound on ΓS(Cn, H), and the second claim establishes an upper bound on ΓS(Cn, H). Claim 3.7. ΓS(Cn, H) ≥ kn+ ⌈n 3 ⌉ . Proof. Let f : V (G) → V (H) be a constant function, that is, we select h ∈ V (H) and for every vertex g ∈ V (G), we set f(g) = h. Let DG be a γ-set of G. Thus, |DG| = γ(Cn) = ⌈n/3⌉. By property (a) in Definition 3.4, for every vertex g ∈ V (G), there exists a γ-set of gH that contains the vertex (g, f(g)) = (g, h). If g ∈ DG, let Dg be a γ-set of gH that contains the vertex (g, f(g)) = (g, h), and so |Dg| = γ(H) = k + 1. If g ∈ V (G) \DG, let Dg be a γ-set of gH − (g, f(g)) = gH − (g, h), and so in this case |Dg| = γ(H − h) = γ(H)− 1 = k. Let D = ⋃ g∈V (G) Dg. The set D is a dominating set of G⊗f H , and so γ(G⊗f H) ≤ |D| = γ(G)(k + 1) + (n− γ(G))k = kn+ γ(G) = kn+ ⌈n 3 ⌉ . (3.1) For the fixed vertex h chosen earlier, we note that the set of vertices (g, h) for all g ∈ V (G) induces a subgraph of G ⊗f H that is isomorphic to G ∼= Cn. We denote this M. A. Henning et al.: The Sierpiński domination number 479 copy of G by Gh. Among all γ-sets of G ⊗f H , let D∗ be chosen to contain as many vertices of Gh as possible. Let D∗g = D ∗ ∩ V (gH) for every g ∈ V (G). Further let D∗G = {(g, h) ∈ D∗ : g ∈ V (G)}, that is, D∗G is the restriction of D∗ to the copy of G. If a vertex (g, h) /∈ D∗G and (g, h) is not dominated by D∗G, then D∗g is a γ-set of gH by the minimality of the set D∗. However in this case, we could replace the set D∗g be a γ-set of gH that contains the vertex (g, h) to produce a new γ-set of G ⊗f H that contains more vertices from the copy of G than does D∗, a contradiction. Hence, the set D∗G is a dominating set in the copy of G, and so |D∗g(G)| ≥ γ(G). By the minimality of the set D∗ and by property (a) in Definition 3.4, for each vertex g ∈ V (G), we have |D∗g | = γ(H) = k + 1 if the vertex (g, h) ∈ D∗G and |D∗g | = γ(H − h) = k if the vertex (g, h) /∈ D∗G. Therefore, γ(G⊗f H) = |D∗| = |D∗G|(k + 1) + (n− |D∗G|)k = kn+ |D∗G| ≥ kn+ γ(G) = kn+ ⌈n 3 ⌉ . (3.2) By inequalities (3.1) and (3.2), we have γ(G⊗f H) = kn+ ⌈n 3 ⌉ . (3.3) By equation (3.3), we have ΓS(Cn, H) ≥ γ(G⊗f H) = kn+ ⌈n/3⌉. This completes the proof of Claim 3.7. (2) Claim 3.8. ΓS(Cn, H) ≤ kn+ ⌈n 3 ⌉ . Proof. Let f : V (G) → V (H) be an arbitrary function. Let Hi be the ith copy of H corresponding to the vertex gi of G for all i ∈ [n]. Let D be the dominating set of G⊗f H constructed as follows. Let xiyi+1 be the connecting edge from Hi to Hi+1 for all i ∈ [n], where addition is taken modulo n. Thus, the vertex xi ∈ V (Hi) is adjacent to the vertex yi+1 ∈ V (Hi+1) in the graph G⊗fH , that is, xi = (gi, f(gi+1)) and yi+1 = (gi+1, f(gi)). We note that possibly xi = yi. By property (b) in Definition 3.4, there exists a γ-set of Hi that contains both xi and yi. For i ∈ [n], we define the sets Di,1, Di,2, and Di,3 as follows. Let Di,1 be a γ-set of Hi − xi. Let Di,2 be a γ-set of Hi that contains both xi and yi. Let Di,3 be a γ-set of Hi − yi. We note that |Di,1| = |Di,3| = k and |Di,2| = k + 1. For i ∈ [n], we define the set Di as follows. Di =  Di,1; i ≡ 1 (mod 3) and i ̸= n, Di,2; i ≡ 2 (mod 3) or i ≡ 1 (mod 3) and i = n, Di,3; i ≡ 0 (mod 3). For example, the set D1 dominates all vertices of H1 − x1. The set D2 contains the vertex y2, which is adjacent to the vertex x1 of H1, and contains the vertex x2, which is adjacent to the vertex y3 of H3, implying that D2 dominates the vertex x1 of H1, all vertices of H2, and the vertex y3 of H3. The set D3 dominates all vertices of H3 − y3. Thus, D1 ∪D2 ∪D3 dominates all vertices in V (H1)∪ V (H2)∪ V (H3) in the Sierpiński product G⊗fH . Moreover, |D1|+|D2|+|D3| = k+(k+1)+k = 3k+1. More generally, the set D3j−2∪D3j−1∪D3j dominates all vertices in V (H3j−2)∪V (H3j−1)∪V (H3j) in 480 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 the Sierpiński product G⊗f H for all j ∈ {1, . . . , ⌊n/3⌋}. Moreover, |D3j−2|+ |D3j−1|+ |D3j | = k + (k + 1) + k = 3k + 1. If n ≡ 1 (mod 3), then the set Dn is a γ-set of Hn, and in this case |Dn| = k + 1. If n ≡ 2 (mod 3), then the set Dn−1 ∪Dn dominates all vertices in V (Hn−1) ∪ V (Hn), and in this case |Dn−1|+ |Dn| = k + (k + 1) = 2k + 1. The set D = n⋃ i=1 Di is therefore a dominating set of G⊗f H , implying that γ(G⊗f H) ≤ |D| = n∑ i=1 |Di| = kn+ ⌈n 3 ⌉ . This completes the proof of Claim 3.8. (2) The proof of Theorem 3.6 follows as an immediate consequence of Claims 3.7 and 3.8. As a consequence of Proposition 3.5, we have the following special case of Theo- rem 3.6. Corollary 3.9. For n ≥ 3 and k ≥ 1, ΓS(Cn, C3k+1) = kn+ ⌈n 3 ⌉ . We consider next the Sierpiński domination number of Cn and C3k+1, and show that γS(Cn, C3k+1) = kn if n ≡ 0 (mod 4) and γS(Cn, C3k+1) ∈ {kn, kn+ 1}, otherwise. Theorem 3.10. For n ≥ 3 and k ≥ 1, γS(Cn, C3k+1) ∈ {kn, kn+ 1}. Moreover, if n ≡ 0 mod 4, then γS(Cn, C3k+1) = kn. Proof. For n ≥ 3 and k ≥ 1, let G = Cn and let H = C3k+1. Let G be the cycle given by g1g2 . . . gng1. We adopt our notation employed in our earlier proofs. For notational convenience, we let V (H) = {1, 2, . . . , 3k+1} where vertices i and i+1 are consecutive on the cycle H for all i ∈ [3k + 1] (and where addition is taken modulo 3k + 1, and so vertex 1 and vertex 3k + 1 are adjacent). As before, we denote the copy giH of H corresponding to the vertex gi simply by Hi for each i ∈ [n]. Thus, Hi = C3k+1 is the cycle (gi, 1), (gi, 2), . . . , (gi, 3k + 1), (gi, 1) for all i ∈ [n]. Recall that we denote the connecting edge from Hi to Hi+1 by xiyi+1 for all i ∈ [n], where xi ∈ V (Hi), yi+1 ∈ V (Hi+1), and addition is taken modulo n. Thus, yi = (gi, f(gi−1)) and xi = (gi, f(gi+1)) for all i ∈ [n]. By Proposition 3.5, the graph H belongs to the class Hk. Thus, γ(H) = k + 1 and γ(H − v) = k for every vertex v ∈ V (H). Furthermore, if x, y ∈ V (H) where x = y is allowed, then there exists a γ-set of H that contains x and y. By the elementary lower bound on the Sierpiński domination number given in Theo- rem 2.1, γS(G,H) ≥ n(G)γ(H)−m(G) = kn, noting that here n(G) = m(G) = n and γ(H) = k + 1. It follows that γS(Cn, H) ≥ kn. M. A. Henning et al.: The Sierpiński domination number 481 To complete the proof we are going to prove that γS(Cn, H) ≤ kn+ ⌈n 4 ⌉ − ⌊n 4 ⌋ . Let f : V (G) → V (H) be the function defined by f(gi) = { 1; i mod 4 ∈ {1, 2}, 3; otherwise. for all i ∈ [n] where addition is taken modulo n. Adopting our earlier notation, recall that yi = (gi, f(gi−1)) and xi = (gi, f(gi+1)) for all i ∈ [n]. Let n = 4ℓ + j where j ∈ [3]0 = {0, 1, 2, 3}. We note that f(g4i−3) = f(g4i−2) = 1 and f(g4i−1) = f(g4i) = 3 for all i ∈ [ℓ]. Let Di be the unique γ-set of Hi − yi ∼= P3k which consists of all vertices at distance 2 modulo 3 from yi in the graph Hi for all i ∈ [n], and let D = n⋃ i=1 Di. We note that |Di| = k for all i ∈ [n], and so |D| = kn. For all i ∈ {2, 3, . . . , ℓ − 1}, the following four properties hold. (P1) y4i−3 = (g4i−3, 3) and x4i−3 = (g4i−3, 1). (P2) y4i−2 = (g4i−2, 1) and x4i−2 = (g4i−2, 3). (P3) y4i−1 = (g4i−1, 1) and x4i−1 = (g4i−1, 3). (P4) y4i = (g4i, 3) and x4i = (g4i, 1). Hence for all i ∈ {2, 3, . . . , ℓ − 1}, the vertices xi and yi are at distance 2 in Hi, implying that xi ∈ Di. We consider four cases to determine which properties hold for the boundary conditions (that is for i ∈ {1, ℓ}) and finally to set the upper bound on the domination number in each case. Case 1. n ≡ 0 (mod 4), that is n = 4ℓ. In this case, properties (P1) and (P4) also hold for i = 1 and i = ℓ, respectively. Thus, y1 = (g1, 3) and x1 = (g1, 1), and y4ℓ = (g4ℓ, 3) and x4ℓ = (g4ℓ, 1), implying that x1, x4ℓ ∈ D. The set D is therefore a dominating set of G ⊗f H , and so γ(G ⊗f H) ≤ |D| = kn = kn+ ⌈n/4⌉ − ⌊n/4⌋. Case 2. n ≡ 1 (mod 4), that is n = 4ℓ+ 1. In this case, y1 = x1 = (g1, 1), and y4ℓ+1 = (g4ℓ+1, 3) and x4ℓ+1 = (g4ℓ+1, 1). In particular, property (P4) also holds for i = ℓ, and so x4ℓ+1 ∈ D. The set D ∪ {x1} is therefore a dominating set of G ⊗f H , and so γ(G ⊗f H) ≤ |D| + 1 = kn + 1 = kn+ ⌈n/4⌉ − ⌊n/4⌋. Case 3. n ≡ 2 (mod 4), that is n = 4ℓ+ 2. In this case, y1 = x1 = (g1, 1), and y4ℓ+2 = x4ℓ+2 = (g4ℓ+2, 1). We note that neither x1 nor x4ℓ+2 belong to the set D. The set D ∪ {x1} is a dominating set of G ⊗f H , and so γ(G⊗f H) ≤ |D|+ 1 = kn+ 1 = kn+ ⌈n/4⌉ − ⌊n/4⌋. Case 4. n ≡ 3 (mod 4), that is n = 4ℓ+ 3. 482 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 In this case, y1 = (g1, 3) and x1 = (g1, 1), and y4ℓ+3 = x4ℓ+3 = (g4ℓ+3, 1). In particular, property (P1) also holds for i = 1, and so x1 ∈ D. However, x4ℓ+3 /∈ D. The set D ∪ {x4ℓ+3} is therefore a dominating set of G ⊗f H , and so γ(G ⊗f H) ≤ |D| + 1 = kn+ 1 = kn+ ⌈n/4⌉ − ⌊n/4⌋. In all four cases, γ(G⊗f H) ≤ kn+ ⌈n/4⌉ − ⌊n/4⌋. 3.2 The cycle Cn and cycles C3k+2 In this section, we determine the Sierpiński domination number γS(Cn, C3k+2) and the upper Sierpiński domination number ΓS(Cn, C3k+2). Theorem 3.11. For n ≥ 3 and k ≥ 1, we have ΓS(Cn, C3k+2) = (k + 1)n. Proof. For n ≥ 3 and k ≥ 1, let G ∼= Cn and let H ∼= C3k+2. Let f : V (G) → V (H) be a constant function, that is, we select h ∈ V (H) and for every vertex g ∈ V (G), we set f(g) = h. For each vertex g ∈ V (G), let Hg denote the copy of H associated with the vertex g. Let D be a dominating set of G ⊗f H , and let Dg = D ∩ V (Hg), and so Dg is the restriction of D to the copy Hg of H . If the vertex (g, h) does not belong to Dg , then Dg dominates all vertices on the path Hg − (g, h) ∼= P3k+1, and so |Dg| ≥ γ(P3k+1) = k + 1. If the vertex (g, h) does belong to Dg , then Dg dominates all vertices on the cycle Hg ∼= C3k+2, and so |Dg| ≥ γ(C3k+2) = k + 1. In both cases, |Dg| ≥ k + 1. Therefore, γ(G⊗f H) = |D| = ∑ g∈V (G) |Dg| ≥ (k + 1)n, implying that ΓS(Cn, C3k+2) ≥ (k + 1)n. By the upper bound in Theorem 2.1, we have ΓS(G,H) ≤ n(G)γ(H) = (k + 1)n, noting that in this case γ(H) = γ(C3k+2) = k + 1. Consequently, ΓS(Cn, C3k+2) = (k + 1)n. Theorem 3.12. For n ≥ 3 and k ≥ 1, γS(Cn, C3k+2) ∈ {kn+ ⌊n 2 ⌋ , kn+ ⌊n 2 ⌋ + 1}. Moreover, if n ≡ 0 mod 4, then γS(Cn, C3k+2) = kn+ ⌊n 2 ⌋ . Proof. For n ≥ 3 and k ≥ 1, let G ∼= Cn and let H ∼= C3k+2. We adopt our notation employed in our earlier proofs. Thus, the cycle G is given by g1g2 . . . gng1, and V (H) = {1, 2, . . . , 3k + 2} where vertices i and i + 1 are consecutive on the cycle H for all i ∈ [3k + 2] (and where addition is taken modulo 3k + 2, and so vertex 1 and vertex 3k + 2 are adjacent). As before, we denote the copy giH of H corresponding to the vertex gi simply by Hi for each i ∈ [n]. Thus, Hi = C3k+2 is the cycle (gi, 1), (gi, 2), . . . , (gi, 3k+ 2), (gi, 1) for all i ∈ [n]. We adopt our notation from the proof of Theorem 3.6. Thus, we denote the connecting edge from Hi to Hi+1 by xiyi+1 for all i ∈ [n], where xi ∈ V (Hi), yi+1 ∈ V (Hi+1), and addition is taken modulo n. Thus, yi = (gi, f(gi−1)) and xi = (gi, f(gi+1)) for all i ∈ [n]. We proceed further with two claims. The first claim establishes a lower bound on γS(G,H), and the second claim an upper bound on γS(G,H). Combining these two bounds yields the desired result in the statement of the theorem. M. A. Henning et al.: The Sierpiński domination number 483 Claim 3.13. γS(Cn, H) ≥ kn+ ⌊n 2 ⌋ . Proof. Let f : V (G) → V (H) be an arbitrary function. We show that γ(G⊗f H) ≥ kn+ ⌊n 2 ⌋ . (3.4) Let D be a γ-set of G ⊗f H constructed, and let Di = D ∩ V (Hi) for i ∈ [n]. If the vertex xi is not dominated by Di, then either xi ̸= yi, in which case xi is dominated by the vertex yi+1 ∈ D, or xi = yi, in which case xi is dominated by the vertex xi−1 ∈ D or the vertex yi+1 ∈ D. Analogously, if the vertex yi is not dominated by Di, then either xi ̸= yi, in which case yi is dominated by the vertex xi−1 ∈ D, or xi = yi, in which case yi is dominated by the vertex xi−1 ∈ D or the vertex yi+1 ∈ D. If a vertex is not dominated by Di, then such a vertex is xi or yi, and we say that such a vertex is dominated from outside Hi. Similarly as before, we proceed with a claim that delivers properties of sets Di leading to the desired lower bound on the Sierpiński domination number. Claim 3.14. The following properties hold in the graph Hi. (a) If d(xi, yi) ≡ 1 (mod 3), then |Di| = k. Further, both xi and yi are dominated from outside Hi. (b) If d(xi, yi) ̸≡ 1 (mod 3), then |Di| = k + 1. Proof. Suppose that Di contains a vertex wi that dominates xi. Possibly, wi = xi. In order to dominate the 3(k−1)+2 vertices in Hi not dominated by wi, at least k additional vertices are needed even if the vertex yi is dominated outside the cycle Hi. Thus in this case, |Di| ≥ k+1, implying by the minimality of the set D that |Di| = k+1. Analogously, if Di contains a vertex that dominates yi, then |Di| = k + 1. Hence, if xi or yi (or both xi and yi) are dominated by Di, then |Di| = k + 1. Suppose that neither xi nor yi is dominated by Di, implying that both xi and yi are dominated from outside the cycle Hi. Thus, Di is a dominating set of H ′i = Hi−xi−yi. If xi = yi, then H ′i = P3k+1, and by the minimality of D we have |Di| = γ(P3k+1) = k+1. Hence, we may assume that xi ̸= yi. If xi and yi are adjacent, then H ′i = P3k, and by the minimality of D we have |Di| = γ(P3k) = k. Suppose that xi and yi are not adjacent, and so H ′ is the disjoint union of two paths Pk1 and Pk2 , where k1 + k2 = 3k. If k1 = 3j1 +1 and k2 = 3j2 + 2 (or if k1 = 3j1 + 2 and k2 = 3j2 + 1) for some integers ji and j2 where j1+ j2 = k− 1, then |Di| = ⌈k1/3⌉+ ⌈k2/3⌉ = (j1+1)+ (j2+1) = k+1. If k1 = 3j1 and k2 = 3j2 where j1 + j2 = k, then |Di| = ⌈k1/3⌉ + ⌈k2/3⌉ = j1 + j2 = k. Hence if neither xi nor yi is dominated by Di, then either d(xi, yi) ≡ 1 (mod 3), in which case |Di| = k, or d(xi, yi) ̸≡ 1 (mod 3), in which case |Di| = k+1. This proves properties (a) and (b) of the claim. (2) By Claim 3.14, if |Di| = k for some i ∈ [n], then |Di−1| = |Di+1| = k + 1 where addition is taken modulo n. Furthermore in this case when |Di| = k, the vertices xi and yi are distinct and are both dominated from outside Hi, implying that yi+1 ∈ Di+1 and xi−1 ∈ Di−1. This implies that if n is even, then |D| ≥ kn + n/2, and if n is odd, then |D| ≥ kn+ (n+ 1)/2. This proves inequality (3.4). Claim 3.15. γS(Cn, H) ≤ kn+ ⌊n 2 ⌋ + ⌈n 4 ⌉ − ⌊n 4 ⌋ . 484 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 Proof. Let f : V (G) → V (H) be the function defined by f(gi) =  1; i ≡ 1 (mod 4), 2; i ≡ 2 (mod 4), 3; otherwise. for all i ∈ [n] where addition is taken modulo n. Adopting our earlier notation, recall that yi = (gi, f(gi−1)) and xi = (gi, f(gi+1)) for all i ∈ [n]. Let n = 4ℓ+ j where j ∈ [3]0 = {0, 1, 2, 3}. We note that f(g4i−3) = 1, f(g4i−2) = 2, and f(g4i−1) = f(g4i) = 3 for all i ∈ [ℓ]. Case 1. n ≡ 0 (mod 4). Thus, n = 4ℓ. We note that y4i−3 = (g4i−3, 3) and x4i−3 = (g4i−3, 2) for all i ∈ [n], and so in the graph H4i−3 the vertices x4i−3 and y4i−3 are at distance 1. Moreover, y4i−1 = (g4i−1, 2) and x4i−1 = (g4i−1, 3) for all i ∈ [n], and so in the graph H4i−1 the vertices x4i−1 and y4i−1 are at distance 1. This implies that H4i−j − {x4i−j , y4i−j} ∼= C3k for j ∈ {1, 3}. Let D4i−j be a γ-set of H4i−j − {x4i−j , y4i−j} for j ∈ {1, 3}, and so |D4i−j | = k. We also note that y4i−2 = (g4i−2, 1) and x4i−2 = (g4i−2, 3) for all i ∈ [n], and so in the graph H4i−2 the vertices x4i−2 and y4i−2 are at distance 2. Moreover, y4i = (g4i, 3) and x4i = (g4i, 1) for all i ∈ [n], and so in the graph H4i the vertices x4i and y4i are at distance 2. This implies that H4i−j − N [{x4i−j , y4i−j}] ∼= C3(k−1) for j ∈ {0, 2}. Let D4i−j be a γ-set of H4i−j that contains both vertices x4i−j and y4i−j for j ∈ {0, 2}, and so |D4i−j | = k + 1. The set D = 4ℓ⋃ i=1 Di is a dominating set of G⊗f H , and so γ(G⊗f H) ≤ |D| = 4kℓ+ 2ℓ = kn+ n/2. Case 2. n ≡ 2 (mod 4). Thus, n = 4ℓ + 2 and in this case, f(g4ℓ+1) = 1 and f(g4ℓ+2) = 2. We note that in the graph H4ℓ+1, the vertices x4ℓ+1 and y4ℓ+1 are at distance 1 and in the graph H4ℓ+2 we have x4ℓ+2 = y4ℓ+2. For i ∈ [4ℓ], we define the set Di exactly as in the previous case. Further, let D4ℓ+1 be a γ-set of H4ℓ+1 − {x4ℓ+1, y4ℓ+1} ∼= C3k, and let D4ℓ+2 be a γ-set of H4ℓ+2 containing x4ℓ+2. We note that |D4ℓ+1| = k and |D4ℓ+2| = k + 1. The set D = 4ℓ+2⋃ i=1 Di is a dominating set of G⊗f H , and so γ(G⊗f H) ≤ |D| = 4kℓ+2k+2ℓ+1 = kn+n/2. Case 3. n ≡ 1 (mod 4). Thus, n = 4ℓ+ 1, and in this case, f(g4ℓ+1) = 1. Thus, y4ℓ+1 = (g4ℓ+1, 3) and x4ℓ+1 = (g4ℓ+1, 1), and so in the graph H4ℓ+1, the vertices x4ℓ+1 and y4ℓ+1 are at distance 2. For i ∈ [4ℓ], we define the set Di exactly as in the previous cases. Further, let D4ℓ+1 be a γ-set of H4ℓ+1 that contains the vertex x4ℓ+1. We note that |D4ℓ+1| = k + 1. The set D = 4ℓ+1⋃ i=1 Di M. A. Henning et al.: The Sierpiński domination number 485 is a dominating set of G⊗fH , and so γ(G⊗fH) ≤ |D| = 4kℓ+k+2ℓ+1 = kn+(n+1)/2. Case 4. n ≡ 3 (mod 4). Thus, n = 4ℓ + 3, and in this case, f(g4ℓ+1) = 1, f(g4ℓ+2) = 2, and f(g4ℓ+3) = 3. In particular, y4ℓ+3 = (g4ℓ+3, 2) and x4ℓ+3 = (g4ℓ+3, 1), and so in the graph H4ℓ+3, the vertices x4ℓ+3 and y4ℓ+3 are at distance 1. For i ∈ [4ℓ + 2], we define the set Di exactly as in Case 2. Further, let D4ℓ+3 be a γ-set of H4ℓ+3 containing the vertex x4ℓ+3. We note that |D4ℓ+3| = k + 1. The set D = 4ℓ+3⋃ i=1 Di is a dominating set of G ⊗f H , and so γ(G ⊗f H) ≤ |D| = 4kℓ + 3k + 2ℓ + 2 = kn+ (n+ 1)/2. The desired result of the claim now follows from the four cases above. (2) The proof of Theorem 3.12 follows as an immediate consequence of Claim 3.13 and Claim 3.15. 3.3 The cycle Cn and cycles C3k In this section, we determine the Sierpiński domination number γS(Cn, C3k) and the upper Sierpiński domination number ΓS(Cn, C3k). Theorem 3.16. For n ≥ 3 and k ≥ 1, γS(Cn, C3k) = ΓS(Cn, C3k) = kn. Proof. We adopt our notation from the earlier sections. Let G ∼= Cn be the cycle g1g2 . . . gng1, and let Hi be the ith copy of C3k corresponding to the vertices gi of G for i ∈ [n]. As before, we denote the connecting edge from Hi to Hi+1 by xiyi+1 for all i ∈ [n]. Let f : V (G) → V (H) be an arbitrary function. Let D be a γ-set of G ⊗f H , and let Di = D∩V (Hi) for i ∈ [n]. We show that |Di| = k for all i ∈ [n]. If both vertices xi and yi are dominated by Di, then Di is a γ-set of Hi ∼= C3k, and so |Di| = k. If exactly one of xi and yi is dominated by Di, say xi, then by the minimality of the set D, the set Di is a γ-set of Hi − yi ∼= P3k−1, and so |Di| = k. Hence, we may assume that neither xi nor yi is dominated by Di, for otherwise, |Di| = k and the desired bound follows. With our assumption that neither xi nor yi is dominated by Di, the set Di is a γ-set of H ′i = Hi − {xi, yi}. If xi = yi, then H ′i = P3k−1, and by the minimality of D we have |Di| = γ(P3k−1) = k. Hence, we may assume that xi ̸= yi. If xi and yi are adjacent, then H ′i = P3k−2, and by the minimality of D we have |Di| = γ(P3k−2) = k. Suppose that xi and yi are not adjacent, and so H ′ is the disjoint union of two paths Pk1 and Pk2 , where k1 + k2 = 3k − 2. If k1 = 3j1 + 1 and k2 = 3j2 for some integers ji and j2 where j1+j2 = k−1, then |Di| = ⌈k1/3⌉+⌈k2/3⌉ = (j1+1)+j2 = k. Analogously, if k1 = 3j1 and k2 = 3j2 + 1, then |Di| = k. If k1 = 3j1 + 2 and k2 = 3j2 + 2 for some integers ji and j2 where j1 + j2 = k − 2, then |Di| = ⌈k1/3⌉+ ⌈k2/3⌉ = (j1 + 1) + (j2 + 1) = k. In all cases, |Di| = k, implying that γ(G⊗f H) = |D| = n∑ i=1 |Di| = kn. Since f : V (G) → V (H) was chosen as an arbitrary function, and D as an arbitrary γ-set of G⊗f H , we deduce that γS(Cn, C3k) = ΓS(Cn, C3k) = γ(G⊗f H) = kn. 486 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 4 Concluding remarks It seems to us that in the vast majority of cases where the lower Sierpiński domination num- ber of the Sierpiński product of two cycles is specified to two values exactly, the larger of the two is the correct value. However, the following example, which surprised us, demon- strates that there are also cases where the exact value is the smaller of the two possible values. Let G ∼= C18 with V (G) = [18] and let H ∼= C7 with V (H) = [7] and let the function f : V (G) → V (H) be defined as follows: f(1) = f(4) = f(5) = f(18) = 4, f(2) = f(3) = f(6) = f(7) = 2, f(8) = f(9) = 7, f(10) = f(11) = 5, f(12) = f(13) = 3, f(14) = f(15) = 1, f(16) = f(17) = 6. Then Theorem 3.2 asserts that γ(G⊗f H) ∈ {36, 37} and it is straightforward to check that the exact value is γ(G⊗f H) = 36. ORCID iDs Michael A. Henning https://orcid.org/0000-0001-8185-067X Sandi Klavžar https://orcid.org/0000-0002-1556-4744 Elżbieta Kleszcz https://orcid.org/0000-0002-1413-2413 Monika Pilśniak https://orcid.org/0000-0002-3734-7230 References [1] K. Balakrishnan, M. Changat, A. M. Hinz and D. S. Lekha, The median of Sierpiński graphs, Discrete Appl. Math. 319 (2022), 159–170, doi:10.1016/j.dam.2021.02.005, https://doi. org/10.1016/j.dam.2021.02.005. [2] B. Brešar and J. Ferme, Packing coloring of Sierpiński-type graphs, Aequationes Math. 92 (2018), 1091–1118, doi:10.1007/s00010-018-0561-8, https://doi.org/10.1007/ s00010-018-0561-8. [3] F. Deng, Z. Shao and A. Vesel, On the packing coloring of base-3 Sierpiński graphs and H- graphs, Aequationes Math. 95 (2021), 329–341, doi:10.1007/s00010-020-00747-w, https: //doi.org/10.1007/s00010-020-00747-w. [4] A. Estrada-Moreno, E. D. Rodrı́guez-Bazan and J. A. Rodrı́guez-Velázquez, On distances in generalized Sierpiński graphs, Appl. Anal. Discrete Math. 12 (2018), 49–69, doi:10.2298/ AADM160802001E, https://doi.org/10.2298/AADM160802001E. [5] A. Estrada-Moreno and J. A. Rodrı́guez-Velázquez, On the General Randić index of polymeric networks modelled by generalized Sierpiński graphs, Discrete Appl. Math. 263 (2019), 140–151, doi:10.1016/j.dam.2018.03.032, https://doi.org/10.1016/j. dam.2018.03.032. M. A. Henning et al.: The Sierpiński domination number 487 [6] M. Farrokhi Derakhshandeh Ghouchan, E. Ghorbani, H. R. Maimani and F. R. Mahid, Some algebraic properties of Sierpiński-type graphs, Ars Math. Contemp. 20 (2021), 171–186, doi:10. 26493/1855-3974.2199.97e, https://doi.org/10.26493/1855-3974.2199.97e. [7] S. Gravier, M. Kovše and A. Parreau, Generalized sierpiński graphs, in: Posters at Eu- roComb’11, Budapest, 2011, (2017–12–11), http://www.renyi.hu/conferences/ ec11/posters/parreau.pdf. [8] T. W. Haynes, S. T. Hedetniemi and M. A. Henning (eds.), Topics in Domination in Graphs, volume 64 of Dev. Math., Springer, Cham, 2020, doi:10.1007/978-3-030-51117-3, https: //doi.org/10.1007/978-3-030-51117-3. [9] T. W. Haynes, S. T. Hedetniemi and M. A. Henning (eds.), Structures of Domination in Graphs, volume 66 of Dev. Math., Springer, Cham, 2021, doi:10.1007/978-3-030-58892-2, https: //doi.org/10.1007/978-3-030-58892-2. [10] M. A. Henning and A. Yeo, Total Domination in Graphs, Springer Monographs in Mathemat- ics, Springer, New York, 2013, doi:10.1007/978-1-4614-6525-6, https://doi.org/10. 1007/978-1-4614-6525-6. [11] A. M. Hinz, S. Klavžar and C. Petr, The Tower of Hanoi—Myths and Maths, Birkhäuser/Springer, Cham, 2nd edition, 2018, doi:10.1007/978-3-319-73779-9, https: //doi.org/10.1007/978-3-319-73779-9. [12] A. M. Hinz, S. Klavžar and S. S. Zemljič, A survey and classification of Sierpiński-type graphs, Discrete Appl. Math. 217 (2017), 565–600, doi:10.1016/j.dam.2016.09.024, https://doi. org/10.1016/j.dam.2016.09.024. [13] W. Imrich and I. Peterin, Recognizing generalized Sierpiński graphs, Appl. Anal. Discrete Math. 14 (2020), 122–137, doi:10.2298/aadm180331003i, https://doi.org/10.2298/ aadm180331003i. [14] M. Khatibi, A. Behtoei and F. Attarzadeh, Degree sequence of the generalized Sierpiński graph, Contrib. Discrete Math. 15 (2020), 88–97, doi:10.11575/cdm.v15i3.68174, https://doi. org/10.11575/cdm.v15i3.68174. [15] S. Klavžar and U. Milutinović, Graphs S(n, k) and a variant of the Tower of Hanoi problem, Czechoslovak Math. J. 47(122) (1997), 95–104, doi:10.1023/A:1022444205860, https:// doi.org/10.1023/A:1022444205860. [16] S. Klavžar and S. S. Zemljič, Connectivity and some other properties of generalized Sierpiński graphs, Appl. Anal. Discrete Math. 12 (2018), 401–412, doi:10.2298/AADM170206009K, https://doi.org/10.2298/AADM170206009K. [17] D. Korže and A. Vesel, Packing coloring of generalized Sierpiński graphs, Discrete Math. Theor. Comput. Sci. 21 (2019), Paper No. 7, 18 pp., doi:10.23638/DMTCS-21-3-7, https: //doi.org/10.23638/DMTCS-21-3-7. [18] J. Kovič, T. Pisanski, S. S. Zemljič and A. Žitnik, The Sierpiński product of graphs, Ars Math. Contemp. 23 (2023), Paper No. 1.01, 25 pp., doi:10.26493/1855-3974.1970.29e, https:// doi.org/10.26493/1855-3974.1970.29e. [19] C.-A. Liu, Roman domination and double Roman domination numbers of Sierpiński graphs S(Kn, t), Bull. Malays. Math. Sci. Soc. 44 (2021), 4043–4058, doi:10.1007/ s40840-021-01136-5, https://doi.org/10.1007/s40840-021-01136-5. [20] M. K. Menon, M. R. Chithra and K. S. Savitha, Security in Sierpiński graphs, Discrete Appl. Math. 328 (2023), 10–15, doi:10.1016/j.dam.2022.11.017, https://doi.org/10. 1016/j.dam.2022.11.017. 488 Ars Math. Contemp. 24 (2024) #P3.06 / 473–488 [21] F. Ramezani, E. D. Rodrı́guez-Bazan and J. A. Rodrı́guez-Velázquez, On the Roman domi- nation number of generalized Sierpiński graphs, Filomat 31 (2017), 6515–6528, doi:10.2298/ fil1720515r, https://doi.org/10.2298/fil1720515r. [22] J. A. Rodrı́guez-Velázquez, E. D. Rodrı́guez-Bazan and A. Estrada-Moreno, On generalized Sierpiński graphs, Discuss. Math. Graph Theory 37 (2017), 547–560, doi:10.7151/dmgt.1945, https://doi.org/10.7151/dmgt.1945. [23] J. Varghese and S. Aparna Lakshmanan, Italian domination on Mycielskian and Sierpinski graphs, Discrete Math. Algorithms Appl. 13 (2021), Paper No. 2150037, 9 pp., doi:10.1142/ S1793830921500373, https://doi.org/10.1142/S1793830921500373. [24] E. Vatandoost, F. Ramezani and S. Alikhani, On the zero forcing number of general- ized Sierpiński graphs, Trans. Comb. 8 (2019), 41–50, doi:10.22108/toc.2018.101107.1463, https://doi.org/10.22108/toc.2018.101107.1463. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.07 / 489–507 https://doi.org/10.26493/1855-3974.2835.8f0 (Also available at http://amc-journal.eu) On the minisymposium problem* Peter Danziger † Department of Mathematics, Toronto Metropolitan University, Toronto, ON M5B 2K3, Canada Eric Mendelsohn Department of Mathematics, University of Toronto, Toronto, ON M5S 2E4, Canada Brett Stevens ‡ School of Mathematics and Statistics - Carleton University, Ottawa, ON K1S 5B6, Canada Tommaso Traetta § DICATAM, Università degli Studi di Brescia, Via Branze 43, 25123 Brescia, Italy Received 26 February 2022, accepted 25 July 2023, published online 29 July 2024 Abstract The generalized Oberwolfach problem asks for a factorization of the complete graph Kv into prescribed 2-factors and at most a 1-factor. When all 2-factors are pairwise isomor- phic and v is odd, we have the classic Oberwolfach problem, which was originally stated as a seating problem: given v attendees at a conference with t circular tables such that the ith table seats ai people and ∑t i=1 ai = v, find a seating arrangement over the v−1 2 days of the conference, so that every person sits next to each other person exactly once. In this paper we introduce the related minisymposium problem, which requires a solu- tion to the generalized Oberwolfach problem on v vertices that contains a subsystem on m vertices. That is, the decomposition restricted to the required m vertices is a solution to the generalized Oberwolfach problem on m vertices. In the seating context above, the larger conference contains a minisymposium of m participants, and we also require that pairs of these m participants be seated next to each other for ⌊ m−1 2 ⌋ of the days. When the cycles are as long as possible, i.e. v, m and v−m, a flexible method of Hilton and Johnson provides a solution. We use this result to provide further solutions when v ≡ m ≡ 2 (mod 4) and all cycle lengths are even. In addition, we provide extensive *We thank the anonymous referees for their many useful suggestions that helped strongly improve this paper. †P. Danziger has received support from NSERC Discovery Grants RGPIN-2022-03816. ‡B. Stevens recieved support from NSERC Discovery Grant RGPIN-2017-06392. §Corresponding author. T. Traetta has received support from GNSAGA of Istituto Nazionale di Alta Matem- atica. cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 490 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 results in the case where all cycle lengths are equal to k, solving all cases when m | v, except possibly when k is odd and v is even. Keywords: Minisymposium problem, (generalized) Oberwolfach problem, 2-factorizations, subsys- tems. Math. Subj. Class. (2020): 05C51, 05B30 1 Introduction We assume that the reader is familiar with the fundamentals of graph theory and of design theory and refer them to [41] and [14], respectively. In particular, a factor is a spanning subgraph and an r-factor is a factor which is r-regular, so in a 1-factor every vertex has degree one and a 2-factor is a disjoint union of cycles. Given a collection of factors, F , an F-factorization of a graph G is a decomposition of the edges of G into subgraphs, each of which is isomorphic to some F ∈ F . If F = {F} we speak of an F -factorization. We use Kn to denote the complete graph on n vertices and K∗n to denote the graph Kn when n is odd and Kn − I , where I is a 1-factor, when n is even. Similarly, Km,n denotes the complete bipartite graph with parts of sizes m and n. If the parts are X and Y , respectively, we may also speak of KX,Y . A 2-factor is called uniform if all of its con- stituent cycles are of the same length; it is called Hamiltonian* if its cycles have longest possible lengths given the requirements of the factorization. If a 2-factor, F , consists en- tirely of cycles of a particular length, k say, we refer to an F -factor and an F -factorization as a Ck-factor and Ck-factorization, respectively. Given a graph G, we denote by G[n] the lexicographic product of G with the empty graph on n vertices. Specifically, the vertex set of G[n] is V (G) × Zn (where Zn denotes the cyclic group of order n) and (x, i)(y, j) ∈ E(G[n]) if and only if xy ∈ E(G), i, j ∈ Zn. The well known Oberwolfach problem asks for a 2-factorization of K∗n into 2-factors all of which are isomorphic to a given 2-factor F . A summary of results up until 2006 can be found in [14, Section VI.12], in particular the case of uniform factors has been solved [1, 2, 26]. Theorem 1.1 ([1, 2, 26]). Given integers v, k ≥ 3, there is a Ck-factorization of K∗v if and only if k | v, except that there is no C3-factorization of K∗6 or K∗12. The case when all cycles in F have even length has been completely solved in [6]. The case with exactly two cycles is solved in [38]. The case of the complete graph Kℵ, where ℵ is any infinite cardinal has been completely solved in [15]. In the related Hamilton- Waterloo problem two 2-factors F1 and F2 are specified and we are asked for a factorization of K∗n into a given number of each of the factors. There has been much recent progress in this problem, see [3, 5, 7, 10, 11, 12, 17, 27, 28, 29, 32, 39, 40]. More generally, in the generalized Oberwolfach problem we are given a set of 2-factors F1, F2, . . . , Ft of Kn and positive integers α1, α2, . . . , αt, where ∑ αi = ⌊n−12 ⌋, and are asked for a factorization of K∗n which contains αi copies of the 2-factor Fi, see [6, 13, 21]. A major recent development gives a non-constructive asymptotic existence result for the generalized Oberwolfach problem [23]. E-mail addresses: danziger@torontomu.ca (Peter Danziger), mendelso@math.utoronto.ca (Eric Mendelsohn), brett@math.carleton.ca (Brett Stevens), tommaso.traetta@unibs.it (Tommaso Traetta) P. Danziger et al.: On the minisymposium problem 491 Other graphs have also been considered. In particular, Liu has shown the following for the complete multipartite graph. Theorem 1.2 ([30, 31]). Let k, t and u be positive integers with k ≥ 3. There exists a Ck-factorization of Kt[u] if and only if k | tu, (t − 1)u is even, further k is even if t = 2, and (k, t, u) ̸∈ {(3, 3, 2), (3, 6, 2), (3, 3, 6), (6, 2, 6)}. Originally the Oberwolfach problem was stated as a seating problem: Given an odd number v of attendees at a conference with t circular tables such that the ith table seats ai people and ∑t i=1 ai = v, find a seating arrangement over the v−12 days of the conference, so that every person sits next to each other person exactly once. In this paper we introduce the related minisymposium problem. In this case we require a solution to the generalized Oberwolfach problem on v vertices such that its restriction to a subset of m vertices constitutes a solution to the generalized Oberwolfach problem on m vertices. Another way of considering the problem asks for a solution to the generalized Oberwolfach problem on v vertices which contains a subsystem on m vertices. In the seating context above, the larger conference contains a minisymposium of m participants, and we also require that pairs of these m participants be seated next to each other for⌊ m−1 2 ⌋ of the days. A similar problem has been considered, for example, in [9] for whist tournaments. Section 2 gives the formal definition of a minisymposium factorization and some nec- essary conditions for its existence, as well as introduces some special cases. In Section 3 we show how to use a flexible theorem by Hilton and Johnson [25] to solve the case of Hamiltonian* 2-factors (where the cycles are as long as possible). The same section con- siders the case where all cycles are of even length and v ≡ m ≡ 2 (mod 4). Section 4 considers the uniform case, where all cycles have the same length. We completely solve the case where all cycles are of length m when (v − 1)m is even. In Section 5 we discuss and give some preliminary results on factorizations that contain more than one subsystem. We provide some concluding remarks in the final section. 2 Preliminaries We begin by giving a formal definition of a minisymposium factorization. The minisym- posium problem is equivalent to the original Oberwolfach problem when v = m. Hence we will generally assume that v > m. Definition 2.1. Given positive integers v and m with v ≥ m, let F = { Fi : i = 1, . . . , ⌊ v − 1 2 ⌋ − ⌊ m− 1 2 ⌋} , be a collection of 2-factors on v vertices and let G = { (Ti, Ui) : i = 1, . . . , ⌊ m− 1 2 ⌋} , where the Ti are 2-factors on m vertices and the Ui are 2-factors on v − m vertices. We define a minisymposium factorization MSF(F ,G) as a factorization of K∗v into 2-factors F ∈ F and Gi = Ti ∪Ui, where (Ti, Ui) ∈ G, such that T = {Ti : i = 1, . . . , ⌊ m−1 2 ⌋ } is a factorization of a subgraph of K∗v isomorphic to K ∗ m. 492 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 Note that with the notation MSF(F ,G), we assume that the parameters v, m, T and U = {Ui : i = 1, . . . , ⌊ m−1 2 ⌋ } are defined implicitly. We may also use the notation MSF(F , (T ,U)), if we wish to explicitly refer to the factorizations T and U . An MSF(F ,G) can be thought of as a 2-factorization of K∗v with a subsystem of size m. When m = 1 or 2, this is just a factorization of K∗v into 2-factors in F , which is equivalent to a solution of the generalized Oberwolfach problem, and so we will assume m ≥ 3. Similarly, when m = v, this is a factorization into the Ti, so we assume that m < v. Removing the subsystem, we can talk about a 2-factorization of K∗v with a “hole” of size m. However, care must be taken when either v or m is even as the placement of the various 1-factors must be considered, as noted below. We note that the size of F is⌊ v − 1 2 ⌋ − ⌊ m− 1 2 ⌋ =  v−m 2 v ≡ m (mod 2) v−m+1 2 v + 1 ≡ m ≡ 0 (mod 2) v−m−1 2 v ≡ m+ 1 ≡ 0 (mod 2). • In the case when both v and m are even, we are considering a factorization of K∗v = Kv − I , where I is a 1-factor, containing a factorization of a subgraph G− J of K∗v where G ∼= Km and J is a 1-factor of G contained in I . • When v is even and m is odd, we are considering a factorization of Kv − I , where I is a 1-factor, containing a factorization of a subgraph G ∼= Km. Note that none of the edges of I are contained in G. • When v is odd and m is even, we are considering a factorization of K∗v = Kv containing a factorization of G − J , where G ∼= Km and J is a 1-factor of G. We note that the edges of J are not covered by factors in T , and hence must be covered by factors in F . • When both v and m are odd there is no 1-factor to consider. Since in all cases G ∼= Km, we will henceforth refer to it as the Km. We note that none of the edges of the Km are covered by F , except in the case of m even and v odd; in this case, the edges of the 1-factor J of the Km are covered by F . We use this observation in the proofs of the lemmas below, where for a given 2-factor F , we define ak(F ) to be the number of cycles of length k in F . Lemma 2.2. For a given v and m, if there is a minisymposium factorization, MSF(F ,G), then for each factor F ∈ F using cF edges in the Km, v −m ≥ −cF + v∑ i=3 ai(F ) ⌈ i 2 ⌉ , (2.1) m ≤ cF + v∑ i=3 ai(F ) ⌊ i 2 ⌋ . (2.2) In particular, if v is even, or m is odd, all of the cF = 0 and therefore,∑v i=3 ai(F ) ⌈ i 2 ⌉∑v i=3 ai(F ) ⌊ i 2 ⌋ ≤ v −m m . (2.3) P. Danziger et al.: On the minisymposium problem 493 Proof. We first deal with the case when v is even, or m is odd. In this case none of the edges of the Km appear in any F ∈ F . Therefore, for each F ∈ F , at most ⌊ i 2 ⌋ vertices of any cycle of length i in F are inside the Km, hence m ≤ v∑ i=3 ai(F ) ⌊ i 2 ⌋ . (2.4) Similarly, for any cycle of length i in F , at least ⌈ i2⌉ vertices of the cycle are not in the Km, thus v∑ i=3 ai(F ) ⌈ i 2 ⌉ ≤ v −m. (2.5) Thus Inequality (2.3) follows. Now, if v is odd and m is even the m2 edges in the 1-factor J of the Km must be used in factors from F . Suppose that F ∈ F uses cF of these edges. Each edge of the Km used can increase the right hand side of Inequality (2.4) by no more than one and decrease the left hand side of Inequality (2.5) by no more than one. Theorem 2.3. For a given v and m, if there is a minisymposium factorization, MSF(F ,G), then v ≥ 2m, unless v is odd and m is even, in which case v ≥ 2m− 1. Proof. When v is even or m is odd, the left hand side of Inequality (2.3) is at least 1 and therefore v ≥ 2m. When v is odd and m is even, ∑ F∈F cF = m 2 . Since v is odd, each F ∈ F must contain at least one odd cycle, therefore ∑v i=3 ai(F )⌈ i 2⌉ ≥ ⌈ v 2⌉ = v+1 2 . Also note that the number of factors F ∈ F is v−m+12 . Summing Inequality (2.1) over all of the F ∈ F twice gives (v −m+ 1)(v −m) ≥ 2 ∑ F∈F ( −cF + v∑ i=3 ai(F ) ⌈ i 2 ⌉) = −m+ 2 ∑ F∈F v∑ i=3 ai(F ) ⌈ i 2 ⌉ ≥ −m+ 2 ∑ F∈F v + 1 2 = (v −m+ 1)v + 1 2 −m Hence, 0 ≤ (v −m+ 1)(v −m)− (v −m+ 1)v + 1 2 +m = (v −m+ 1)(v − 2m− 1) + 2m 2 = (v − (2m− 1))(v − (m+ 1)) 2 . 494 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 When v = m + 1, the factors in U are required to be 2-regular graphs on a single vertex, which is not possible, so v − (m + 1) > 0. Thus v − (2m − 1) ≥ 0 and the result follows. In the case where v is odd and m is even and v = 2m − 1, we have that U is a factorization of Kv−m. So we can interchange both the roles of m and v −m, as well as those of T and U . Thus, without loss of generality, we may assume that v ≥ 2m in all cases. There are two cases of initial special interest. Firstly, the case of uniform cycle lengths (when all cycles in a factor are of the same length), which we consider in detail in Section 4. Secondly, the case where the cycles are as long as possible, which in correspondence with the definition of K∗n and the Hamiltonian-like nature of such factorizations we will call Hamiltonian* factorizations. We formally define Hamiltonian* factorizations in Section 3. There we show that a method of Hilton and Johnson completely settles their existence. An MSF(F ,G) in which all of the factors in F , T and U are uniform with the same cycle length k is called uniform and we refer to it as a UMSF(v,m, k). In this case we have the following necessary conditions. Theorem 2.4. If v > m and a UMSF(v,m, k) exists, then k ≥ 3, k | m and k | v. Furthermore, • if k is even, then v ≥ 2m; • if k is odd, then v ≥ 2mkk−1 . Proof. Since we are forming 2-factors with cycles of length k, we require k ≥ 3. The divisibility conditions follow directly from the requirement for a factorization of K∗v and K∗m into k-cycles. If k is even, then v is even, since it is a multiple of k, and Theorem 2.3 gives v ≥ 2m. If k is odd, we note that for a Ck-factor F ∈ F , we have ai(F ) = vk when i = k and 0 otherwise, ⌊k2 ⌋ = k−1 2 and ⌈ k 2 ⌉ = k+1 2 . Thus, when v is even or m is odd, Inequality (2.3) implies that v ≥ 2mkk−1 and the result follows. This leaves the case when v and k are odd and m is even. We sum Inequality (2.1) over all the F ∈ F to obtain∑ F∈F (v −m) ≥ ∑ F∈F ( −cF + ∑v i=3 ai(F ) ⌈ i 2 ⌉) 1 2 (v −m+ 1)(v −m) ≥ −m/2 + ∑ F∈F v k k+1 2 2k(v −m+ 1)(v −m) ≥ v(v −m+ 1)(k + 1)− 2mk. Rearranging and expanding in v gives (k − 1)v2 + (−3km+m+ k − 1)v + 2km2 ≥ 0. (2.6) Let f(v) = (k − 1)v2 + (−3km+m+ k − 1)v + 2km2. By Theorem 2.3, v ≥ 2m− 1, but f (2m− 1) = m(1 + k − 2m) < 0, P. Danziger et al.: On the minisymposium problem 495 so v is at least as large as the larger of the two roots of f . Now f ( 2mk k − 1 − 2 ) = −2(m+ 1− k) < 0, and f ( 2mk k − 1 − 1 ) = (k − 1)m > 0. Thus f has its larger root between 2mkk−1 − 2 and 2mk k−1 − 1. It is left to check that v ̸= ⌊ 2mkk−1 −1⌋, ⌈ 2mk k−1 −1⌉. Recalling that k | v, if v = ⌊ 2mk k−1 −1⌋ or v = ⌈ 2mkk−1 − 1⌉, then there exist a rational number 0 ≤ ϵ < 1 and a positive integer a such that v = 2mk k − 1 − 1± ϵ = ak. Multiplying both sides by k−12k and rearranging we have that m− ak − 1 2 = (1± ϵ)k − 1 2k . Since k is odd, the left side is an integer. However, 0 < (1 ± ϵ)k−12k < 1, a contradiction. We conclude that v ≥ 2mkk−1 . One interesting case in light of these necessary conditions is when m = k, i.e. a UMSF(v, k, k). For these parameters, since k < v and k | v, we must have v ≥ 2k, so the necessary conditions in Theorem 2.3 are satisfied. We consider these types of fac- torizations in Section 4. 3 Hamiltonian* and bipartite factors Considering non-uniform factors, an obvious case to consider is a Hamiltonian* minisym- posium factorization, which is one in which the cycles have the longest possible lengths. Specifically, the factors in F are all v-cycles, factors in T are all m-cycles and the factors in U are all (v −m)-cycles. Such an MSF(F ,G) is denoted by HMSF(v,m). We sometimes refer to the cycles in T and U as ‘short’ cycles. Because of the lengths of these cycles there are no further necessary conditions beyond those of Theorem 2.3. In a paper on the Oberwolfach problem, Hilton and Johnson prove the following theo- rem on a flexible construction technique. Theorem 3.1 ([25]). Let m and n be integers, 1 ≤ m < n. Let (s1, . . . , st), si ∈ 1, 2, 1 ≤ i ≤ t, be a composition of n−1. Let Km be edge coloured with t colours c1, . . . , ct. Let fi be the number of edges coloured ci and Km(ci) be the ith colour class. This colouring can be extended to an edge-colouring of Kn in which the colour class Kn(ci) is an si-factor, 1 ≤ i ≤ t, and when si = 2, Kn(ci) contains exactly one more cycle than Km(ci) if and only if for all i = 1, 2, . . . , t: fi ≥ si(m− n/2), sin is even, ∆(Km(ci)) ≤ si. This theorem is sufficient to provide a solution to the Hamiltonian* minisymposium factorization. 496 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 Corollary 3.2. An HMSF(v,m) exists if and only if m ≥ 3, v ≥ 2m−1 in case m is even, and v ≥ 2m in case m is odd. Proof. The given conditions are necessary by Theorem 2.3. To prove sufficiency, we will define an edge colouring of the Km from a decomposition of the Km into Hamiltonian cycles and possibly a single 1-factor using Theorem 1.1. If m is odd, this defines edge colours ci, 1 ≤ i ≤ (m− 1)/2. If v is also odd, extend this to a (v − 1)/2-edge colouring of the Km by including (v −m)/2 empty colour classes ci, (m+ 1)/2 ≤ i ≤ (v − 1)/2. Let si = 2 for all 1 ≤ i ≤ (v − 1)/2. If v is even, extend the colouring to a v/2-edge colouring of the Km by adding empty colour classes. Let si = 2 for all 1 ≤ i < v/2 and sv/2 = 1. In both cases, it can be verified that Theorem 3.1 now gives an HMSF(v,m) as desired. If m is even, define m/2 edge colour classes of the Km from a decomposition into Hamiltonian cycles and one 1-factor. Let c1 be the colour class of the 1-factor. If v is also even, extend this to a v/2-edge colouring by adding empty colour classes. Let s1 = 1 and si = 2 for 2 ≤ i ≤ v/2. If v is odd, extend this to a (v − 1)/2-edge colouring by adding empty colour classes. Let si = 2 for 1 ≤ i ≤ v/2. In both cases, it can be verified that Theorem 3.1 now gives an HMSF(v,m) as desired. Theorem 3.1 is more than just an existence result; a recursive procedure can be extracted from the proof to algorithmically build the edge decompositions. We have a more direct construction of all HMSF(v,m) which uses difference methods and decompositions of Cayley graphs [16]. Theorem 3.1 can be used much more generally to build minisymposium factorizations. Essentially it shows that it is possible to extend any 2-factorization of the Km to one of Kv , provided that the necessary conditions hold, where the additional 2-factors are Hamiltonian, with an additional 1-factor when v is even. When all of the cycles of the factors in F , U and T are bipartite (i.e. contain only even cycles), we apply the Theorem of Häggkvist [24] (given below) to HMSF(v,m) to give us a solution to the minisymposium problem when v ≡ m ≡ 2 (mod 4). Theorem 3.3 ([24]). If F is a bipartite 2-regular graph of order 2n, then there is a factor- ization of Cn[2] into 2 isomorphic copies of F . We note that in the case where the factors are bipartite, so all cycle lengths are even, v and m are both even and Theorem 2.3 gives v ≥ 2m. Theorem 3.4. If v ≡ m ≡ 2 (mod 4), and F = {Fi : 1 ≤ i ≤ (v −m)/2}, U = {Uj : 1 ≤ j ≤ (m−2)/2} and T = {Tj : 1 ≤ j ≤ (m−2)/2} are sets of bipartite factors with Fi = Fi+1, Uj = Uj+1 and Tj = Tj+1 for every odd i and j, then an MSF(F , (T ,U)) exists if and only v ≥ 2m. Proof. We note that if v ≡ m ≡ 2 (mod 4), the number of the Fi is (v − m)/2 and the number of the Uj and the Tj is (m− 2)/2, so the number of the Fi, Uj and Tj are all even. We take an HMSF(v/2,m/2), which exists by Corollary 3.2, with factors F ′i of order v/2, U ′i of order (v − m)/2 and T ′i of order m/2. We blow up each vertex by 2 and apply Theorem 3.3 to factor each F ′i [2] into 2 copies of Fi, each U ′ j [2] into 2 copies of Uj and each T ′j [2] into 2 copies of Tj . P. Danziger et al.: On the minisymposium problem 497 If v ≡ 0 (mod 4) or m ≡ 0 (mod 4), then v/2 or m/2 would be even and the HMSF(v/2,m/2) would contain 1-factors either in Kv/2 or the Km/2. When a 1-factor is blown up as done in Theorem 3.3, it results in a C4-factor, which prevents constructing the desired MSF unless F , T , and U already contain this kind of factor. An immediate consequence of Theorem 3.4 is the following relating to uniform factors. Corollary 3.5. If k > 3, k ≡ 2 (mod 4), v ≡ m ≡ k (mod 2k), then a UMSF(v,m, k) exists if and only if v ≥ 2m. 4 Uniform factors In this section we consider the case of uniform factors, i.e. when all cycles are of the same length, k. We recall from Theorem 2.4 that in order for a UMSF(v,m, k) to exist, we require that k ≥ 3, which we will assume throughout this section. We also require k | m and k | v. Additionally, if k is even, then v ≥ 2m and if k is odd, then v ≥ 2mkk−1 . Corollary 3.5 gives a powerful result in the case when k ≡ 2 (mod 4) and v ≡ m ≡ k (mod 2k). The case where k = 3 has been considered in [34, 35, 36] when v and m are both odd, and [18, 19, 20, 22, 37] when they are both even. However, the case when m and v have opposite parities appears to be completely open. We summarize these results in the following theorem. Theorem 4.1 ([20, 35]). If v ≡ m (mod 2), there exists a UMSF(v,m, 3) if and only if v ≥ 3m, v ≡ m ≡ 0 (mod 3), and if v,m are even, then v,m > 12. We will find the following results useful. A corollary of a result in [4] yields the fol- lowing. Theorem 4.2 ([4]). If G is a Hamiltonian decomposable graph, then G[n] is also Hamil- tonian decomposable. In particular, Cm[n] has a Cmn-factorization for every m ≥ 3. Piotrowski [33] has shown the following result for m ≥ 4. The case m = 3 is covered by Theorem 1.2. Theorem 4.3 ([33]). There exists a Cm-factorization of Cm[n], except if n = 2 and m is odd, or when (m,n) = (3, 6). Piotrowski [33] has also shown the following result. Theorem 4.4 ([33]). Let F be a bipartite 2-regular graph of order 2n. The complete bipartite graph K2[n] has an F -factorization if and only if n is even, except when n = 6 and F consists of two 6-cycles. We now give some recursive constructions for uniform minisymposium factorizations. Theorem 4.5. Let m ≥ k ≥ 3 and t ≥ 2 be integers. If (t − 1)m is even and k | m, then there is a UMSF(mt,m, k), except that there is no UMSF(6t, 6, 3) UMSF(12t, 12, 3), UMSF(12, 6, 6), or UMSF(2m,m, k) when k is odd. Proof. The non-existence of a UMSF(6t, 6, 3) and UMSF(12t, 12, 3) are covered by The- orem 4.1. Since a UMSF(2m,m, k) is equivalent to a Ck-factorization of the complete bipartite graph K2[m], it clearly does not exist when the cycle length k is odd, or when k = m = 6 by Theorem 4.4. In all remaining cases, the following conditions simultane- ously hold: 498 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 1. (m, k) ̸∈ {(6, 3), (12, 3)}, 2. (t,m, k) ̸= (2, 6, 6), 3. if k is odd, then t > 2. The assumptions of Theorems 1.1 and 1.2 are then satisfied. Hence there is a Ck-factorization of Kt[m] and a Ck-factorization of K∗m, which we use to fill in the parts of size m in Kt[m]. This completes the proof. Considering the necessary conditions in Theorem 2.4, we get the following corollaries. Corollary 4.6. Suppose that either k is even or v is odd, and m | v. Then there exists a UMSF(v,m, k) if and only if k | m, v ≥ 2m when k is even, and v ≥ 3m when k is odd, except that UMSF(v, 6, 3), UMSF(v, 12, 3) and UMSF(12, 6, 6) do not exist. Proof. Taking v = mt, Theorem 2.4 gives the necessary conditions k | m and v ≥ 2m when k is even. When k is odd, the necessary condition from Theorem 2.4 is v ≥ 2mkk−1 , but since m | v, this implies k ≥ 3m. Given the conditions of k and v, the sufficiency comes from Theorem 4.5. We note that if m | v this corollary completely solves all cases except when k is odd and v is even. One case of particular interest is when k = m, in this case m | v is necessary. Corollary 4.7. Let m(t− 1) be even. Then a UMSF(tm,m,m) exists if and only if t ≥ 2 when m is even, t ≥ 3 when m is odd, and (t,m) ̸= (2, 6). The previous results all require m | v, however the next theorem allows us to recur- sively construct solutions to cases where m does not divide v. Theorem 4.8. Assume there is a UMSF(v,m, k) and let t ≥ 1. Then there exists a UMSF(vtk,mtk, ℓ), with ℓ ∈ {k, kt}, in each of the following cases: (1) v and m have the same parity; (2) v and t are even, ℓ = tk, and m and k are both odd, except possibly when (k, t) = (3, 2). Proof. Letting w ∈ {m, v}, we factorize K∗wtk into Γw = K∗w[tk] and Γw = K∗wtk − Γw. Note that Γw is the vertex disjoint union of (1) w copies of K∗tk when w is odd, or (2) w/2 copies of K∗2tk when w is even. Without loss of generality, we can assume that Γm ⊆ Γv and Γm ⊆ Γv , except when v is odd and m is even. In this case the components of Γm are copies of K∗2tk, while those of Γv are isomorphic to K∗tk, therefore Γm ⊆ Γv cannot hold. We proceed by constructing (a) a Cℓ-factorization of Γv containing a Cℓ-factorization of Γm, and (b) a Cℓ-factorization of Γv containing a Cℓ-factorization of Γm, P. Danziger et al.: On the minisymposium problem 499 which together will provide the desired UMSF (vtk,mtk, ℓ). We blow up each vertex of the UMSF(v,m, k) by tk, to obtain a Ck[tk]-factorization of Γv containing a Ck[tk]-factorization of Γm. To construct (a) it is therefore enough to factorize Ck[tk] into Cℓ-factors, ℓ ∈ {k, tk}. By Theorem 4.3 there is a Ck-factorization of Ck[tk], except when (t, k) = (2, 3). In this case, the desired UMSF(6v, 6m, 3) exists by Theorem 4.1. Considering that Ck[tk] = Ck[t][k], by Theorem 4.2 there exists a Ckt- factorization of Ck[t] which we blow up by k to obtain Ckt[k]-factorization of Ck[tk]. By Theorem 4.3, each Ckt[k]-factor can be further decomposed into Ckt-factors yielding a Ckt-factorization of Ck[tk]. It is left to construct (b). If m and v have the same parity, the components of Γm and Γv are pairwise isomorphic: they are copies of K∗tk or K ∗ 2tk. It is then enough to build a Cℓ-factorization of K∗tk and K ∗ 2tk for ℓ ∈ {k, kt}. They exist by Theorem 1.1 except when ℓ = k = 3 and one of the following two conditions hold, 1. mv is odd and t ∈ {2, 4}, or 2. m and v are even, and t ∈ {1, 2}. In each of these cases, the existence of the desired UMSF (3vt, 3mt, 3) is guaranteed by Theorem 4.1. If v and t are even, ℓ = tk, and both m and k are odd, the components of Γm are isomorphic to K∗tk, while those of Γv are isomorphic to K ∗ 2tk. Since we can factorize K∗2tk into K2[tk] and two copies of K ∗ tk, it is enough to decompose both K2[tk] and K ∗ tk into Ctk-factors. These factorizations exist by Theorem 4.4 and Theorem 1.1, respectively, except possibly when (k, t) = (3, 2). We may now use the result on triples (Theorem 4.1) to obtain the following. Corollary 4.9. Let v ≡ m ≡ 0, 3 (mod 6), with v ≥ 3m and m ̸∈ {0, 6, 12}. Then there exists a UMSF(3tv, 3tm, 3t) for all t > 0. Additionally, we may use Theorem 3.5 to obtain the following result. Corollary 4.10. Let 3 < k, k ≡ 2 (mod 4), v ≡ m ≡ k (mod 2k) and v ≥ 2m. Then there exists a UMSF(vtk,mtk, k) and a UMSF(vtk,mtk, tk) for all t > 0. We note that the above result can be used to obtain UMSF’s with cycle length, subsys- tem size or number of vertices congruent to 0 (mod 4) by taking t even. However, in all cases, the number of vertices and subsystem size will be divisible by k2. 5 Multiple subsystems A natural question to ask is if a system can have multiple subsystems. In general, it seems likely to be hard to navigate through the lattice of subsystems and all the possible ways the subsystems can be distributed across the main system. However, when the subsystems are disjoint, have small common intersections or are nested, the problem is more tractable. We give some preliminary results in the next three subsections. 500 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 5.1 Disjoint subsystems In the uniform case, the flexibility of Theorem 1.2 allows us to create a large number of disjoint subsystems. We refer to a factorization of K∗v into k-cycles with subsystems on disjoint vertex sets of sizes mj for 1 ≤ j ≤ n as a UMSF(v, {mj}, k). Lemma 5.1. Let k | mj for 1 ≤ j ≤ n. Let m be an integer such that there is a UMSF(m,mj , k) for each 1 ≤ j ≤ n. Then there exists a UMSF(ms, {mj}, k) for all s ≥ max{2, n} if k is even, and for all s ≥ max{3, n} such that (s − 1)m is even if k is odd, except when (k, s,m) = (6, 2, 6). Proof. Theorem 1.2 guarantees the existence of a Ck-factorization of Ks[m]. For each 1 ≤ j ≤ n, place a copy of the ingredient UMSF(m,mj , k) on the jth part of size m of Ks[m], and any Ck-factorization of K∗m on each of the remaining parts. The definite exception (k, s,m) = (6, 2, 6) follows from the non-existence of a UMSF(12, 6, 6) (see Theorem 4.5). As with the uniform factorizations containing a single subsystem in this paper, the easiest case is when mj | m for all 1 ≤ j ≤ n and either k is even or m is odd. Corollary 5.2. Let m = lcm{mj : j = 1, . . . , n}, and assume the following conditions are all satisfied: (1) k | mj for all j ∈ {1, . . . , n} and m | v; (2) k(m− 1) is even; (3) if k = 3, then mj ̸∈ {6, 12} for all j; (4) if (k,m) = (6, 12), then mj ̸= 6 for all j; (5) (v,m, k) ̸= (12, 6, 6); (6) v/m ≥ max{2, n} if k is even; (7) v/m ≥ max{3, n} and v is odd if k is odd. Then there exists a UMSF(v, {mj}, k). Proof. Since each mj is a divisor of m and conditions (1)–(4) hold, we can apply either Theorem 1.1 or Corollary 4.6, as needed, to ensure the existence of a UMSF(m,mj , k) for every j ∈ {1, . . . , n}. These are the ingredient designs needed in Lemma 5.1, which can be applied in view of conditions (5)–(7). We note that for any fixed multiset of mj , this corollary constructs UMSF(v, {mj}, k) for all but a finite number of v permitted by the necessary conditions when mj | v and either k is even or m is odd. P. Danziger et al.: On the minisymposium problem 501 5.2 Scattered subsystems The proof of Lemma 5.1 builds systems whose factors intersect either all of the subsystems or none of them. A balancing of the sizes of these intersections could be an interesting property. For instance, we could ask for systems whose factors do not intersect more than one subsystem. In other words, we ask for a Ck-factorization F of K∗v that contains n subsystems of sizes m1,m2, . . . ,mn, such that no two factors of any of the subsystems are contained in the same factor of F . We denote such a factorization by UMSF(v, [mj ], k) and say that the subsystems are scattered. Partial results in this direction can be easily obtained by making use of cycle frames. We recall that a k-cycle frame (k-CF) of Ks[m] is a decomposition of Ks[m] into holey Ck-factors; a holey Ck-factor is a vertex-disjoint union of k-cycles covering all vertices Ks[m] except those belonging to one part. The following result, proven in [8], provides necessary and sufficient conditions for the existence of k-cycle frames. Theorem 5.3 ([8]). Let m ≥ 2 and k, s ≥ 3. There exists a k-cycle frame of Ks[m] if and only if m is even, m(s− 1) ≡ 0 (mod k), k is even when s = 3, and (k,m, s) ̸= (6, 6, 3). By making use of Theorem 5.3, we obtain the following. Lemma 5.4. Let u ∈ {1, 2}. If there exists a UMSF(2m+ u,mj , k) for each 1 ≤ j ≤ n, then there exists a UMSF(2ms+ u, [mj ], k) whenever 2s ≡ 2 (mod k) and s ≥ n. Proof. Let n ≥ 1, 2s ≡ 2 (mod k) and s ≥ n. It follows that 2m(s − 1) ≡ 0 (mod k), k = 4 when s = 3, and (k, 2m, s) ̸= (6, 6, 3). Therefore, Theorem 5.3 guarantees the existence of a k-cycle frame F of Ks[2m]. Let Pi denote the i-th part of Ks[2m], for 1 ≤ i ≤ s. Also, let F = {Fij : 1 ≤ i ≤ s, 1 ≤ j ≤ m}, where the Fijs are the holey Ck-factors of F missing Pi, for 1 ≤ i ≤ s. By assumption, there is a UMSF(2m + u,mj , k) on Pi ∪ {∞1,∞u}, say Hi = {Hij : 1 ≤ j ≤ m}. It follows that F∗ = {Fij ∪Hij : 1 ≤ i ≤ s, 1 ≤ j ≤ m} is a Ck-factorization of K2ms+u with scattered subsystems of sizes m1,m2, . . . ,mn. Indeed, the factors of the subsystems belong to the Hijs, each of which belongs to exactly one factor of F∗. In the UMSF(2ms+u, [mj ], k) constructed in the proof of Lemma 5.4, two subsystems may intersect in 0, 1, or 2 vertices, which are necessarily in the set {∞1,∞u}. Theorem 4.5 provides sufficient conditions for the existence of a UMSF(v,m, k) if m is a divisor of v. From that, we easily obtain the following corollary. Corollary 5.5. Let u ∈ {1, 2}, k ≥ 3, and let k | mj | (2m + u) for each 1 ≤ j ≤ n. Then there exists a UMSF(2ms+ u, [mj ], k) whenever the following conditions hold: (1) mj is even or (2m+ u)/mj is odd, (2) 2n ≤ 2s and 2s ≡ 2 (mod k), except when (mj , k) ∈ {(6, 3), (12, 3)}, and except possibly when (2m + u,mj , k) = (12, 6, 6), or k is odd and 2m+ u = 2mj , for some j ∈ {1, . . . , n}. Note that for values of the triple (2m + u,mj , k) determining a possible exception in Corollary 5.5 it is possible for a UMSF(2ms + u, [mj ], k) to exist. However, our method cannot construct them because the UMSF(2m+u,mj , k) to use in the construction does not exist. It is possible that other construction methods would build a UMSF(2ms+u, [mj ], k). 502 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 5.3 Nested subsystems A scenario complementary to the susbsystems being all disjoint is when the subsystems are completely nested, on vertex sets M1 ⊃ M2 ⊃ · · · ⊃ Mn−1. We modify our notation slightly for this section to make it less cumbersome in this specific context. Definition 5.6. Let v = m0 > m1 > · · · > mn−1 > mn = 0 be non-negative integers. For 1 ≤ i ≤ ⌊m0−12 ⌋ and 0 ≤ j < n, let Ui,j be a 2-regular graph of order |V (Ui,j)| =  mj −mj+1, if 1 ≤ i ≤ ⌊mj+1−12 ⌋, mj , if ⌊mj+1+12 ⌋ ≤ i ≤ ⌊ mj−1 2 ⌋, 0, if ⌊mj+12 ⌋ ≤ i ≤ ⌊ m0−1 2 ⌋. A nested minisymposium factorization nMSF({Ui,j : 1 ≤ i ≤ ⌊m0−12 ⌋, 0 ≤ j < n}) is a 2-factorization F = {Fi : 1 ≤ i ≤ ⌊m0−12 ⌋} of K ∗ v such that • V (K∗v ) = M0 ⊃ M1 ⊃ · · · ⊃ Mn−1 ⊃ Mn = ∅ are nested sets with |Mj | = mj ; • each 2-factor Fi = ⋃n−1 j=0 Fi,j , where V (Fi,j) =  Mj \Mj+1 if 1 ≤ i ≤ ⌊mj+1−12 ⌋ Mj if ⌊mj+1+12 ⌋ ≤ i ≤ ⌊ mj−1 2 ⌋ ∅ otherwise, and each Fi,j is isomorphic to Ui,j ; • for every 0 ≤ ℓ < n, { ⋃n−1 j=ℓ Fi,j : 1 ≤ i ≤ ⌊ mℓ−1 2 ⌋} is a 2-factorization of a graph isomorphic to K∗mℓ . In other words, the factorization F of K∗v restricted to vertex set Mℓ factorizes a graph isomorphic to K∗mℓ into 2-factors whose structure is determined by the Ui,js. Our construction of nested minisymposium factorizations is most tidily expressed by defining holey factorizations. Definition 5.7. Given positive integers v and m with v ≥ m, let U = { Ui : 1 ≤ i ≤ ⌊ v − 1 2 ⌋} , be a collection of 2-regular graphs on v−m vertices for 1 ≤ i ≤ ⌊m−12 ⌋, and on v vertices for ⌊m+12 ⌋ ≤ i ≤ ⌊ v−1 2 ⌋. A holey factorization HF(U) is a decomposition F = {Fi : 1 ≤ i ≤ ⌊ v−1 2 ⌋ } of K∗v −G (i.e., K∗v minus the edges of G) where each Fi ∼= Ui and G ∼= K∗m. If v is even, then there is a 1-factor, Iv , on the vertices of K∗v whose edges are not present in K∗v − G. If v and m are both even there is a 1-factor, Jm, on the vertices of G which is a subgraph of Iv . If v is even and m odd, then no edges of Iv are induced on the vertices of G. If v is odd and m is even, then there is a 1-factor Jm on the vertices of G whose edges are present in K∗v −G. By removing the 2-factors of a subsystem or “filling the hole” with them (making the Jm in the hole coincide with the Im of the subsystem as required by the parities of v and m) we have an equivalence between the existence of minisymposium factorizations and holey factorizations. P. Danziger et al.: On the minisymposium problem 503 Theorem 5.8. Let T = {Ti : 1 ≤ i ≤ ⌊m−12 ⌋} be a 2-factorization of K ∗ m and U = { Ui : 1 ≤ i ≤ ⌊ v − 1 2 ⌋} , be a collection of 2-regular graphs on v−m vertices for 1 ≤ i ≤ ⌊m−12 ⌋ and on v vertices for ⌊m+12 ⌋ ≤ i ≤ ⌊ v−1 2 ⌋. Then a HF(U) exists if and only if a MSF({Ui : ⌊m+12 ⌋ ≤ i ≤ ⌊ v−1 2 ⌋}, {(Ti, Ui) : 1 ≤ i ≤ ⌊ m−1 2 ⌋}) exists. Because in a nested minisymposium factorization the holes are nested and emptying or filling them does not affect the edges outside the hole, this equivalence extends to nested minisymposium factorizations and shows that they can be constructed exactly when the various holey factorizations of K∗mj with holes of size mj+1 exist. Theorem 5.9. Let v = m0 > m1 > · · · > mn−1 > mn = 0 be positive integers. For 1 ≤ i ≤ ⌊m0−12 ⌋ and 0 ≤ j < n, let Ui,j be a 2-regular graph of order |V (Ui,j)| =  mj −mj+1, if 1 ≤ i ≤ ⌊mj+1−12 ⌋, mj , if ⌊mj+1+12 ⌋ ≤ i ≤ ⌊ mj−1 2 ⌋, 0, if ⌊mj+12 ⌋ ≤ i ≤ ⌊ m0−1 2 ⌋. A nested minisymposium factorization nMSF({Ui,j : 1 ≤ i ≤ ⌊m0−12 ⌋, 0 ≤ j < n}) exists if and only if a HF({Ui,j : 1 ≤ i ≤ ⌊mj−12 ⌋}) exists for each 0 ≤ j < n. Proof. The forward direction is proved simply by restricting the system to Mj and re- moving the subsystem on Mj+1. The converse is proved by a recursive construction starting with j = n − 1: in this case, a HF({Ui,n−1 : 1 ≤ i ≤ ⌊mn−1−12 ⌋}) is an nMSF({Ui,n−1 : 1 ≤ i ≤ ⌊mn−1−12 ⌋}), say Fn−1. At stage j < n − 1, use Theorem 5.8 to construct an nMSF({Ui,ℓ : 1 ≤ i ≤ ⌊mj−12 ⌋, j ≤ ℓ < n}), say Fj , by filling the hole in the HF({Ui,j : 1 ≤ i ≤ ⌊ mj−1 2 ⌋}) with the nMSF({Ui,ℓ : 1 ≤ i ≤ ⌊mj+1−12 ⌋, j + 1 ≤ ℓ < n}), denoted by Fj+1, built at stage j + 1. Between the extremes of disjoint and nested subsystems, there are factorizations with multiple subsystems with arbitrary intersections. Some structured instances of this much more general problem may be amenable to solution but we leave this to future work. 6 Conclusions and further work We have introduced the minisymposium problem: a subsystem variant of the generalized Oberwolfach problem. This variant asks for a solution to a generalized Oberwolfach prob- lem that contains a subsystem of a given size. When v, the number of vertices, is even, it is traditional in 2-factor decomposition problems to ask for decompositions of Kv − I where I is a 1-factor. When the number of vertices in the system and the subsystem are both even, then we require that the 1-factor in the subsystem be a subgraph of the 1-factor in the full system. Therefore when the parities of the system and the subsystem agree, the 504 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 problem becomes more tractable. When the parities are opposite, either the 1-factor of the full system must avoid the subsystem, or the edges of the 1-factor in the subsystem must be in 2-factors of the whole system. Clearly, this is a very broad statement and we identify some particularly interesting cases, Hamiltonian* and uniform. In the Hamiltonian* minisymposium problem there are as few cycles as possible and in the uniform minisymposium problem all cycles are of the same length. We have shown that the work of Hilton and Johnson provides a complete solution for the Hamiltonian* minisymposium problem in Corollary 3.2. In the case when v ≡ m ≡ 2 (mod 4), Theorem 3.4 uses this Hamiltonian* construction to provide a wide range of solutions when the resulting factors are all bipartite. In particular, a uniform factorization with k ≡ 2 (mod 4), v ≥ 2m and v ≡ m ≡ k (mod 2k) always exists. Corollary 4.10 can be used to extend this to uniform factorizations where k ≡ 0 (mod 4) or v ≡ m ≡ 0 (mod 2k). In Section 4 we considered the uniform case. We have solved a large part of the spec- trum. In particular, when k is even or v is odd, Corollary 4.6 gives all cases when m | v and Corollary 4.7 completely solves all cases when k = m has the same parity as v. Theo- rem 4.8 gives a powerful recursive construction which is applicable in cases where m does not divide v. By applying it to the case when k = 3, we obtain uniform factorizations with cycle lengths divisible by 3. The case when m is odd and v is even seems to be the hardest. Even in the simplest case when k = 3, which has been well studied otherwise [18, 19, 20, 22, 34, 35, 36, 37], the case with v and m having opposite parities has not been previously considered and remains open. While the Hamiltonian* problem is solved and we have made significant inroads into the uniform case, the general problem remains wide open. We expect that when m and v have the same parity solutions will be easier to find. When m and v have opposite parity we expect that v odd with m even is more tractable than the reverse. Considering 2-factorizations where a solution to the Oberwolfach problem is known might be a good starting point. A natural case to consider is the case when all factors are isomorphic i.e. Fi ∼= Tj ∪ Uj for all i and j. Uniform factorizations are an example of this, but other variations are possible, for example, requiring all factors to be isomorphic to Cv−m ∪Cm. Indeed, Theorem 3.4 solves all these cases when the factors are bipartite and v ≡ m ≡ 2 (mod 4), but this broader variant remains open. More complex variants can also be considered. We have briefly considered systems with multiple subsystems. When these subsystems are completely nested the problem es- sentially reduces to the existence of the necessary ingredients as described in Theorem 5.9. Let {mj} be a multiset of subsystem sizes. When the subsystems are pairwise disjoint, v is divisible by each mj and either v is odd or at least one subsystem is even, then Lemma 5.1 and Corollary 5.2 use Theorem 1.2 to construct a UMSF(v, {mj}, k) for all but a finite number of admissible v. Even in the seemingly simple case when the subsystems are all disjoint the problem remains generally open even for the uniform case. Further partial re- sults are obtained when the subsystems are scattered, that is, when no two minisymposia have meetings taking place on the same day. Cycle frames in Theorem 5.3 allow us to construct uniform factorizations as described in Lemma 5.4 and Theorem 5.5. The more general case when the intersections of multiple subsystems are arbitrary is completely open. P. Danziger et al.: On the minisymposium problem 505 ORCID iDs Tommaso Traetta https://orcid.org/0000-0001-8141-0535 References [1] B. Alspach and R. Häggkvist, Some observations on the Oberwolfach problem, J. Graph Theory 9 (1985), 177–187, doi:10.1002/jgt.3190090114, https://doi.org/10.1002/ jgt.3190090114. [2] B. Alspach, P. J. Schellenberg, D. R. Stinson and D. Wagner, The Oberwolfach problem and factors of uniform odd length cycles, J. Comb. Theory Ser. A 52 (1989), 20–43, doi:10.1016/ 0097-3165(89)90059-9, https://doi.org/10.1016/0097-3165(89)90059-9. [3] J. Asplund, D. Kamin, M. Keranen, A. Pastine and S. Özkan, On the Hamilton-Waterloo problem with triangle factors and C3x-factors, Australas. J. Comb. 64 (2016), 458–474, https://ajc.maths.uq.edu.au/pdf/64/ajc_v64_p458.pdf. [4] Z. Baranyai and G. R. Szász, Hamiltonian decomposition of lexicographic product, J. Comb. Theory Ser. B 31 (1981), 253–261, doi:10.1016/0095-8956(81)90028-9, https://doi. org/10.1016/0095-8956(81)90028-9. [5] S. Bonvicini and M. Buratti, Octahedral, dicyclic and special linear solutions of some Hamilton-Waterloo problems, Ars Math. Contemp. 14 (2018), 1–14, doi:10.26493/1855-3974. 1088.414, https://doi.org/10.26493/1855-3974.1088.414. [6] D. Bryant and P. Danziger, On bipartite 2-factorizations of Kn − I and the Oberwolfach prob- lem, J. Graph Theory 68 (2011), 22–37, doi:10.1002/jgt.20538, https://doi.org/10. 1002/jgt.20538. [7] D. Bryant, P. Danziger and M. Dean, On the Hamilton-Waterloo problem for bipartite 2-factors, J. Comb. Des. 21 (2013), 60–80, doi:10.1002/jcd.21312, https://doi.org/10.1002/ jcd.21312. [8] M. Buratti, H. Cao, D. Dai and T. Traetta, A complete solution to the existence of (k, λ)- cycle frames of type gu, J. Comb. Des. 25 (2017), 197–230, doi:10.1002/jcd.21523, https: //doi.org/10.1002/jcd.21523. [9] M. Buratti and F. Zuanni, Perfect Cayley Designs as Generalizations of Perfect Mendel- sohn Designs, Des. Codes Cryptogr. 23 (2001), 233—-248, doi:10.1023/a:1011272801679, https://doi.org/10.1023/a:1011272801679. [10] A. C. Burgess, P. Danziger and T. Traetta, On the Hamilton-Waterloo problem with odd or- ders, J. Comb. Des. 25 (2017), 258–287, doi:10.1002/jcd.21552, https://doi.org/10. 1002/jcd.21552. [11] A. C. Burgess, P. Danziger and T. Traetta, On the Hamilton-Waterloo problem with cycle lengths of distinct parities, Discrete Math. 341 (2018), 1636–1644, doi:10.1016/j.disc.2018. 02.020, https://doi.org/10.1016/j.disc.2018.02.020. [12] A. C. Burgess, P. Danziger and T. Traetta, On the Hamilton-Waterloo problem with odd cycle lengths, J. Comb. Des. 26 (2018), 51–83, doi:10.1002/jcd.21586, https://doi.org/10. 1002/jcd.21586. [13] N. J. Cavenagh, S. I. El-Zanati, A. Khodkar and C. Vanden Eynden, On a generalization of the Oberwolfach problem, J. Comb. Theory Ser. A 106 (2004), 255–275, doi:10.1016/j.jcta.2004. 02.003, https://doi.org/10.1016/j.jcta.2004.02.003. [14] C. J. Colbourn and J. H. Dinitz (eds.), Handbook of Combinatorial Designs, Discrete Mathe- matics and its Applications (Boca Raton), Chapman & Hall/CRC, Boca Raton, FL, 2nd edition, 2007. 506 Ars Math. Contemp. 24 (2024) #P3.07 / 489–507 [15] S. Costa, A complete solution to the infinite Oberwolfach problem, J. Comb. Des. 28 (2020), 366–383, doi:10.1002/jcd.21702, https://doi.org/10.1002/jcd.21702. [16] P. Danziger, E. Mendelsohn, B. Stevens and T. Traetta, Hamiltonian* minisymposium factor- izations, 2023, preprint. [17] P. Danziger, G. Quattrocchi and B. Stevens, The Hamilton-Waterloo problem for cycle sizes 3 and 4, J. Comb. Des. 17 (2009), 342–352, doi:10.1002/jcd.20219, https://doi.org/10. 1002/jcd.20219. [18] D. Deng, R. Rees and H. Shen, Further results on nearly Kirkman triple systems with sub- systems, Discrete Math. 270 (2003), 99–114, doi:10.1016/S0012-365X(02)00749-5, https: //doi.org/10.1016/S0012-365X(02)00749-5. [19] D. Deng, R. Rees and H. Shen, On the existence and application of incomplete nearly Kirkman triple systems with a hole of size 6 or 12, Discrete Math. 261 (2003), 209–233, doi:10.1016/ S0012-365X(02)00469-7, papers on the occasion of the 65th birthday of Alex Rosa, https: //doi.org/10.1016/S0012-365X(02)00469-7. [20] D. Deng, R. Rees and H. Shen, On the existence of nearly Kirkman triple systems with sub- systems, Des. Codes Cryptogr. 48 (2008), 17–33, doi:10.1007/s10623-007-9166-2, https: //doi.org/10.1007/s10623-007-9166-2. [21] S. I. El-Zanati, S. K. Tipnis and C. Vanden Eynden, A generalization of the Oberwolfach prob- lem, J. Graph Theory 41 (2002), 151–161, doi:10.1002/jgt.10058, https://doi.org/10. 1002/jgt.10058. [22] G. Ge and R. Rees, On group-divisible designs with block size four and group-type 6um1, Dis- crete Math. 279 (2004), 247–265, doi:10.1016/S0012-365X(03)00272-3, https://doi. org/10.1016/S0012-365X(03)00272-3. [23] S. Glock, F. Joos, J. Kim, D. Kühn and D. Osthus, Resolution of the Oberwolfach problem, Acta Math. Univ. Comenian. (N.S.) 88 (2019), 735–741. [24] R. Häggkvist, A lemma on cycle decompositions, in: Cycles in graphs (Burnaby, B.C., 1982), North-Holland, Amsterdam, volume 115 of North-Holland Math. Stud., pp. 227–232, 1985, doi:10.1016/S0304-0208(08)73015-9, https://doi.org/10.1016/ S0304-0208(08)73015-9. [25] A. J. W. Hilton and M. Johnson, Some results on the Oberwolfach problem, J. London Math. Soc. (2) 64 (2001), 513–522, doi:10.1112/s0024610701002666, https://doi.org/10. 1112/s0024610701002666. [26] D. G. Hoffman and P. J. Schellenberg, The existence of Ck-factorizations of K2n−F , Discrete Math. 97 (1991), 243–250, doi:10.1016/0012-365x(91)90440-d, https://doi.org/10. 1016/0012-365x(91)90440-d. [27] M. S. Keranen and S. Özkan, The Hamilton-Waterloo problem with 4-cycles and a single factor of n-cycles, Graphs Comb. 29 (2013), 1827–1837, doi:10.1007/s00373-012-1231-6, https://doi.org/10.1007/s00373-012-1231-6. [28] M. S. Keranen and A. Pastine, A generalization of the Hamilton-Waterloo problem on complete equipartite graphs, J. Comb. Des. 25 (2017), 431–468, doi:10.1002/jcd.21560, https:// doi.org/10.1002/jcd.21560. [29] H. Lei and H. Shen, The Hamilton-Waterloo problem for Hamilton cycles and triangle- factors, J. Comb. Des. 20 (2012), 305–316, doi:10.1002/jcd.20311, https://doi.org/ 10.1002/jcd.20311. [30] J. Liu, A generalization of the Oberwolfach problem and Ct-factorizations of complete equipar- tite graphs, J. Comb. Des. 8 (2000), 42–49, doi:10.1002/(SICI)1520-6610(2000)8:1⟨42:: P. Danziger et al.: On the minisymposium problem 507 AID-JCD6⟩3.0.CO;2-R, https://doi.org/10.1002/(SICI)1520-6610(2000) 8:1<42::AID-JCD6>3.0.CO;2-R. [31] J. Liu, The equipartite Oberwolfach problem with uniform tables, J. Comb. Theory Ser. A 101 (2003), 20–34, doi:10.1016/S0097-3165(02)00011-0, https://doi.org/10.1016/ S0097-3165(02)00011-0. [32] U. Odabaşıand S. Özkan, The Hamilton-Waterloo problem with C4 and Cm factors, Dis- crete Math. 339 (2016), 263–269, doi:10.1016/j.disc.2015.08.013, https://doi.org/ 10.1016/j.disc.2015.08.013. [33] W.-L. Piotrowski, The solution of the bipartite analogue of the Oberwolfach problem, Discrete Math. 97 (1991), 339–356, doi:10.1016/0012-365x(91)90449-c, https://doi.org/10. 1016/0012-365x(91)90449-c. [34] R. Rees and D. R. Stinson, Kirkman triple systems with maximum subsystems, Ars Comb. 25 (1988), 125–132. [35] R. Rees and D. R. Stinson, On the existence of Kirkman triple systems containing Kirkman subsystems, Ars Comb. 26 (1988), 3–16. [36] D. R. Stinson, Frames for Kirkman triple systems, Discrete Math. 65 (1987), 289–300, doi:10.1016/0012-365x(87)90060-4, https://doi.org/10.1016/0012-365x(87) 90060-4. [37] S. Tang and H. Shen, Embeddings of nearly Kirkman triple systems, J. Statist. Plann. In- ference 94 (2001), 327–333, doi:10.1016/s0378-3758(00)00263-9, second Shanghai Confer- ence on Designs, Codes and Finite Geometries (1996), https://doi.org/10.1016/ s0378-3758(00)00263-9. [38] T. Traetta, A complete solution to the two-table Oberwolfach problems, J. Comb. Theory Ser. A 120 (2013), 984–997, doi:10.1016/j.jcta.2013.01.003, https://doi.org/10.1016/j. jcta.2013.01.003. [39] L. Wang and H. Cao, A note on the Hamilton-Waterloo problem with C8-factors and Cm- factors, Discrete Math. 341 (2018), 67–73, doi:10.1016/j.disc.2017.06.024, https://doi. org/10.1016/j.disc.2017.06.024. [40] L. Wang, F. Chen and H. Cao, The Hamilton-Waterloo problem for C3-factors and Cn- factors, J. Comb. Des. 25 (2017), 385–418, doi:10.1002/jcd.21561, https://doi.org/ 10.1002/jcd.21561. [41] D. B. West, Introduction to Graph Theory, Prentice Hall, Inc., Upper Saddle River, NJ, 2001. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.08 / 509–534 https://doi.org/10.26493/1855-3974.3035.6ac (Also available at http://amc-journal.eu) Two kinds of partial Motzkin paths with air pockets* Jean-Luc Baril † LIB, Université de Bourgogne, B.P. 47 870, 21078 Dijon Cedex, France Paul Barry School of Science, South East Technological University (SETU), Ireland Received 23 December 2022, accepted 23 October 2023, published online 16 August 2024 Abstract Motzkin paths with air pockets (MAP) are defined as a generalization of Dyck paths with air pockets by allowing some horizontal steps with certain conditions. In this paper, we introduce two generalizations. The first one consists of lattice paths in N2 starting at the origin, made of steps U = (1, 1), Dk = (1,−k), k ⩾ 1 and H = (1, 0), where two down steps cannot be consecutive, while the second one are lattice paths in N2 starting at the origin, made of steps U , Dk and H , where each step Dk and H is necessarily followed by an up step, except for the last step of the path. We provide enumerative results for these paths according to the length, the type of the last step, and the height of its end-point. A similar study is made for these paths read from right to left. As a byproduct, we obtain new classes of paths counted by the Motzkin numbers. Finally, we express our results using Riordan arrays. Keywords: Enumeration, Motzkin paths, kernel method, Riordan array. Math. Subj. Class. (2020): 05A05, 05A15, 05A15 1 Introduction In a recent paper [2], the authors introduce, study and enumerate special classes of lattice paths, called Dyck paths with air pockets (DAP for short). Such paths are non empty lattice paths in the first quadrant of Z2 starting at the origin, and consisting of up-steps U = (1, 1) and down-steps Dk = (1,−k), k ⩾ 1, where two down steps cannot be consecutive. *The authors would like to thank the anonymous referees for useful remarks and comments. †Corresponding author. E-mail addresses: barjl@u-bourgogne.fr (Jean-Luc Baril), pbarry@wit.ie (Paul Barry) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 510 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 The length of a path is the number of its steps. These paths can be viewed as ordinary Dyck paths where each maximal run of down-steps is condensed into one large down step. As mentioned in [2], they also correspond to a stack evolution with (partial) reset opera- tions that cannot be consecutive (see for instance [7]). Whenever the last point is on the x-axis, they prove that DAP of length n are in one-to-one correspondence with the peakless Motzkin paths of length n − 1. They also investigate the popularity of many patterns in these paths and they give asymptotic approximations. We also refer to [10] for the enu- meration of DAP with respect to the length, the type (up or down) of the last step and the height of the end-point (i.e., the ordinate of the end-point). In a second work [3], the au- thors make a study for a generalization of these paths by allowing them to go below the x-axis. They call these paths Grand Dyck paths with air pockets (GDAP), and they also yield enumerative results for these paths according to the length and several restrictions on the height. In this paper, we introduce two generalizations of partial Dyck paths of air pockets by allowing some possible horizontal steps H = (1, 0) with some conditions. These two kinds of paths can be viewed as special partial Motzkin paths (lattice paths in N2 starting at the origin and made of U = (1, 1), D = (1,−1), and H = (1, 0)), where each maximal run of down-steps is condensed into one large down step. Definition 1.1. A partial Motzkin path with air pocket of the first kind is a lattice path in N2 starting at the origin, consisting of steps U , Dk and H , where two down steps cannot be consecutive. Let M1 be the set of these paths. Definition 1.2. A partial Motzkin path with air pocket of second kind is a lattice path in N2 starting at the origin, consisting of steps U , Dk and H , where any step H and Dk (except the last step of the path) is immediately followed by an up step U . Let M2 be the set of these paths. Whenever these paths end on the x-axis, we call them Motzkin paths with air pockets of the first and second kinds, respectively. For short, we denote by PMAP all paths in Mi, i ∈ {1, 2}, and we denote by MAP all paths ending on the x-axis. For instance, UUUDHUD2 ∈ M1 and UUUDUHUD2 ∈ M2 are two PMAP of the first and second kinds, respectively. The paths UUUDHUD3UD and UUUDUHUD4 UD are MAP of the first and second kinds, respectively. See also Figure 1 and Figure 5 for other examples of PMAP and MAP. We also consider lattice paths obtained from MAP in M1 and M2 by reading them from right to left and by replacing down-steps with up-steps and vice versa. More precisely, we define the two sets M′1 and M′2 as follows. Definition 1.3. Let M′1 be the set of paths in N2 starting at the origin, consisting of up steps Uk = (1, k), k ⩾ 1, horizontal step H and down step D, where two up steps cannot be consecutive. Definition 1.4. Let M′2 be the set of paths in N2 starting at the origin, consisting of up steps Uk = (1, k), k ⩾ 1, horizontal step H and down step D, where any H and Uk, k ⩾ 1, (except the first step of the path) is preceded by a down step D. All paths in M′1 and M′2 will be be called PMAP from right to left, and in the case where they end on the x-axis, we call them MAP from right to left. For instance, we have UDU3DHUD ∈ M′1, UDU4DHDU ∈ M′2, and the path UDU3DHUDDD J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 511 (resp. UDU4DHDUDDD) is a MAP of the first (resp. second) kind from right to left. See also Figure 3 and Figure 7 for other examples of PMAP and MAP from right to left. Notice that D1 = D and U1 = U , and we will use both notations. Throughout the paper, and for each class of paths Mi and M′i, i ∈ {1, 2}, described above, we will use the following notations. For k ⩾ 0, we consider the generating function fk = fk(z) (resp. gk = gk(z), resp. hk = hk(z)), where the coefficient of zn in the series expansion is the number of partial Motzkin paths with air pockets of length n ending at height k with an up-step, (resp. with a down-step, resp. with a horizontal step H). We introduce the bivariate generating functions F (z, u) = ∑ k⩾0 ukfk(z), G(z, u) = ∑ k⩾0 ukgk(z), and H(z, u) = ∑ k⩾0 ukhk(z). For short, we also use the notation F (u), G(u) and H(u) for these functions, and we introduce the generating function Total(z, u) = F (u) +G(u) +H(u). Finally, for any bivariate generating function H(z, u) = H(u), we will use the notation [uk]H(u) for the coefficient of uk in H(u). The outline of this paper is the following. In Section 2, we present enumerative results for partial Motzkin paths with air pockets of the first kind, and for these paths when we read them from right to left. We provide bivariate generating functions that count these paths with respect to the length, the type of the last step (up, down or horizontal step) and the height of the end-point. In Section 3, we make a similar study for PMAP of second kind, and we present new classes of lattice paths counted by the well known Motzkin numbers. All these results are obtained algebraically by using the famous kernel method for solving several systems of functional equations. More precisely, Sections 2.1, 2.2, 3.1 and 3.2 have the same structure: we show how to obtain a system of equations involving fk, gk and hk, and we apply the kernel method in order to provide some expressions of F (u), G(u) and H(u). Finally, in Section 4 we express our results using Riordan arrays and we deduce closed forms for PMAP of length n ending at height k. 2 PMAP of the first kind In this section, we focus on PMAP of the first kind, i.e. lattices paths in N2 starting at the origin, made of steps U , Dk and H , such that two down steps cannot be consecutive. The first subsection considers the paths in M1, while the second subsection handles the paths in M′1 (see Introduction for the definition of these two sets). We yield enumerative results for these paths according to the length, the type of the last step, and the height of its end-point. 2.1 PMAP in M1 - From left to right In this part, we consider PMAP in M1. Figure 1 shows two examples of such paths. Let P be a length n PMAP in M1 ending at height k ⩾ 0. If the last step of P is U , then k ⩾ 1 and P can be written P = QU where Q is a length (n − 1) PMAP ending at height k − 1. So, we obtain the first relation fk = zfk−1 + zgk−1 + zhk−1 for k ⩾ 1, anchored with f0 = 1 by considering the empty path. If the last step of P is a down step Da, a ⩾ 1, then we have P = QDa where Q is a length (n − 1) PMAP ending at height 512 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Figure 1: The left drawing shows a Motzkin path with air pockets of length 18 in M1. The right drawing shows a partial Motzkin path with air pockets of length 18 ending at height 3 in M1. ℓ = a + k ⩾ k + 1 with no down step at its end. So, we obtain the second relation gk = z ∑ ℓ⩾k+1 fℓ + z ∑ ℓ⩾k+1 hℓ. If the last step of P is a horizontal step H , then we have P = QH where Q is a length (n − 1) PMAP ending at height k, which implies hk = zfk + zgk + zhk. Therefore, we have to solve the following system of equations. f0 = 1, and fk = zfk−1 + zgk−1 + zhk−1, k ⩾ 1, gk = z ∑ ℓ⩾k+1 fℓ + z ∑ ℓ⩾k+1 hℓ, k ⩾ 0, hk = zfk + zgk + zhk, k ⩾ 0. (2.1) Summing the recursions in (2.1), we have: F (u) = 1 + z ∑ k⩾1 ukfk−1 + z ∑ k⩾1 ukgk−1 + z ∑ k⩾1 ukhk−1 = 1 + zuF (u) + zuG(u) + zuH(u), G(u) = z ∑ k⩾0 uk ( ∑ ℓ⩾k+1 fℓ ) + z ∑ k⩾0 uk ( ∑ ℓ⩾k+1 hℓ ) = z ∑ k⩾1 fk(1 + u+ . . .+ u k−1) + z ∑ k⩾1 hk(1 + u+ . . .+ u k−1) = z ∑ k⩾1 uk − 1 u− 1 fk + z ∑ k⩾1 uk − 1 u− 1 hk = z u− 1 (F (u)− F (1) +H(u)−H(1)), H(u) = zF (u) + zG(u) + zH(u). Notice that we have F (1) − H(1) = 1 by considering the difference of the first and third equations. Now, setting a := F (1) and solving these functional equations, we obtain F (u) = 2au z2 − u z2 + uz + z2 − u− z + 1 u2z + u z2 + z2 − u− z + 1 , G(u) = −z (2auz + 2az − uz − 2a− z + 2) u2z + u z2 + z2 − u− z + 1 , H(u) = z (2az − u− 2z + 1) u2z + u z2 + z2 − u− z + 1 . J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 513 In order to compute a = F (1), we use the kernel method (see [1, 9]) on F (u). This method consists in cancelling the denominator by finding u as an algebraic function of z, s(z). So, if we substitute u with s(z) in the numerator, then it necessarily equals zero (in order to counterbalance the cancellation of the denominator), which allows to find a = F (1). Finally, we can deduce the generating function F (u). We can write the denominator (which is a polynomial in u of degree 2), as z(u − r) (u− s) with r = 1− z2 + √ z4 − 4z3 + 2z2 − 4z + 1 2z , s = 1− z2 − √ z4 − 4z3 + 2z2 − 4z + 1 2z . Replacing u with s (which have a Taylor expansion at z = 0) in order to cancel the numer- ator of F (u), we obtain the equation 2as z2 − s z2 + sz + z2 − s− z + 1 = 0. Using zrs = z2 − z + 1, a straightforward calculation provides a = F (1) = H(1) + 1 = r(s− 1) 2z . Finally, after simplifying by the factor (u − s) in the numerators and denominators, we obtain F (u) = r r − u , G(u) = r(s− 1)− z r − u , and H(u) = 1 r − u . Extracting the coefficient of uk in the series expansion of each generating function, we obtain [uk]F (u) := fk = 1 rk , [uk]G(u) := gk = s− 1 rk − z rk+1 , [uk]H(u) := hk = 1 rk+1 . Theorem 2.1. The bivariate generating function for the total number of PMAP in M1 with respect to the length and the height of the end-point is given by Total(z, u) = 1 z(r − u) , and we have [uk]Total(z, u) = 1 zrk+1 . Finally, setting t(n, k) := [zn][uk]Total(z, u), we have for n ⩾ 2, k ⩾ 1, t(n, k) = t(n, k − 1) + t(n− 1, k)− t(n− 1, k − 2)− t(n− 2, k)− t(n− 2, k − 1), and setting tn := t(n, 0), then we have t0 = t1 = 1, and for n ⩾ 2, tn = tn−1 + tn−2 + n−3∑ k=0 tktn−k−3 + n−1∑ k=2 (tk − tk−1) tn−k−1. 514 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Proof. Since Total(z, u) = F (u)+G(u)+H(u), the first two equalities are immediately deduced from the previous results. The third equality is obtained using the Mathematica package Guess.m (written by Manuel Kauers [6]) for guessing recurrence relations. After this, it suffices to check algebraically: Total(z, u) = (u+ z − u2z − z2 − uz2)Total(z, u)− u+ (z 2 − z + 1)(1 + rs− z) r . Now, let us prove the last equality. Any length n MAP of the first kind is of the form (i) HP , or (ii) UDP , or (iii) UPHDQ where P,Q are some MAP so that the length k of P lies into [0, n− 3], and the length of Q is n− k − 3, or (iv) P ♯Q where P ♯ = UP ′Dℓ, ℓ ⩾ 1, and P ′ is a PMAP such that P ′Dℓ−1 is a MAP of length k lying into [2, n− 1], and the length of Q is n− k − 1. Taking into account all these cases, we obtain the result. Let T be the infinite matrix T := [t(n, k)]n⩾0,k⩾0. The first few rows of the matrix T are T =  1 0 0 0 0 0 0 0 · · · 1 1 0 0 0 0 0 0 · · · 2 2 1 0 0 0 0 0 · · · 5 5 3 1 0 0 0 0 · · · 13 14 9 4 1 0 0 0 · · · 36 40 28 14 5 1 0 0 · · · 105 118 87 48 20 6 1 0 · · · 317 359 273 161 75 27 7 1 · · · ... ... ... ... ... ... ... ... . . .  . Corollary 2.2. The generating function that counts the PMAP in M1 with respect to the length is given by Total(z, 1) = 1 z(r − 1) . The first few terms of the series expansion of Total(z, 1) are 1 + 2z + 5z2 + 14z3 + 41z4 + 124z5 + 385z6 + 1220z7 + 3929z8 + 12822z9 + O(z10), which correspond to the sequence A159771 in [13] counting the n-leaf binary trees that do not contain (()((()())((()())()))) as a subtree (see [11]). See Figure 2 for an illustration of the 14 PMAP of length 3. Corollary 2.3. The generating function that counts the MAP in M1 with respect to the length is given by Total(z, 0) = 1 zr . The first few terms of the series expansion of Total(z, 0) are 1 + z + 2z2 + 5z3 + 13z4+36z5+105z6+317z7+982z8+3105z9+O(z10) which correspond to the sequence A114465 in [13] counting Dyck paths of length 2n having no ascents of length 2 that start at an odd level. We leave open the question of finding a constructive bijection between these sets. J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 515 Figure 2: The 14 PMAP of length three in M1. Notice that five paths end on the x-axis, five paths end at height 1, three paths end at height 2, and one path ends at height 3, which correspond to the fourth row of T . 2.2 PMAP in M′1 - From right to left Here, we consider the paths of the previous section, but we read them from right to left. This means that down steps become up steps and vice versa, and horizontal steps are unchanged, which implies that two up steps cannot be consecutive now. See Definition 1.3 and Figure 3 for two examples of such paths. Figure 3: The left drawing shows a Motzkin path with air pockets of length 18 in M′1. The right drawing shows a partial Motzkin path with air pockets of length 18 ending at height 2 in M′1. Let P be a length n PMAP in M′1 ending at height k ⩾ 0. If the last step of P is Ua, a ⩾ 1, then k ⩾ a and we have P = QUa where Q is a length (n − 1) PMAP ending at height ℓ = k − a with a horizontal step or a down step. So, we obtain the first relation fk = z + z(g0 + g1 + . . . + gk−1) + z(h0 + h1 + . . . + hk−1) for k ⩾ 1, anchored with f0 = 1 by considering the empty path. If the last step of P is a down step D, then we have P = QD where Q is a length (n − 1) PMAP ending at height k + 1. So, we obtain the second relation gk = zfk+1 + zgk+1 + zhk+1. If the last step of P is a horizontal step H , then we have P = QH where Q is a length (n − 1) PMAP ending at height k, which implies hk = zfk + zgk + zhk. Therefore, we have to solve the following system of equations.  f0 = 1, fk = z + z(g0 + g1 + . . .+ gk−1) + z(h0 + h1 + . . .+ hk−1), k ⩾ 1, gk = zfk+1 + zgk+1 + zhk+1, k ⩾ 0, hk = zfk + zgk + zhk, k ⩾ 0. (2.2) 516 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Summing the recursions in (2.2), we have: F (u) = 1 + zu 1− u + z ∑ k⩾1 (g0 + . . .+ gk−1)u k + z ∑ k⩾1 (h0 + . . .+ hk−1)u k = 1 + zu 1− u + z ∑ k⩾0 gk uk+1 1− u + z ∑ k⩾0 hk uk+1 1− u = 1 + zu 1− u (1 +G(u) +H(u)), G(u) = z ∑ k⩾0 fk+1u k + z ∑ k⩾0 gk+1u k + z ∑ k⩾0 hk+1u k = z u (F (u)− F (0) +G(u)−G(0) +H(u)−H(0)), H(u) = zF (u) + zG(u) + zH(u). Notice that we have H(0) = z(1+G(0))1−z by considering the third equation and F (0) = 1. Solving these functional equations using a := G(0), we obtain F (u) = −u2z3 + a uz2 + 3u2z2 − uz3 − 3u2z + 2uz2 + u2 + uz − z2 − u+ z (1− z) (u2z2 − u2z + uz2 + u2 − u+ z) , G(u) = z ( a uz2 − a uz + a u+ a z + uz − a ) (−1 + z) (u2z2 − u2z + uz2 + u2 − u+ z) , H(u) = ( a uz2 + u2z2 − a uz − 2u2z + uz2 + a z + u2 − u+ z ) z (1− z) (u2z2 − u2z + uz2 + u2 − u+ z) . In order to compute a = G(0), we use the kernel method on G(u). We can write the denominator (which is a polynomial in u of degree 2), as (z−1)(z2−z+1)(u−r)(u−s) with r = 1− z2 + √ z4 − 4 z3 + 2 z2 − 4 z + 1 2(z2 − z + 1) , s = 1− z2 − √ z4 − 4 z3 + 2 z2 − 4 z + 1 2(z2 − z + 1) . Replacing u with s (which have a Taylor expansion at z = 0) in order to cancel the numer- ator of G(u), we obtain the equation a(sz2 − sz + s+ z − 1) + sz = 0. Using sr(z2 − z + 1) = z, we deduce a = G(0) = 1− r r − sz. Finally, after simplifying by the factor (u− s) in the numerators and denominators, we obtain F (u) = 1 + sru r − u , G(u) = s(1− r + rz) r − u , and H(u) = sr − sru(1− z) r − u , J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 517 which implies that [uk]F (u) := fk = [k = 0] + [k ̸= 0] · s rk−1 , [uk]G(u) := gk = s(1− r + rz) rk+1 , [uk]H(u) := hk = s rk − [k ̸= 0] · (1− z)s rk−1 , where [k = 0] (resp. [k ̸= 0]) equals 1 whenever k = 0 (resp. k ̸= 0), and 0 otherwise. Theorem 2.4. The bivariate generating function for the total number of PMAP (read from right to left) with respect to the length and the height of the end-point is given by Total(z, u) = 1 + s(1 + rz + ruz) r − u , and we have [uk]Total(z, u) = [k = 0] + s(rz + 1) rk+1 + [k ̸= 0] · sz rk−1 . Finally, setting t(n, k) := [zn][uk]Total(z, u), we have for n ⩾ 2, k ⩾ 1, t(n, k) = t(n− 2, k− 1) + t(n− 2, k)− t(n− 1, k− 1) + t(n− 1, k+ 1) + t(n, k− 1), and setting tn := t(n, 0), we have t0 = t1 = 1, and for n ⩾ 2, tn = tn−1 + tn−2 + n−3∑ k=0 tktn−k−3 + n−1∑ k=2 (tk − tk−1) tn−k−1. Proof. The first two equalities are directly deduced from the previous results. Since we have (u− u2z2 − uz2 + zu2 − z − u2)Total(z, u) + u2(1− z)− (1 + s− sz)u+ s = 0, we deduce the third relation. The last equality is already given in Theorem 2.1, since the number of MAP in M′1 is obviously equal to the number of MAP in M1. Let T be the infinite matrix T := [t(n, k)]n⩾0,k⩾0. The first few rows of the matrix T are T =  1 0 0 0 0 0 · · · 1 1 1 1 1 1 · · · 2 3 3 3 3 3 · · · 5 8 10 12 14 16 · · · 13 23 33 43 53 63 · · · 36 69 107 149 195 245 · · · 105 212 348 512 704 924 · · · 317 665 1141 1753 2509 3417 · · · ... ... ... ... ... ... . . .  . 518 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Since there is an infinite number of PMAP of length n, we do not provide an ordinary generating function (with respect to the length) for these paths. So, we get around this by counting PMAP ending on a point (x, n− x) for a given n ⩾ 0. Corollary 2.5. The generating function that counts the partial PMAP ending on the line y = n− x is given by Total(z, z) = 1 + s(1 + rz + rz2) r − z . The first few terms of the series expansion of Total(z, z) are 1 + z + 3z2 + 9z3 + 25z4 + 73z5 + 223z6 + 697z7 + 2217z8 + 7161z9 + O(z10), which correspond to the sequence A101499 in [13], which is a Chebyshev transform of the Catalan number that counts peakless Motzkin paths of length n where horizontal steps at level at least one come in 2 colors. See Figure 4 for an illustration of the 9 PMAP in M′1 ending on the line y = 3− x. Notice that, from Theorem 2.4 we have Total(z, 0) = 1 + s(1 + rz) r , which is obviously equal to the expression derived in Corollary 2.3 that counts MAP of the first kind with respect to the length. Figure 4: The 9 PMAP ending on the line y = 3− x in M′1. Notice that five paths end on the x-axis, three paths end at height 1, and one path ends at height 2, which correspond to the fourth diagonal of T . 3 PMAP of the second kind In this section, we focus on PMAP of the second kind. See Figure 5 for an illustration of such paths. The first subsection considers paths in M2, while the second handles paths in M′2. We yield enumerative results for these paths according to the length, the type of the last step, and the height of the end-point. 3.1 PMAP in M2 - From left to right In this part, we consider PMAP in M2, i.e. lattice paths in N2 starting at the origin, consisting of steps U , Dk and H , and where any down step or horizontal step (except for the last step of the path) is immediately followed by an up step. Figure 5 shows two examples of such paths. J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 519 Figure 5: The left drawing shows a MAP of length 18 in M2. The right drawing shows a PMAP of length 18 ending at height 3 in M2. Let P be a length n PMAP in M2 ending at height k ⩾ 0. If the last step of P is U , then k ⩾ 1 and we have P = QU where Q is a length (n−1) PMAP ending at height k−1. So, we obtain the first relation fk = zfk−1 + zgk−1 + zhk−1 for k ⩾ 1, anchored with f0 = 1 by considering the empty path. If the last step of P is a down step Da, a ⩾ 1, then we have P = QDa where Q is a length (n− 1) PMAP ending at height ℓ = a+ k ⩾ k+1 with an up step. So, we obtain the second relation gk = z ∑ ℓ⩾k+1 fℓ. If the last step of P is a horizontal step H , then we have P = QH where Q is a length (n− 1) PMAP ending at height k with an up step, which implies hk = zfk. So, we have to solve the following system of equations. f0 = 1, and fk = zfk−1 + zgk−1 + zhk−1, k ⩾ 1, gk = z ∑ ℓ⩾k+1 fℓ, k ⩾ 0, hk = zfk, k ⩾ 0. (3.1) Summing the recursions in (3.1), we have: F (u) = 1 + z ∑ k⩾1 ukfk−1 + z ∑ k⩾1 ukgk−1 + z ∑ k⩾1 ukhk−1 = 1 + zuF (u) + zuG(u) + zuH(u), G(u) = z ∑ k⩾0 uk ( ∑ ℓ⩾k+1 fℓ ) = z ∑ k⩾1 uk − 1 u− 1 fk = z u− 1 (F (u)− F (1)), H(u) = zF (u). Now, setting a := F (1) and solving these functional equations, we deduce F (u) = au z2 − u+ 1 u2z2 + u2z − uz − u+ 1 , G(u) = − z ( au z2 + auz − a+ 1 ) u2z2 + u2z − uz − u+ 1 , H(u) = z ( au z2 − u+ 1 ) u2z2 + u2z − uz − u+ 1 . In order to compute a = F (1), we use the kernel method on F (u). We can write the denominator (which is a polynomial in u of degree 2), as (z2 + z)(u− r)(u− s) with r = 1 + z + √ −3z2 − 2z + 1 2z (z + 1) , and s = 1 + z − √ −3z2 − 2z + 1 2z (z + 1) . 520 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Replacing u with s (which have a Taylor expansion at z = 0) in order to cancel the numer- ator of F (u), we obtain the equation as z2 − s+ 1 = 0, and thus a = F (1) = s− 1 sz2 . Finally using z(1 + z)rs = 1 and simplifying by the factor (u − s) in the numerators and denominators, we obtain F (u) = r r − u , G(u) = s− 1 sz(r − u) , and H(u) = zr r − u , which implies that [uk]F (u) := fk = 1 rk , [uk]G(u) := gk = (1 + z) · s− 1 rk , [uk]H(u) := hk = z rk . Theorem 3.1. The bivariate generating function for the total number of PMAP with respect to the length and the height of the end-point is given by Total(z, u) = 1 z(r − u) , and we have [uk]Total(z, u) = 1 zrk+1 . Finally, setting t(n, k) = [zn][uk]Total(z, u), we have for n ⩾ 2, k ⩾ 1, t(n, k) = t(n, k − 1) + t(n− 1, k − 1)− t(n− 1, k − 2)− t(n− 2, k − 2), and setting tn := t(n, 0), we have t0 = 1, and for n ⩾ 1, tn = tn−1 + n−2∑ k=1 tktn−1−k. Proof. The first two equalities are immediately deduced from the previous results. The third equality is obtained using the Mathematica package Guess.m ([6]) for guessing recurrence relations. After this, it suffices to check algebraically: Total(z, u) = (u+ uz − u2z − u2z2)Total(z, u)− u(1 + z) + 1 zr . For the last equality, it suffices to remark that the generating function of the sequence (tn)n⩾0, that is 1/(zr), generates a shift of the well known Motzkin sequence A001006 in [13]. J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 521 Let T be the infinite matrix T := [t(n, k)]n⩾0,k⩾0. The first few rows of the matrix T are T =  1 0 0 0 0 0 0 0 · · · 1 1 0 0 0 0 0 0 · · · 1 2 1 0 0 0 0 0 · · · 2 3 3 1 0 0 0 0 · · · 4 6 6 4 1 0 0 0 · · · 9 13 13 10 5 1 0 0 · · · 21 30 30 24 15 6 1 0 · · · 51 72 72 59 40 21 7 1 · · · ... ... ... ... ... ... ... ... . . .  . Corollary 3.2. The generating function that counts the PMAP with respect to the length is given by Total(z, 1) = 1 z(r − 1) . The first few terms of the series expansion of Total(z, 1) are 1 + 2z + 4z2 + 9z3 + 21z4+51z5+127z6+323z7+835z8+2188z9+O(z10), which correspond to a shift of the sequence A001006 in [13] that counts the Motzkin paths of a given length. See Figure 6 for an illustration of the 9 paths of length 3. Corollary 3.3. The generating function that counts the MAP with respect to the length is given by Total(z, 0) = 1 zr . The first few terms of the series expansion of Total(z, 0) are 1+ z+ z2+2z3+4z4+ 9z5+21z6+51z7+127z8+323z9+O(z10) which correspond to a shift of the sequence A001006 in [13] that counts Motzkin paths of a given length. Figure 6: The 9 PMAP of length three in M2. Notice that two paths end on the x-axis, three paths end at height 1, three paths end at height 2, and one path ends at height 3, which correspond to the fourth row of T . 522 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 3.2 PMAP in M′2 - From right to left Here, we consider the paths of the previous section, but we read them from right to left. This means that down steps become up steps and vice versa, and horizontal steps are un- changed, which implies that any up step or horizontal step (except the first step of the path) is preceded by a down step. See Definition 1.4 and Figure 7 for two examples of such paths. Figure 7: The left drawing shows a Motzkin path with air pockets of length 18 in M′2 (read from right to left). The right drawing shows a partial Motzkin path with air pockets of length 18 ending at height 2 in M′2. Let P be a length n PMAP in M′2 ending at height k ⩾ 0. If the last step of P is Ua, a ⩾ 1, then k ⩾ a and we have P = QUa where Q is a length (n− 1) PMAP ending at height ℓ = k−a with a down step. So, we obtain the first relation fk = z+z(g0+g1+. . .+gk−1) for k ⩾ 1, anchored with f0 = 1 by considering the empty path. If the last step of P is a down step D, then we have P = QD where Q is a length (n− 1) PMAP ending at height k+1. So, we obtain the second relation gk = zfk+1+zgk+1+zhk+1. If the last step of P is a horizontal step H , then we have P = QH where Q is a length (n− 1) PMAP ending at height k with a down step whenever it is nonempty. If k ⩾ 1, then we have hk = zgk; if k = 0 then we have h0 = z + zg0 where the monomial z corresponds to the path P = H . So, we have to solve the following system of equations. f0 = 1, and fk = z(1 + g0 + g1 + . . .+ gk−1), k ⩾ 1, gk = zfk+1 + zgk+1 + zhk+1, k ⩾ 0, h0 = z + zg0, and hk = zgk, k ⩾ 1. (3.2) Using the same notations as in the previous sections, and summing the recursions in (3.2), we have: F (u) = 1 + ∑ k⩾1 ukfk = 1 + z ∑ k⩾1 (1 + g0 + . . .+ gk−1)u k = 1 + zu 1− u (1 +G(u)), G(u) = z ∑ k⩾0 (fk+1 + gk+1 + hk+1)u k = z u (F (u)− F (0) +G(u)−G(0) +H(u)−H(0)), H(u) = z + zG(u). Notice that F (0) = 1 and H(0) = z + zG(0) by the third relation. Now, setting J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 523 a := G(0) and solving these functional equations, we deduce F (u) = a uz3 + a uz2 + uz3 − u2z + uz2 + u2 − uz + z2 − u+ z u2 − uz + z2 − u+ z , G(u) = −z (a uz + a u− a z + uz − a) u2 − uz + z2 − u+ z , H(u) = − z ( a uz2 + a uz − a z2 + uz2 − a z − u2 + uz − z2 + u− z ) u2 − uz + z2 − u+ z . In order to compute a = G(0), we use the kernel method on F (u). We can write the denominator (which is a polynomial in u of degree 2), as (u− r)(u− s) with r = 1 + z + √ −3 z2 − 2 z + 1 2 , and s = 1 + z − √ −3 z2 − 2 z + 1 2 . Replacing u with s (which have a Taylor expansion at z = 0) in oder to cancel the numer- ator of F (u), we obtain the equation a sz3 + a sz2 + sz3 − s2z + sz2 + s2 − sz + z2 − s+ z = 0. Using rs = z(1 + z), we deduce a = G(0) = 1− r r . Finally, after simplifying by the factor (u − s) in the numerators and denominators, we obtain F (u) = u(z − 1) + r r − u , G(u) = 1− r r − u , and H(u) = z(1− u) r − u , which implies that [uk]F (u) := fk = 1 rk + [k ̸= 0] · z − 1 rk , [uk]G(u) := gk = 1− r rk+1 , [uk]H(u) := hk = z rk+1 − [k ̸= 0] · z rk . Theorem 3.4. The bivariate generating function for the total number of PMAP with respect to the length and the height of the end-point is given by Total(z, u) = 1− u+ z r − u . [uk]Total(z, u) = z + 1 rk+1 − [k ̸= 0] · 1 rk . Finally, setting t(n, k) = [zn][uk]Total(z, u), we have for n ⩾ 1, k ⩾ 1, t(n, k) = t(n, k − 1)− t(n− 1, k) + t(n− 2, k + 1) + t(n− 1, k + 1), and setting tn := t(n, 0), we have t0 = 1, and for n ⩾ 2, tn = tn−1 + n−2∑ k=1 tktn−1−k. 524 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Proof. The proof are obtained mutatis mutandis as for the previous theorems. Let T be the infinite matrix T := [t(n, k)]n⩾0,k⩾0. The first few rows of the matrix T are T =  1 0 0 0 0 0 0 · · · 1 1 1 1 1 1 1 · · · 1 1 1 1 1 1 1 · · · 2 3 4 5 6 7 8 · · · 4 6 8 10 12 14 16 · · · 9 15 22 30 39 49 60 · · · 21 36 54 75 99 126 156 · · · 51 91 142 205 281 371 476 · · · ... ... ... ... ... ... ... . . .  . Since there is an infinite number of PMAP of length n, we do not provide an ordinary generating function (with respect to the length) for these paths. So, we get around this by counting PMAP ending on a point (x, n− x) for a given n ⩾ 0. Corollary 3.5. The generating function that counts the partial PMAP ending on the line y = n− x is given by Total(z, z) = 1 r − z . The first few terms of the series expansion of Total(z, z) are 1 + z + 2z2 + 4z3 + 9z4+21z5+51z6+127z7+323z8+835z9+O(z10), which correspond to the sequence A001006 in [13] that counts the Motzkin paths with respect to the length. See Figure 8 for the illustration of the 9 PMAP in M′2 ending on the line y = 4− x. Notice that we obviously retrieve the results of Corollary 3.3, i.e., the generating func- tion Total(z, 0) that counts the MAP with respect to the length is also a shift of the Motzkin sequence A001006 in [13]. Figure 8: The 9 PMAP ending on the line y = 4 − x in M′2. Notice that four paths end on the x-axis, three paths end at height 1, one path ends at height 2, and one path ends at height 3, which correspond to the fifth diagonal of T . J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 525 4 A Riordan array point of view In this section, we make links between the previous matrices T = [tn,k]n⩾0,k⩾0 and some Riordan arrays or almost Riordan arrays. We first give a short background on Riordan arrays [4, 5, 12]. An infinite column vector (a0, a1, . . . )T has generating function f(z) if f(z) =∑ n⩾0 anz n. A Riordan array is an infinite lower triangular matrix whose k-th column has generating function g(z)f(z)k for all k ⩾ 0, for some formal power series g(z) and f(z), with g(0) ̸= 0, f(0) = 0, and f ′(0) ̸= 0. Such a Riordan array is denoted by (g(z), f(z)). If we multiply this matrix by a column vector (b0, b1, . . . )T having generat- ing function h(z) = ∑ n⩾0 bnz n, then the resulting column vector has generating function g(z)h(f(z)). This property is known as the fundamental theorem of Riordan arrays. The product of two Riordan arrays (g(z), f(z)) and (h(z), l(z)) is defined by (g(z), f(z)) ∗ (h(z), l(z)) = (g(z)h(f(z)), l(f(z))) . Under the operation “∗”, the set of all Riordan arrays is a group [12]. The identity element is I = (1, z), and the inverse of (g(z), f(z)) is (g(z), f(z))−1 = ( 1/ ( g ◦ f ) (z), f(z) ) , where f(z) denotes the compositional inverse of f(z). Moreover, if a matrix T = [tn,k]n⩾0,k⩾0 is a Riordan array (g(z), f(z)) then tn,k equals the coefficient of znuk in the series expansion of the bivariate generating function g(z) 1− uf(z) , and we say that this is the bivariate generating function of the matrix T . An almost Riordan array T ′ is a matrix that consists of an initial column vector (d0, d1, . . . ) T with generating function g0(z) = ∑ n⩾0 dnz n, followed by a vertically shifted Riordan array (g(z), f(z)) as illustrated below. T ′ =  d0 0 · · · 0 0 · · · d1 d2 d3 (g(z), f(z)) d4 d5 ... ... ... ... ... ...  Therefore, the bivariate generating function for this matrix is given by g0(z) + zu g(z) 1− uf(z) . Finally, we will say that a matrix M = [mn,k]n⩾0,k⩾0 is the rectification of the Riordan array (g(z), f(z)) whenever mn,k equals the coefficient of znuk in the series expansion of 526 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 the bivariate generating function g(z) 1− u f(z)z . In this section, C(z) = 1− √ 1−4z 2z will be the generating function where the coefficient of zn in its series expansion is the Catalan number cn = 1n+1 ( 2n n ) . 4.1 Comment on Section 2.1 Let T = [t(n, k)]n⩾0,k⩾0 be the matrix given in Section 2.1 where t(n, k) is the number of length n PMAP in M1 ending at height k. Proposition 4.1. The matrix T = [t(n, k)]n⩾0,k⩾0 is a Riordan array defined by ( 1 1− z2 C ( z(1− z + z2) (1− z2)2 ) , z 1− z2 C ( z(1− z + z2) (1− z2)2 )) . Proof. Considering r defined in Section 2.1, we have [uk]Total(z, u) = 1 zrk+1 = ( 1 zr )( 1 r )k = 2 1− z2 + √ 1− 4z + 2z2 − 4z3 + z4 ( 2z 1− z2 + √ 1− 4z + 2z2 − 4z3 + z4 )k . Therefore, the array T satisfies T = ( 2 1− z2 + √ 1− 4z + 2z2 − 4z3 + z4 , 2z 1− z2 + √ 1− 4z + 2z2 − 4z3 + z4 ) = ( 1− z2 − √ 1− 4z + 2z2 − 4z3 + z4 2z(1− z + z2) , 1− z2 − √ 1− 4z + 2z2 − 4z3 + z4 2(1− z + z2) ) = ( 1 1− z2 C ( z(1− z + z2) (1− z2)2 ) , z 1− z2 C ( z(1− z + z2) (1− z2)2 )) . Proposition 4.2. The general term t(n, k) equals n−k∑ i=0 ( k + n−k−i2 n−k−i 2 ) 1 + (−1)n−k−i 2 i∑ j=0 Ck+j,k j∑ m=0 ( j m ) (−1)m m∑ p=0 ( m p ) (−1)p ( 2j − 1 + i−j−m−p2 i−j−m−p 2 ) 1 + (−1)i−j−m−p 2 , where Cn,k = k+1n+1 ( 2n−k n−k ) is the general term of the Catalan matrix (A033184 in [13]). J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 527 Proof. Setting Z = z(1−z+z 2) (1−z2)2 , we have t(n, k) = [zn]zk 1 (1− z2)k+1 C(Z)k+1 = [zn−k] 1 (1− z2)k+1 C(Z)k+1 = n−k∑ i=0 [zn−k−i] 1 (1− z2)k+1 [zi]C(Z)k+1 (product rule [8]) = n−k∑ i=0 [zn−k−i] 1 (1− z2)k+1 i∑ j=0 [zj ]C(z)k+1[zi]Zj (composition rule [8]) = n−k∑ i=0 [zn−k−i] 1 (1− z2)k+1 i∑ j=0 [zj ] 1 zk C(z)(zC(z))k[zi]Zj = n−k∑ i=0 ( k + n−k−i2 n−k−i 2 ) 1 + (−1)n−k−i 2 i∑ j=0 Ck+j,k[z i]Zj (see [8]). Since we have [zi]Zj = [zi] ( z(1− z + z2) (1− z2)2 )j = j∑ m=0 ( j m ) (−1)m m∑ p=0 ( m p ) (−1)p ( 2j − 1 + i−j−m−p2 i−j−m−p 2 ) 1 + (−1)i−j−m−p 2 , the result follows. 4.2 Comment on Section 2.2 Let T = [t(n, k)]n⩾0,k⩾0 be the matrix given in Section 2.2 where t(n, k) is the number of length n PMAP in M′1 ending at height k. Proposition 4.3. The matrix T = [t(n, k)]n⩾0,k⩾0 can be written T =  1 0 0 0 0 0 · · · 1 1 1 1 1 1 · · · 2 3 3 3 3 3 · · · 5 8 10 12 14 16 · · · 13 23 33 43 53 63 · · · 36 69 107 149 195 245 · · · ... ... ... ... ... ... . . .  = A ·B 528 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 where A =  1 0 0 0 0 0 · · · 1 1 0 0 0 0 · · · 2 3 0 0 0 0 · · · 5 8 2 0 0 0 · · · 13 23 10 0 0 0 · · · 36 69 38 4 0 0 · · · ... ... ... ... ... ... . . .  and B =  1 0 0 0 0 0 · · · 0 1 1 1 1 1 · · · 0 0 1 2 3 4 · · · 0 0 0 1 3 6 · · · 0 0 0 0 1 4 · · · 0 0 0 0 0 1 · · · ... ... ... ... ... ... . . .  are defined as follows: • The matrix B = [bn,k]n,k⩾0 is defined by b0,0 = 1, and bn,0 = b0,n = 0 if n ⩾ 1, and bn,k = ( k−1 n−1 ) otherwise, which is a kind of Pascal matrix. • The matrix A = [an,k]n,k⩾0 is the almost Riordan array with initial column of gen- erating function g0(z) = 1− z2 − √ 1− 4z + 2z2 − 4z3 + z4 2z(1− z + z2) , followed by the shifted Riordan array( 1− 3z + z2 − z3 − (1− z) √ 1− 4z + 2z2 − 4z3 + z4 2z3(1− z + z2) , 1− 2z − z2 − √ 1− 4z + 2z2 − 4z3 + z4 2z ) . The second column of T 1, 3, 8, 23, 69, . . . with generating function 1− 3z + z2 − z3 − (1− z) √ 1− 4z + 2z2 − 4z3 + z4 2z3(1− z + z2) is the convolution of the first column of T 1, 1, 2, 5, 13, 36, . . . (A114465 in [13]) and the sequence 1, 2, 4, 10, 28, . . . (A187256 in [13]). Proof. An almost Riordan array is represented by an initial column vector with generating function g0(z), followed by a vertically shifted Riordan array (g(z), f(z)). The bivariate generating function of this matrix is then given by g0(z) + zu g(z) 1−uf(z) . In our case, for the almost Riordan array A, we have g0(z) = 1− z2 − √ 1− 4z + 2z2 − 4z3 + z4 2z(1− z + z2) , g(z) = 1− 3z + z2 − z3 − (1− z) √ 1− 4z + 2z2 − 4z3 + z4 2z3(1− z + z2) , f(z) = 1− 2z − z2 − √ 1− 4z + 2z2 − 4z3 + z4 2z . J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 529 We let G(z, u) = g0(z) + zu g(z) 1−uf(z) , the bivariate generating function of the almost Rior- dan array. Now, it suffices to check that the generating function corresponding to the matrix A ·B, that is G(z, u 1− u ) = (1− u+ uz)(1− 2z + 2uz − (1 + 2u)z2 − √ 1− 4z + 2z2 − 4z3 + z4 2(1− z + z2)(u(u− 1) + (1− u2)z + u(u+ 1)z2) , coincides with the generating function Total(z, u) of the matrix T . Proposition 4.4. The matrix [t(n, k)]n⩾1,k⩾1 =  1 1 1 1 1 · · · 3 3 3 3 3 · · · 8 10 12 14 16 · · · 23 33 43 53 63 · · · 69 107 149 195 245 · · · ... ... ... ... ... . . .  is the rectification of the Riordan array (g(z), h(z)) with h(z) = 1− z2 − √ 1− 4z + 2z2 − 4z3 + z4 2 , and g(z) = 1− 3z + z2 − z3 − (1− z) √ 1− 4z + 2z2 − 4z3 + z4 2z3(1− z + z2) . Proof. It suffices to check that the generating function of T , i.e. Total(z, u), given in Theorem 2.4, equals to Total(z, 0) + zu g(z) 1− uh(z)z . We can express h(z) and g(z), respectively, in the following form h(z) = z(1− z + z2) 1− z2 C ( z(1− z + z2) (1− z2)2 ) , g(z) = 1 1− 3z + z2 − z3 C ( z3(1− z + z2) (1− 3z + z2 − z3)2 ) . Then g(z) expands to give the first column 1, 3, 8, 23, . . ., whose n-th term vn can be expressed vn = n∑ k=0 k∑ j=0 ( k j ) (−1)j j∑ i=0 ( j i ) (−1)i n−3k−j−i∑ ℓ=0 ( 2k + ℓ ℓ ) ℓ∑ m=0 ( ℓ m ) 3ℓ−m(−1)m ( m n− 3k − j − i− ℓ−m ) (−1)n−3k−j−i−ℓ−mck. Using vn, we can deduce the following. 530 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 Proposition 4.5. The general term t(n, k) equals n+k∑ i=0 vn+k−i i∑ j=0 j∑ m=0 Mm,k m∑ p=0 ( m p ) (−1)p p∑ q=0 ( p q ) (−1)q ( 2m− 1 + j−m−p−q2 j−m−p−q 2 ) 1 + (−1)j−m−p−q 2 ( k i−j 2 ) (−1) i−j 2 1 + (−1)i−j 2 , where Mn,k = { [k = 0] if n = 0, n k ( 2n−k−1 n−k ) otherwise, is the general term of Riordan array (1, zC(z)) (see A106566 in [13]). 4.3 Comment on Section 3.1 Let T = [t(n, k)]n⩾0,k⩾0 be the matrix given in Section 3.1 where t(n, k) is the number of length n PMAP in M2 ending at height k. Proposition 4.6. The matrix T = [t(n, k)]n⩾0,k⩾0 is the Riordan array T = (1 + zM(z), z(1 + zM(z))) = ( C ( z 1 + z ) , zC ( z 1 + z )) , where M(z) = 1−z− √ 1−2z−3z2 2z2 is the generating function of the Motzkin numbers (see A001006 in [13]). Proof. It suffices to check that Total(z, u) given in Theorem 3.1 satisfies Total(z, u) = C ( z 1+z ) 1− uzC ( z 1+z ) , and that 1 + zM(z) = C ( z 1+z ) . This triangle T corresponds to A091836 in [13] where the coefficient of row n− 1 and column k is the number of Motzkin paths of length n having k points on the horizontal axis (besides the first and last point). As mentioned in [13], we obtain t(n, k) =  1, if n = k, k+1 n+1 n−k∑ j=1 j(−1)n−k−j ( n+j j ) n−k∑ i=0 1 n−k ( i n−k−i+j )( n−k i ) , otherwise. A second expression for t(n, k) is given by the following proposition. Proposition 4.7. The general term t(n, k) of the Riordan array (1+zM(z), z(1+zM(z))) is given by t(n, k) = { 1 if n = k, k+1 n−k ∑k j=0 ( k j )∑n−k i=0 ( n−k i )( i n−k−i−j−1 ) otherwise. J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 531 Proof. We prove this using Lagrange inversion, using the fact that z 1 + z + z2 M ( z 1 + z + z2 ) = z, which means that the compositional inverse (zM(z))−1 of zM(z) is (zM(z))−1 = z 1 + z + z2 . Thus we have t(n, k) = [zn](1 + zM(z))(z(1 + zM(z)))k = [zn−k](1 + zM(z))k+1 = [zn−k]G(zM(z)), with G(z) = (1 + z)k+1 = 1 n− k [zn−k−1]G′(z) ( z (zM(z))−1 )n−k (Lagrange inversion) = 1 n− k [zn−k−1](k + 1)(1 + z)k(1 + z + z2)n−k = k + 1 n− k [zn−k] k∑ j=0 ( k j ) zj n−k∑ i=0 ( n− k i ) zi(1 + z)i = k + 1 n− k [zn−k] k∑ j=0 ( k j ) zj n−k∑ i=0 ( n− k i ) zi i∑ ℓ=0 ( i ℓ ) zℓ = k + 1 n− k k∑ j=0 ( k j ) n−k∑ i=0 ( n− k i )( i n− k − i− j − 1 ) . Remark 4.8. The Riordan array (1 + zM(z), z(1 + zM(z))) is a pseudo-involution in the Riordan group (see [5, Example 8]), that is, the matrix [(−1)ktn,k]n,k⩾0 is idempotent. Thus, this work yields a significant lattice path interpretation of this array. 4.4 Comment on Section 3.2 Let T = [t(n, k)]n⩾0,k⩾0 be the matrix given in Section 3.2 where t(n, k) is the number of length n PMAP in M′2 ending at height k. Proposition 4.9. The matrix T = [t(n, k)]n⩾0,k⩾0 can be written T =  1 0 0 0 0 0 · · · 1 1 1 1 1 1 · · · 1 1 1 1 1 1 · · · 2 3 4 5 6 7 · · · 4 6 8 10 12 14 · · · 9 15 22 30 39 49 · · · ... ... ... ... ... ... . . .  = A ·B 532 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 where A =  1 0 0 0 0 0 · · · 1 1 0 0 0 0 · · · 1 1 0 0 0 0 · · · 2 3 1 0 0 0 · · · 4 6 2 0 0 0 · · · 9 15 7 1 0 0 · · · ... ... ... ... ... ... . . .  and B =  1 0 0 0 0 0 · · · 0 1 1 1 1 1 · · · 0 0 1 2 3 4 · · · 0 0 0 1 3 6 · · · 0 0 0 0 1 4 · · · 0 0 0 0 0 1 · · · ... ... ... ... ... ... . . .  are defined as follows: • The matrix B = [bn,k]n,k⩾0 is defined by b0,0 = 1, and bn,0 = b0,n = 0 if n ⩾ 1, and bn,k = ( k−1 n−1 ) otherwise, which is the same as in Proposition 4.3. • The matrix A = [an,k]n,k⩾0 is the almost Riordan array with initial column whose generating function is g0(z) = 1 + zM(z) = 1 + z − √ 1− 2z − 3z2 2z , which is followed by the shifted Riordan array (g(z), z2g(z)) where g(z) = 1− z − 2z2 − √ 1− 2z − 3z2 2z3(1 + z) . Proof. The almost Riordan array A has generating function g0(z) + zu g(z) 1− z2ug(z) . Setting G(z, u) = g0(z) + zu g(z) 1−z2ug(z) , it suffices to check that the generating function corresponding to the matrix A·B, that is G(z, u1−u ), coincides with the generating function Total(z, u) of the matrix T given in Theorem 3.4. Proposition 4.10. The matrix [t(n, k)]n⩾1,k⩾1 =  1 1 1 1 1 · · · 1 1 1 1 1 · · · 3 4 5 6 7 · · · 6 8 10 12 14 · · · 15 22 30 39 49 · · · 36 54 75 99 126 · · · ... ... ... ... ... . . .  is the rectification of the Riordan array (M(z), zR(z)) where M(z) = 1−z− √ 1−2z−3z2 2z2 is the generating function of the Motzkin numbers, and R(z) = 1+z− √ 1−2z−3z2 2z(1+z) is the generating function of the Riordan numbers (A005043 in [13]). J.-L. Baril and P. Barry: Two kinds of partial Motzkin paths with air pockets 533 Proof. It suffices to check that the generating function of T , i.e. Total(z, u), given in Theorem 3.4, equals to g0(z) + zu M(z) 1−u zR(z)z . We let mn denote the n-th Motzkin number mn = ∑⌊n2 ⌋ k=0 ( n 2k ) ck where ck is the k-th Catalan defined above. Corollary 4.11. We have t(n, k) = { [k = 0] if n = 0, r(n− 1, k) otherwise, where r(n, k) = n∑ i=0 mi · (k + [n = i]) n+ k − i+ [n = i] n−i∑ j=0 (−1)n−i−j ( n+ k − i+ j − 1 j ) n+k−i∑ ℓ=0 ( n+ k − i ℓ )( ℓ n− i− j − ℓ ) . Proof. We have (M(z), zR(z))−1 = ( (1−x)2 1−x+x2 , x(1−x) 1−x+x2 ) . If we denote by (v(z), u(z)) this inverse Riordan array, then we obtain (M(z), zR(z)) = ( 1 v(ū(z)) , ū(z) ) , where ū is the compositional inverse [8] of u. Using the definition of a Riordan array, and Lagrange inversion, we find that the Riordan array (M(z), zR(z)) has general term r̃(n, k) given by n∑ i=0 mi · (k + [n = k + i]) n− i+ [n = k + i] n−k−i∑ j=0 (−1)n−k−i−j ( n− i+ j − 1 j ) n−i∑ ℓ=0 ( n− i ℓ )( ℓ n− k − i− j − ℓ ) . This array begins as follows: 1 0 0 0 0 0 · · · 1 1 0 0 0 0 · · · 2 1 1 0 0 0 · · · 4 3 1 1 0 0 · · · 9 6 4 1 1 0 · · · 21 15 8 5 1 1 · · · ... ... ... ... ... ... . . .  . To rectify this array and thus to obtain the array (M(z), R(z)), we change n to n+ k, and we obtain r(n, k) above. To this array we must now prepend the row (1, 0, 0, 0, . . .), and the result follows. ORCID iDs Jean-Luc Baril https://orcid.org/0000-0002-9811-8028 Paul Barry https://orcid.org/0000-0003-1909-8725 534 Ars Math. Contemp. 24 (2024) #P3.08 / 509–534 References [1] C. Banderier, M. Bousquet-Mélou, A. Denise, P. Flajolet, D. Gardy and D. Gouyou- Beauchamps, Generating functions for generating trees, Discrete Math. 246 (2002), 29–55, doi: 10.1016/S0012-365X(01)00250-3, https://doi.org/10.1016/S0012-365X(01) 00250-3. [2] J.-L. Baril, S. Kirgizov, R. Maréchal and V. Vajnovszki, Enumeration of Dyck paths with air pockets, J. Integer Seq. 26 (2023), Article 23.3.2, 23 pp., https://cs.uwaterloo.ca/ journals/JIS/VOL26/Kirgizov/kirg5.html. [3] J.-L. Baril, S. Kirgizov, R. Maréchal and V. Vajnovszki, Grand Dyck paths with air pockets, Art Discrete Appl. Math. 7 (2024), Article #P1.07, doi:10.26493/2590-9770.1587.b2a, https: //doi.org/10.26493/2590-9770.1587.b2a. [4] P. Barry, Riordan Arrays: A Primer, Logic Press, Kilcock, Co. Kildare, 2017. [5] A. Burstein and L. W. Shapiro, Pseudo-involutions in the Riordan group, J. Integer Seq. 25 (2022), Article 22.3.6, 54 pp., https://cs.uwaterloo.ca/journals/JIS/VOL25/ Burstein/burstein14.html. [6] M. Kauers, Guessing Handbook, Version 0.32, 2009, Technical Report 09-07, RISC-Linz, https://www3.risc.jku.at/research/combinat/software/ergosum/ RISC/Guess.html. [7] A. Krinik, G. Rubino, D. Marcus, R. J. Swift, H. Kasfy and H. Lam, Dual processes to solve single server systems, J. Stat. Plann. Inference 135 (2005), 121–147, doi:10.1016/j.jspi.2005. 02.010, https://doi.org/10.1016/j.jspi.2005.02.010. [8] D. Merlini, R. Sprugnoli and M. C. Verri, The method of coefficients, Am. Math. Monthly 114 (2007), 40–57, doi:10.1080/00029890.2007.11920390, https://doi.org/10.1080/ 00029890.2007.11920390. [9] H. Prodinger, The kernel method: a collection of examples, Sémin. Lothar. Comb. 50 (2004), Article B50f, 19 pp., https://www.mat.univie.ac.at/%7Eslc/ wpapers/s50proding.html. [10] H. Prodinger, Partial Dyck paths with air pockets, Integers 22 (2022), Paper A94, 8 pp., doi:10. 5281/zenodo.10997311, http://math.colgate.edu/˜integers/vol22.html. [11] E. Rowland and R. Yassawi, Automatic congruences for diagonals of rational functions, J. Théor. Nombres Bordeaux 27 (2015), 245–288, doi:10.5802/jtnb.901, https://doi.org/ 10.5802/jtnb.901. [12] L. W. Shapiro, S. Getu, W. J. Woan and L. C. Woodson, The Riordan group, Discrete Appl. Math. 34 (1991), 229–239, doi:10.1016/0166-218X(91)90088-E, https://doi.org/10. 1016/0166-218X(91)90088-E. [13] N. J. A. Sloane et al., The On-Line Encyclopedia of Integer Sequences, https://oeis.org. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.09 / 535–551 https://doi.org/10.26493/1855-3974.3113.6f7 (Also available at http://amc-journal.eu) Independent coalition in graphs: existence and characterization Mohammad Reza Samadzadeh , Doost Ali Mojdeh * Department of Mathematics, Faculty of Mathematical Sciences, University of Mazandaran, Babolsar, Iran Received 2 May 2023, accepted 3 November 2023, published online 23 August 2024 Abstract An independent coalition in a graph G consists of two disjoint sets of vertices V1 and V2 neither of which is an independent dominating set but whose union V1 ∪ V2 is an inde- pendent dominating set. An independent coalition partition, abbreviated, ic-partition, in a graph G is a vertex partition π = {V1, V2, . . . , Vk} such that each set Vi of π either is a singleton dominating set, or is not an independent dominating set but forms an independent coalition with another set Vj ∈ π. The maximum number of classes of an ic-partition of G is the independent coalition number of G, denoted by IC(G). In this paper, we study the concept of ic-partition. In particular, we discuss the possibility of the existence of ic- partitions in graphs and introduce a family of graphs for which no ic-partition exists. We also determine the independent coalition number of some classes of graphs and investi- gate graphs G of order n with IC(G) ∈ {1, 2, 3, 4, n} and the trees T of order n with IC(T ) = n− 1. Keywords: Independent coalition, independent coalition partition, independent dominating set, ido- matic partition. Math. Subj. Class. (2020): 05C69 1 Introduction Let G = (V,E) denote a simple graph of order n with vertex set V = V (G) and edge set E = E(G). The open neighborhood of a vertex v ∈ V is the set N(v) = {u|{u, v} ∈ E}, and its closed neighborhood is the set N [v] = N(v)∪{v}. Each vertex of N(v) is called a neighbor of v, and the cardinality of N(v) is called the degree of v, denoted by deg(v) or *Corresponding author. E-mail addresses: m.samadzadeh02@umail.umz.ac.ir (Mohammad Reza Samadzadeh), damojdeh@umz.ac.ir (Doost Ali Mojdeh) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 536 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 degG(v). A vertex v of degree 1 is called a pendant vertex or leaf, and its neighbor is called a support vertex. A vertex of degree n− 1 is called a full vertex while a vertex of degree 0 is called an isolated vertex. The minimum and maximum degree of G are denoted by δ(G) and ∆(G), respectively. For a set S of vertices of G, the subgraph induced by S is denoted by G[S]. For two sets X and Y of vertices, let [X,Y ] denote the set of edges between X and Y . If every vertex of X is adjacent to every vertex of Y , we say that [X,Y ] is full, while if there are no edges between them, we say that [X,Y ] is empty. A subset Vi ⊆ V is called a singleton set if |Vi| = 1, and is called a non-singleton set if |Vi| ≥ 2. The join G+H of two disjoint graphs G and H is the graph obtained from the union of G and H by adding every possible edge between the vertices of G and the vertices of H . We denote the family of paths, cycles, complete graphs and stars of order n by Pn, Cn, Kn and K1,n−1, respectively, and the complete k-partite graph with partite sets of order n1, n2, . . . , nk, by Kn1,...,nk . A double star with respectively p and q leaves connected to each support vertex is denoted by Sp,q . The complete graph K3 is called a triangle, and a graph is triangle-free if it has no K3 as an induced subgraph. The girth of a graph with a cycle is the length of its shortest cycle. For a graph G, the girth of G is denoted by g(G). For a graph G of order n, let G denote the complement of G with V (G) = V (G) and E(G) = E(Kn)−E(G) [13]. A set S ⊆ V in a graph G = (V,E) is called a dominating set if every vertex v ∈ V is either an element of S or is adjacent to an element of S. A set S ⊆ V is called an independent set if its vertices are pairwise nonadjacent. The vertex independence number, denoted by α(G), is the maximum cardinality of an independent set of G. An independent dominating set in a graph G is a set which is both independent and dominating. A partition of the vertices of G into dominating sets (independent dominating sets) is called a domatic partition (idomatic partition). The maximum number of classes of a domatic partition (idomatic partition) of G is called the domatic number (idomatic number) of G, denoted by d(G) (id(G)). The concepts of domination and domatic partition and their variations have been studied widely in the literature. See, for example, [1, 2, 3, 4, 5, 6, 12]. The term coalition was introduced by Haynes et al, [7] and has been studied further in [8, 9, 10, 11]. Definition 1.1 ([7]). A coalition in a graph G consists of two disjoint sets of vertices V1, V2 ⊂ V , neither of which is a dominating set but whose union V1 ∪ V2 is a dominating set. We say that the sets V1 and V2 form a coalition, and are coalition partners. Definition 1.2 ([7]). A coalition partition, henceforth called a c-partition, in a graph G is a vertex partition π = {V1, V2, . . . , Vk} such that every set Vi of π is either a singleton dominating set, or is not a dominating set but forms a coalition with another set Vj in π. The coalition number C(G) equals the maximum order k of a c-partition of G, and a c-partition of G having order C(G) is called a C(G)-partition. Herein we will focus on coalitions involving independent dominating sets in graphs. In other words, we will study the concepts of independent coalition and independent coalition partition which have been introduced in [7] as an area for future research. We begin with the following definitions. Definition 1.3. An independent coalition in a graph G consists of two disjoint sets of independent vertices V1 and V2, neither of which is an independent dominating set but whose union V1 ∪ V2 is an independent dominating set. We say the sets V1 and V2 form an independent coalition, and are independent coalition partners (or ic-partners). M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 537 Definition 1.4. An independent coalition partition, abbreviated ic-partition, in a graph G is a vertex partition π = {V1, V2, . . . , Vk} such that every set Vi of π is either a singleton dominating set, or is not an independent dominating set but forms an independent coalition with another set Vj ∈ π. The independent coalition number IC(G) equals the maximum number of classes of an ic-partition of G, and an ic-partition of G having order IC(G) is called an IC(G)-partition. Definition 1.5 ([8]). Let G be a graph of order n with vertex set V = {v1, v2, . . . , vn}. The singleton partition, denoted π1, of G is the partition of V into n singleton sets, that is, π1 = {{v1}, {v2}, . . . , {vn}}. This paper is organized as follows. Section 1 is devoted to terminology and definitions. We discuss the possibility of the existence of ic-partitions in graphs and derive some bounds on independent coalition number in Section 2. In Section 3, we determine the independent coalition number of some classes of graphs. The graphs G with IC(G) ∈ {1, 2, 3, 4} are investigated in Section 4. In Section 5, we characterize triangle-free graphs G with IC(G) = n and trees T with IC(T ) = n − 1. Finally, we end the paper with some research problems. 2 Independent coalition partition: existence and bound This section is divided into two subsections. In the first subsection, we show that not all graphs admit an ic-partition, and in the second subsection, we present some bounds on IC(G) whenever the graph G admits an ic-partition. 2.1 Existence In the following definition, we construct graphs with arbitrarily large order for which no ic-partition exists. Definition 2.1. Let B be the set of all graphs obtained from the complete graph Kn, (n ≥ 4) with the vertices vi, (1 ≤ i ≤ n), and two additional vertices vn+1, vn+2 such that vn+1 and vn+2 are adjacent to vn, and vn+1 is adjacent to vn−1. Figure 1 illustrates such a graph for n = 4. v1 v2 v3 v4 v5 v6 Figure 1: The graph G in B for n = 4. Proposition 2.2. Let G be a graph. If G ∈ B, then G has no ic-partition. Proof. Suppose, to the contrary, that G has an ic-partition π. The vertices v1, v2, . . . , vn−1 are pairwise adjacent, so they must be in different classes. Further, vn is a full vertex, so it must be in a singleton class. Since vn−1 is adjacent to all vertices except vn+2, and 538 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 {vn−1, vn+2} dominates G, it follows that {vn−1} ∈ π. Further, since {vn−1} can only form an independent coalition with {vn+2}, it follows that {vn+2} ∈ π. If {vn+1} ∈ π, then π is a singleton partition. In this case, {vn+1} has no ic-partner, a contradiction. Hence, {vn+1} /∈ π. It follows that π consists of a non-singleton set {vn+1, vi} such that vi ∈ {v1, v2, . . . , vn−2}, and n singleton sets. Assume, without loss of generality, that {vn+1, v1} ∈ π. Now for each 2 ≤ i ≤ n − 2, the set {vi} has no ic-partner, a contradiction. 2.2 Bounds Definition 1.4 implies that an ic-partition of a graph G is also a c-partition. Further, we note that an ic-partition of G is a proper coloring as well. Hence, we have the following two sharp bounds on IC(G). To see the sharpness of them, consider the complete graph Kn. Observation 2.3. Let G be a graph. If G has an ic-partition, then IC(G) ≤ C(G). Furthermore, this bound is sharp. Observation 2.4. Let G be a graph. If G has an ic-partition, then IC(G) ≥ χ(G). Fur- thermore, this bound is sharp. Given a connected graph G and an ic-partition π of it, the following theorem shows that each set in π admits at most ∆(G) ic-partners. Theorem 2.5. Let G be a connected graph with maximum degree ∆(G), and let π be an ic- partition of G. If X ∈ π, then X is in at most ∆(G) independent coalitions. Furthermore, this bound is sharp. Proof. Let π be an ic-partition of G, and let X be a set in π. If X is a dominating set, then it has no ic-partner. Hence, we may assume that X does not dominate G. Let x be a vertex that is not dominated by X . Now every ic-partner of X must dominate x, that is, it must contain a vertex in N [x]. Hence, there are at most |N [x]| ≤ ∆(G) + 1 sets in π that can form an independent coalition with X . Now we show that X cannot form an independent coalition with ∆(G) + 1 sets. Suppose, to the contrary, that X has ∆(G) + 1 ic-partners (name V1, V2, . . . , V∆+1). Consequently, [X,Vi] is empty for each 1 ≤ i ≤ ∆(G) + 1. Let U = ⋃∆(G)+1 i=1 Vi, and G ′ = G[U ]. Consider an arbitrary vertex v ∈ U (say v ∈ Vi) and an arbitrary set Vj such that 1 ≤ j ≤ ∆(G) + 1 and j ̸= i. Since X ∪ Vj dominates G and [X,Vi] is empty, it follows that v has a neighbor in Vj . Choosing Vj arbitrarily, we conclude that degG′(v) ≥ ∆(G), and so degG′(v) = ∆(G). Hence, for each v ∈ U , we have degG′(v) = ∆(G). Now since G is connected, there is a path P = (v0, v1, . . . , vk) connecting U to X such that v0 ∈ U and vk ∈ X . Note that [U,X] is empty, and so V (P ) \ (U ∪ X) ̸= ∅. Let i be the smallest index for which vi /∈ U ∪ X . It follows that vi−1 ∈ U , and so degG′(vi−1) = ∆(G). Thus, we have degG(vi−1) ≥ degG′(vi−1) + 1 = ∆(G) + 1, a contradiction. To prove the sharpness, let G be the graph that is obtained from the complete graph Kn with vertices vi, (1 ≤ i ≤ n), and a path P2 = (a, b), where b is adjacent to v1. Let A = {a}, B = {b} and Vi = {vi}, for 1 ≤ i ≤ n. One can observe that ∆(G) = n and that the singleton partition π1 = {V1, V2, . . . , Vn, A,B} is an ic-partition of G such that A forms an independent coalition with Vi, for each 1 ≤ i ≤ n. This completes the proof. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 539 Note that the bound presented in Theorem 2.5 does not hold for disconnected graphs. As a counterexample, consider the graph G = K2 ∪ K2 and the singleton partition π1 of it. On can verify that π1 is an ic-partition of G such that each set in π1 has two ic-partners, while ∆(G) = 1. The next bound relates independent coalition number of a graph to its idomatic number. As we will see in the proof of Theorem 2.6, any graph admitting an idomatic partition has an ic-partition. However, the converse is not necessarily true. For example, the singleton partition of the cycle C5 is an ic-partition of it, while C5 has no idomatic partition. Or the cycle C7 has the ic-partition π = {{v1, v5}, {v2, v4}, {v3}, {v6}, {v7}}, while it has no idomatic partition. Theorem 2.6. Let G be a connected graph, and let r ≥ 0 be the number of full vertices of G. If G admits an idomatic partition, then IC(G) ≥ 2id(G)− r. Proof. Let F = {v1, v2, . . . , vr} be the set of full vertices of G, and let π = {V1, V2, . . . , Vid(G)} be an idomatic partition of G of order id(G). Note that each full vertex must be in a singleton set of π. Without loss of generality, assume that vi ∈ Vi, for each 1 ≤ i ≤ r. It follows that for each r + 1 ≤ i ≤ k, we have |Vi| ≥ 2. Now for each r + 1 ≤ i ≤ k, we partition Vi into two nonempty subsets Vi,1 and Vi,2. Note that no proper subset of Vi is a dominating set. Thus, neither Vi,1 nor Vi,2 is an in- dependent dominating set, and so Vi,1 and Vi,2 are ic-partners. It follows that the par- tition π′ = {V1, V2, . . . , Vr, Vr+1,1, Vr+1,2, Vr+2,1, Vr+2,2, . . . , Vid(G),1, Vid(G),2} is an ic-partition of G of order 2id(G)− r. Hence, IC(G) ≥ 2id(G)− r. 3 Independent coalition number for some classes of graphs Let us begin this section with some routine results. Observation 3.1. For n ≥ 1, we have IC(Kn) = n. Observation 3.2. For n ≥ 3, we have IC(K1,n−1) = 3. Observation 3.3. For p, q ≥ 1, we have IC(Sp,q) = 4. For complete multipartite graphs, the following result is obtained. Proposition 3.4. Let G = Kn1,n2,...,nk be a complete k-partite graph with m ≥ 0 full vertices (m partite sets of cardinality 1). Then IC(G) = 2k −m. Proof. Let π = {V1, V2, . . . , Vk} be the partition of G into its partite sets. Assume, without loss of generality, that the sets Vi, for 1 ≤ i ≤ m, are those containing full vertices. Now for each m+ 1 ≤ i ≤ k, we partition Vi into two sets Vi,1 and Vi,2. Observe that Vi,1 and Vi,2 are ic-partners, and so the partition π′ = {{V1}, {V2}, . . . , {Vm}, {Vm+1,1, Vm+1,2}, {Vm+2,1, Vm+2,2}, . . . , {Vk,1, Vk,2}} is an ic-partition of G of order 2k − m. Thus, IC(G) ≥ 2k −m. Now let π′′ be an ic-partition of G. We note that π′′ has the following properties: • For any set S ∈ π′′, all vertices in S are in the same partite set of G. • For any set Vi ∈ π, the vertices in Vi are in at most two sets of π′′. Hence, we have IC(G) ≤ 2k −m, and so IC(G) = 2k −m. 540 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 Next we determine the independent coalition number of all paths and cycles. Lemma 3.5 ([7]). For any path Pn, C(Pn) ≤ 6. Theorem 3.6. For the path Pn, IC(Pn) =  n if n ≤ 4; 4 if n = 5; 5 if n = 6, 7, 8, 9; 6 if n ≥ 10. Proof. It is clear that for 1 ≤ n ≤ 4, we have IC(Pn) = n. Now let n = 5. Con- sider the path P5 with V (P5) = {v1, v2, v3, v4, v5}. It is easily seen that IC(P5) ̸= 5. Thus, IC(P5) ≤ 4. The partition {{v1, v3}, {v2}, {v4}, {v5}} is an ic-partition of P5, so IC(P5) = 4. Now assume n = 6. Consider the path P6 with V (P6) = {v1, v2, v3, v4, v5, v6}. It is clear that IC(P6) ̸= 6. The partition {{v1, v6}, {v2}, {v3}, {v4}, {v5}} is an ic-partition of P6, so IC(P6) = 5. Next assume n = 7. Consider the path P7 with V (P7) = {v1, v2, v3, v4, v5, v6, v7}. By Lemma 3.5 and Observation 2.3, we have IC(Pn) ≤ 6. Now we show that IC(P7) ̸= 6. Suppose that IC(P7) = 6. Let π be an IC(P7)-partition. We note that π consists of a set (name A) of cardinality 2 and five sin- gleton sets. Since γi(P7) = 3, each singleton set must be an ic-partner of A. On the other hand, Theorem 2.5 implies that A has at most two ic-partners, a contradiction. The parti- tion {{v1, v6}, {v2, v7}, {v3}, {v4}, {v5}} is an ic-partition of P7. Therefore, IC(P7) = 5. Next we assume n = 8. Consider the path P8 with V (P8) = {v1, v2, v3, v4, v5, v6, v7, v8}. By Lemma 3.5 and Observation 2.3, we have IC(Pn) ≤ 6. Now we show that IC(P8) ̸= 6. Suppose that IC(P8) = 6. Let π be an IC(P8)-partition. We consider two cases. Case 1. π consists of a set (name A) of cardinality 3 and five singleton sets. Since γi(P8) = 3, each singleton set must be an ic-partner of A. On the other hand, Theorem 2.5 implies that A has at most two ic-partners, a contradiction. Case 2. π consists of two sets of cardinality 2 and four singleton sets. Since γi(P8) = 3, each singleton set must be an ic-partner of a set of cardinality 2. Therefore, using Theorem 2.5, we deduce that for any two ic-partners C and D, it holds that |C∪D| = 3. On the other hand, v3 and v6 are not present in any independent dominating set of cardinality 3, a contradiction. The partition {{v1, v3, v6}, {v2, v7}, {v8}, {v4}, {v5}} is an ic-partition of P8. Therefore, IC(P8) = 5. Now let n = 9. Consider the path P9 with V (P9) = {v1, v2, v3, v4, v5, v6, v7, v8, v9}. By Lemma 3.5 and Observation 2.3, we have IC(Pn) ≤ 6. Now we show that IC(P9) ̸= 6. Suppose that IC(P9) = 6. Let π be an IC(P9)-partition. There exist three cases. Case 1. π consists of a set (name A) of cardinality 4 and five singleton sets. Since γi(P9) = 3, each singleton set must be an ic-partner of A. On the other hand, by Theo- rem 2.5, A has at most two ic-partners, a contradiction. Case 2. π consists of a set (name A) of cardinality 3, a set (name B) of cardinality 2 and four singleton sets. Since γi(P9) = 3, no two singleton sets in π are ic-partners. Furthermore, by Theorem 2.5, A has at most two ic-partners, so at least two singleton sets of π must be ic-partners of B, which is impossible, as P9 has a unique independent dominating set of cardinality 3. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 541 Case 3. π consists of three sets of cardinality 2, and three singleton sets. We note that each singleton set in π must be an ic-partner of a set of cardinality 2, which is impossible, as P9 has a unique independent dominating set of cardinality 3. The partition {{v1, v3, v5}, {v2, v4, v9}, {v6}, {v7}, {v8}} is an ic-partition of P9. Therefore, IC(P9) = 5. Finally, let n ≥ 10. Consider the path Pn with V (Pn) = {v1, v2, . . . , vn}. By Lemma 3.5 and Observation 2.3, we have IC(Pn) ≤ 6. Now we consider the sets V1 = {v1, v6} ∪ {v2n−1 : n ≥ 5}, V2 = {v2, v5} ∪ {v2n : n ≥ 5}, V3 = {v3}, V4 = {v4}, V5 = {v7}, V6 = {v8}. Then π = {V1, V2, V3, V4, V5, V6} is an ic-partition of Pn, where V3 and V4 are ic-partners of V1, and V5 and V6 are ic-partners of V2. So the proof is complete. Lemma 3.7 ([7]). For any cycle Cn, C(Cn) ≤ 6. Lemma 3.7 and Observation 2.3 imply the following result. Lemma 3.8. For any cycle Cn, IC(Cn) ≤ 6. Lemma 3.9. For any cycle Cn with n ≥ 8 and n ≡ 0 (mod 2), it holds that IC(Cn) = 6. Proof. Let V (Cn) = {v1, v2, . . . , vn}. Consider the sets V1 = {v1, v6}∪{v2n−1 : n ≥ 5}, V2 = {v2, v5} ∪ {v2n : n ≥ 5}, V3 = {v3}, V4 = {v4}, V5 = {v7}, V6 = {v8}. Then π = {V1, V2, V3, V4, V5, V6} is an ic-partition of Cn, for n ≥ 8, where V3 and V4 are ic-partners of V1, and V5 and V6 are ic-partners of V2. Hence, by Lemma 3.8 and Observation 2.3, we have IC(Cn) = 6. Lemma 3.10. For any cycle Cn with n ≥ 8 and n ≡ 0 (mod 3), it holds that IC(Cn) = 6. Proof. Let V (Cn) = {v1, v2, . . . , v3k}. Consider the sets V1 = {v3i+1}, V2 = {v3i+2} and V3 = {v3i+3}, for 0 ≤ i ≤ k − 1. Now for each 1 ≤ i ≤ 3, we partition Vi into two nonempty sets Vi,1 and Vi,2. Observe that Vi,1 and Vi,2 are ic-partners. Hence, by Lemma 3.8 and Observation 2.3, we have IC(Cn) = 6. Lemma 3.11. For any cycle Cn with n ≥ 8 and n ≡ 5 (mod 6), it holds that IC(Cn) = 6. Proof. Assume n = 6k − 1, (k ≥ 2). Let V (Cn) = {v1, v2, . . . , vn}. Consider the sets A = k−1⋃ i=0 {v3i+1}, A1 = 2k−1⋃ i=k {v3i+1}, A2 = 2k−1⋃ i=k {v3i}, B = 2k⋃ i=k {v3i−1}, B1 = k−1⋃ i=1 {v3i−1}, B2 = k−1⋃ i=1 {v3i}. Let π = {A,A1, A2, B,B1, B2}. One can observe that π is an ic-partition of Cn, where A1 and A2 are ic-partners of A, and B1 and B2 are ic-partners of B. Now using Lemma 3.8 and Observation 2.3, we have IC(Cn) = 6. Lemma 3.12. For any cycle Cn with n ≥ 8 and n ≡ 1 (mod 6), it holds that IC(Cn) = 6. 542 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 Proof. Assume n = 6k + 1, (k ≥ 2). Let V (Cn) = {v1, v2, . . . , vn}. Consider the sets A = ( k⋃ i=0 {v3i+1} ) ∪ {v3k+3}, A1 = 2k⋃ i=k+2 {v3i}, A2 = 2k⋃ i=k+2 {v3i−1}, B = ( 2k⋃ i=k+1 {v3i+1} ) ∪ {v3k+2}, B1 = k⋃ i=1 {v3i−1}, B2 = k⋃ i=1 {v3i}. Let π = {A,A1, A2, B,B1, B2}. One can observe that π is an ic-partition of Cn, where A1 and A2 are ic-partners of A, and B1 and B2 are ic-partners of B. Now using Lemma 3.8 and Observation 2.3, we have IC(Cn) = 6. Theorem 3.13. For the cycle Cn, IC(Cn) =  n if n ≤ 6; 5 if n = 7; 6 if n ≥ 8. Proof. If 1 ≤ n ≤ 6, then it is easy to check that IC(Cn) = n. Now assume n = 7. Con- sider the cycle C7 with V (C7) = {v1, v2, v3, v4, v5, v6, v7}. First we show that IC(C7) ̸= 6. Suppose, to the contrary, that IC(C7) = 6. Let π be an IC(C7)-partition. We note that π consists of five singleton sets and a set of cardinality 2 (name A). By Theorem 2.5, A has at most two ic-partners. Hence, π contains two singleton sets that are ic-partners, which contradicts the fact that γi(C7) = 3. The partition {{v1, v3}, {v5}, {v6}, {v4, v7}, {v2}} is an ic-partition of C7, so IC(C7) = 5. Furthermore, by Lemmas 3.9, 3.10, 3.11 and 3.12 we have IC(Cn) = 6, for n ≥ 8. 4 Graphs with small independent coalition number In this section we investigate graphs G with IC(G) ∈ {1, 2, 3, 4}. We will make use of the following two lemmas. Lemma 4.1. Let G be a graph of order n containing r ≥ 1 full vertices, and let F = {v1, v2, . . . , vr} be the set of full vertices of G. Then IC(G) = k, if and only if IC(G[V \ F ]) = k − r, where r < k ≤ n. Proof. Assume first that IC(G[V \ F ]) = k − r. Let π = {V1, V2, . . . , Vk−r} be an IC(G[V \F ])-partition. Now the partition π′ = {V1, V2, . . . , Vk−r, {v1}, {v2}, . . . , {vr}}, is an ic-partition of G, so IC(G) ≥ k. Now we prove that IC(G) = k. Suppose, to the contrary, that IC(G) > k. Let π be an IC(G)-partition. Now the partition π′ = π \ {{v1}, {v2}, . . . , {vr}} is an ic-partition of G[V \F ] such that |π′| > k−r, a contradiction. Hence, IC(G) = k. Conversely, assume that IC(G) = k. Let π be an IC(G)-partition. Now the partition π′ = π \ {{v1}, {v2}, . . . , {vr}} is an ic-partition of G[V \ F ], so IC(G[V \ F ]) ≥ k − r. Now we prove that IC(G[V \ F ]) = k − r. Suppose, to the contrary, that IC(G[V \ F ]) > k − r. Let π be an IC(G[V \ F ])-partition. Now the partition π′ = π ∪ {{v1}, {v2}, . . . , {vr}} is an ic-partition of G such that |π′| > k, a contradiction. Hence, IC(G[V \ F ]) = k − r. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 543 Lemma 4.2. Let G be a graph containing a nonempty set of isolated vertices I . If IC(G) ≥ 3, then for any IC(G)-partition π, there is a set Vr ∈ π such that Vr = I . Proof. First we show that all vertices in I are in the same set of π. Suppose, to the contrary, that there are sets Vi ∈ π and Vj ∈ π such that both Vi and Vj contain isolated vertices. Let Vk ∈ π be an arbitrary set in π such that Vk /∈ {Vi, Vj}. (Since IC(G) ≥ 3, such a set exists). Then Vk has no ic-partner, a contradiction. Now let Vr be the set in π containing isolated vertices. Further, let v be an arbitrary vertex in Vr, and let u ∈ V (G) be an arbitrary vertex such that u ̸= v. If u ∈ Vr, then u is not adjacent to v. Otherwise, the set in π containing u is an ic-partner of Vr, which again implies that u is not adjacent to v. Hence, we have deg(v) = 0. Choosing v arbitrarily, we conclude that Vr = I . Proposition 4.3. Let G be a graph of order n. Then (1) IC(G) = 1 if and only if G ≃ K1. (2) IC(G) = 2 if and only if G ≃ K2 or G ≃ Kn, for some n ≥ 2. Proof. (1) It is clear that IC(G) = 1 if and only if G ≃ K1. (2) If G ≃ K2, then we clearly have IC(G) = 2. Now assume G ≃ Kn, for some n ≥ 2. Let π be an ic-partition of G. Note that no more than two sets in π contain isolated vertices, for otherwise, no two sets in π are ic-partners. Thus, |π| ≤ 2. Partitioning vertices of G into two nonempty sets yields an ic-partition of G. Hence, IC(G) = 2. Conversely, suppose that IC(G) = 2. Let π = {V1, V2} be an IC(G)-partition. If both V1 and V2 are singleton dominating sets, then G ≃ K2. Hence, we may assume that at least one of them (say V1) is not a singleton dominating set. It follows that V2 is not a singleton dominating set either, for otherwise, G is a star, and so by Observation 3.2, we have IC(G) = 3. Hence, V1 and V2 are ic-partners, and so V = V1 ∪ V2 is an independent set. Hence, G ≃ Kn, for some n ≥ 2. Definition 4.4. Let B1 represent the family of bipartite graphs H with partite sets H1 and H2 such that |H1| ≥ 2, |H2| ≥ 2, δ(H) ≥ 1 and id(H) = 2. Definition 4.5. For m ≥ 1, let B2 represent the family of graphs H ∪Km, where H is a bipartite graph with δ(H) ≥ 1 and id(H) = 2. Definition 4.6. For m ≥ 1, let B3 represent the family of graphs H ∪Km, where H is a 3-partite graph with δ(H) ≥ 1 and id(H) = 3. Proposition 4.7. Let G be a graph. Then IC(G) = 3, if and only if G ∈ {K3,K1,n−1} ∪ B2. Proof. Observations 3.1 and 3.2 imply that IC(K3) = 3 and that IC(K1,n−1) = 3, re- spectively. Now let G ∈ B2. Let I be the set of isolated vertices of G, and let {H1, H2} be a partition of G− I into its partite sets. We observe that the partition {I,H1, H2} is an ic-partition of G, so IC(G) ≥ 3. Now we show that IC(G) = 3. Suppose, to the con- trary, that IC(G) ≥ 4. Let π = {V1, V2, . . . , Vk} be an IC(G)-partition. By Lemma 4.2, we have I ∈ {V1, V2, . . . , Vk}. Assume, without loss of generality, that I = V1. Now for each 2 ≤ i ≤ k, Vi forms an independent coalition with V1, and so Vi dominates H . Hence, the partition {V2, . . . , Vk} is an idomatic partition of H , which contradicts the as- sumption. Hence, IC(G) = 3. Conversely, let G be a graph with IC(G) = 3, and let 544 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 π = {V1, V2, V3} be an IC(G)-partition. We consider four cases depending on the number of full vertices of G. Case 1. G has three full vertices. In this case, the sets V1, V2 and V3 are all singleton dominating sets, so G ≃ K3. Case 2. G has two full vertices. Note that this case never occurs. Case 3. G has one full vertex. Let v1 be the full vertex of G. Lemma 4.1 implies that IC(G − v1) = 2. Thus, by Proposition 4.3, either G − v1 ≃ K2, implying that G ≃ K3, or G− v1 ≃ Kn, for some n ≥ 2, which implies that G ≃ K1,n−1, for some n ≥ 3. Case 4. G has no full vertex. Let I be the set of isolated vertices of G. First we note that V1, V2 and V3 are not pairwise ic-partners, for otherwise, we have G ≃ Kn, and so by Proposition 4.3, we have IC(G) = 2, a contradiction. Hence, π contains a set (say V1) that forms an independent coalition with V2 and V3, while V2 and V3 are not ic- partners. Therefore, each vertex in V1 is an isolated vertex, so it follows from Lemma 4.2 that I = V1. Further, the sets V2 and V3 are independent dominating sets of G[V2 ∪ V3], implying that id(G[V2 ∪ V3]) ≥ 2. It remains to show that id(G[V2 ∪ V3]) = 2. Suppose, to the contrary, that id(G[V2 ∪ V3]) ≥ 3. Let π′ = {U1, U2, . . . , Uk}, (k ≥ 3), be an idomatic partition of G[V2 ∪ V3]. Then the partition π′′ = {U1, U2, . . . , Uk, V1} is clearly an ic-partition of G, implying that IC(G) ≥ 4, a contradiction. Hence, G ∈ B2. Proposition 4.8. Let G be a graph. If IC(G) = 4, then G ∈ {K4,K2 +Kn,K1 +B} ∪ B1 ∪ B3, where n ≥ 2 and B ∈ B2. Proof. Let π = {V1, V2, V3, V4} be an ic-partition of G. We consider two cases. Case 1. G has a full vertex. Let v1 be a full vertex of G. Lemma 4.1 implies that IC(G−v1) = 3. Thus, by Proposition 4.7, we have G−v1 ≃ K3, implying that G ≃ K4, or G − v1 ≃ K1,n, for some n ≥ 2, implying that G ≃ K2 + Kn, for some n ≥ 2, or G− v1 ∈ B2, which implies that G ≃ K1 +B, where B ∈ B2. Case 2. G has no full vertex. First assume that G contains a nonempty set I of isolated vertices. Then by Lemma 4.2, we have I ∈ π. Without loss of generality, assume that I = V4. Now for each 1 ≤ i ≤ 3, Vi must form an independent coalition with I . Thus, U = G[V1 ∪ V2 ∪ V3] is a 3-partite graph with id(U) ≥ 3. Since IC(G) = 4, the case id(U) > 3 is impossible. Hence, id(U) = 3, and so G ∈ B3. Now assume that G contains no isolated vertex. Since G has neither full vertices nor isolated vertices, each set of π has either one or two ic-partners. If there is a set of π, (say V1) having one ic-partner, (say V2), then it follows that V3 and V4 are ic-partners, and so G is a bipartitie graph with partite sets V1 ∪ V2 and V3 ∪ V4. Otherwise, assume, without loss of generality, that V2 and V3 are ic-partners of V1. It follows that V4 has an ic-partner in {V2, V3}. By symmetry, we may assume that V4 and V3 are ic-partners. Then G is again a bipartitie graph with partite sets V1 ∪ V2 and V3 ∪ V4. Now using Theorem 2.6, we have id(G) = 2, and so G ∈ B1. 5 Graphs with large independent coalition number Our main goal in this section is to investigate structure of graphs G of order n with IC(G) = n, under specified conditions. In addition, we will characterize all trees T of order n with IC(T ) = n − 1. Let us begin with an observation that characterizes all disconnected graphs G of order n with IC(G) = n. Observation 5.1. Let G be a disconnected graph of order n. Then IC(G) = n if and only if G ≃ Ks ∪Kr, for some s ≥ 1, and r ≥ 1. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 545 Now we introduce two sufficient conditions for a graph G of order n to have indepen- dent coalition number n. Observation 5.2. If G is a graph of order n with α(G) = 2, then IC(G) = n. Proof. Let G be a graph of order n such that α(G) = 2. Consider the singleton partition π1 of G. Note that for any two non-adjacent vertices v and u in V (G), the sets {v} and {u} in π1, are ic-partners. Hence, π1 is an ic-partition of G, and so IC(G) = n. Observation 5.3. Let G be a graph of order n. If G admits a partition of its vertices into two maximal cliques, then IC(G) = n. 5.1 Graphs G with δ(G) = 1 and IC(G) = n In this subsection, we characterize graphs G of order n with δ(G) = 1 and IC(G) = n. We need the following definition. Definition 5.4. Let G be a graph of order n, (n ≥ 3), and let δ(G) = 1. Furthermore, let x be a pendant vertex of G, and let y be the support vertex of x. Then G ∈ F if and only if V (G) \ {x, y} induces a clique. Theorem 5.5. Let G be a graph of order n with δ(G) = 1. Then IC(G) = n if and only if either G ≃ K2, or G ∈ F . Proof. Obviously, IC(K2) = 2. Now assume that G ∈ F . Let x be a pendant vertex of G, and let y be the support vertex of x. Further, let U = V (G) \ {x, y}. Note that U contains no full vertex. If y is a full vertex, then G is obtained from the complete graph Kn−1, where one of its vertices is adjacent to a leaf. In this case, we clearly have IC(G) = n. Thus, we may assume that y is not a full vertex, that is, there is a vertex u ∈ U such that u is not adjacent to y. Then it is easy to verify that the sets {y} and {u} are ic-partners, and that each vertex in U \ {u} forms an independent coalition with {x}. Therefore, IC(G) = n. Conversely, suppose that G is a graph with δ(G) = 1 and IC(G) = n. Let x be a leaf of G, and let y be the support vertex of x. Consider the singleton partition π1 of G. Note that each set in π1\{{x}, {y}} must be an ic-partner of {x} or {y}, to dominate x. Let A = N(y) \ {x}, and B = V (G) \ ({x, y} ∪A). We consider four cases. Case 1. A = ∅ and B = ∅. In this case, we have G ≃ K2. Case 2. A = ∅ and B ̸= ∅. By Observation 5.1, we have G ≃ K2 ∪ Kr, for some r ≥ 1. Thus, G ∈ F . Case 3. A ̸= ∅ and B = ∅. For each v ∈ A, the set {v} cannot be an ic-partner of {y}, so it must be an ic-partner of {x}. This implies that A induces a clique. Hence, G ∈ F . Case 4. A ̸= ∅ and B ̸= ∅. For each v ∈ A, the set {v} cannot be an ic-partner of {y}, so it must be an ic-partner of {x}. This implies that [A,B] is full and that A induces a clique. Now for each vertex u ∈ B, in order for the set {u} to be an ic-partner of {x} or {y}, u must be adjacent to all other vertices in B. Hence, B induces a clique, and so G ∈ F , which completes the proof. As an immediate result from Theorem 5.5 we have: Corollary 5.6. Let T be a tree of order n. Then IC(T ) = n if and only if T ∈ {P1, P2, P3, P4}. 546 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 5.2 Triangle-free graphs G with IC(G) = n In this subsection, we characterize graphs G of order n with g(G) = 4 and IC(G) = n. This will lead to characterization of all triangle-free graphs G of order n with IC(G) = n. We will make use the following lemmas. Lemma 5.7. Let G be a triangle-free graph of order n with IC(G) = n. Then g(G) ≤ 6. Proof. Let G be a graph of order n with IC(G) = n, and suppose, to the contrary, that g(G) ≥ 7. Let C ⊆ G be a cycle of order g(G). Consider an arbitrary vertex v ∈ V (C). Note that γi(C) ≥ 3, and so {v} is not an ic-partner of any set {u} ⊂ V (C). Therefore, it must be an ic-partner of a set {u} ⊆ V (G) \ V (C). It follows that, {u} dominates V (C) \Nc[v], which implies that G contains triangles, a contradiction. Lemma 5.8. Let G be a graph of order n with g(G) = 6. Then IC(G) = n if and only if G ≃ C6. Proof. Let G be a graph of order n with g(G) = 6. If G ≃ C6, then by Theorem 3.13, we have IC(G) = 6. Conversely, assume that IC(G) = n. Let C ⊆ G be a cycle of order 6, and suppose, to the contrary, that V (G) \ V (C) ̸= ∅. Consider an arbitrary vertex v ∈ V (G) \ V (C). If {v} is an ic-partner of a set {u} ⊂ V (C), then {v} must dominate V (C) \ Nc[u], which implies that G contains triangles, a contradiction. Otherwise, {v} must be an ic-partner of a set {u} ⊂ V (G) \ V (C). Now since {u, v} dominates C, it follows that G contains triangles, or induces cycles of order 4, a contradiction. Our next result can be established almost the same way as Lemma 5.8, so we state it without proof. Lemma 5.9. Let G be a graph of order n with g(G) = 5. Then IC(G) = n if and only if G ≃ C5. In order to characterize graphs G of order n with IC(G) = n and g(G) = 4, we need the following definitions. Definition 5.10. Let K0 represent a bipartite graph with partite sets H1 = {v1, v2, v3, v4} and H2 = {u1, u2, u3, u4} such that for each 1 ≤ i ≤ 4, vi is adjacent to all vertices in H2, except ui (see Figure 2). Definition 5.11. Let K represent a family of 4-partite graphs with partite sets H1 = {v1, v2, v3, v4}, H2 = {u1, u2, u3, u4}, H3 = {n1, n2, . . . , nk} and H4 = {m1,m2, . . . ,mk}, for k ≥ 1, with the following properties: • [H1, H3] is full and [H2, H4] is full, • [H1, H4] is empty and [H2, H3] is empty, • For each 1 ≤ i ≤ 4, vi is adjacent to all vertices in H2, except ui, • For each 1 ≤ i ≤ k, ni is adjacent to all vertices in H4, except mi. Figure 3 illustrates such a graph for k = 3. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 547 v1 v2 v3 v4 u1 u2 u3 u4 Figure 2: The graph K0. v1 v2 v3 v4 u4 u3 u2 u1 n1 n2 n3 m3 m2 m1 Figure 3: The graph in K for k = 3. Theorem 5.12. Let G be a graph of order n with g(G) = 4. Then IC(G) = n if and only if G ∈ {C4,K0} ∪ K. Proof. It is easy to check that IC(C4) = 4 and that IC(K0) = 8. Now let G ∈ K. We observe that for each 1 ≤ i ≤ 4, {vi} and {ui} are ic-partners, and that for each 1 ≤ i ≤ k, {ni} and {mi} are ic-partners. Thus, IC(G) = n. Conversely, let G be a graph of order n with g(G) = 4 and IC(G) = n, and let C be a cycle of G of order 4 with V (C) = {x, y, z, t} and E(C) = {xy, yz, zt, tx}. If G = C, then the desired result follows. Hence, we assume that G ̸= C. Since x is adjacent to y and t, neither {y} nor {t} is an ic-partner of {x}. Now consider two cases. Case 1. {x} and {z} are ic-partners. In this case, G is dominated by {x, z}. Let A = N(x) \ {y, t} and B = N(z) \ {y, t}. If A ̸= ∅, (say v ∈ A), then it is not difficult to check that {v} has no ic-partner. Thus, A = ∅, and so by symmetry, we have B = ∅. Hence, G ≃ C4. Case 2. {x} and {z} are not ic-partners. Let {e} be an ic-partner of {x}. Since {x, e} dominates G and z is not adjacent to x, it must be adjacent to e. Let A = N(x) \ {y, t} and B = N(e) \ {z}. It is not difficult to verify that A ∩ B = ∅. Now if A = ∅, then {z} cannot form an independent coalition with any other set, so A ̸= ∅. Let {f} ⊆ A be an ic-partner of {z}. We note that if a set {g} forms an independent coalition with {y}, then g ∈ B. Further, if a set {h} forms an independent coalition with {t} then h ∈ B. Let {g} and {h} be ic-partners of {y} and {t}, respectively. Observe that {g} ≠ {h}. Now let A′ = A \ {f} and B′ = B \ {g, h}. There exist the following subcases. 548 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 Subcase 2.1. A′ = ∅ and B′ = ∅. In this case, we have G ≃ K0. Subcase 2.2. A′ = ∅ and B′ ̸= ∅. Let v ∈ B′. One can verify that {v} cannot form an independent coalition with any other set. Thus, this case is impossible. Subcase 2.3. A′ ̸= ∅ and B′ = ∅. Let v ∈ A′. One can verify that {v} cannot form an independent coalition with any other set. Thus, this case is impossible. Subcase 2.4. A′ ̸= ∅ and B′ ̸= ∅. Let v ∈ A′. If a set {u} forms an independent coalition with {v}, then u ∈ B′. Furthermore, for each vertex u ∈ B′, {u} cannot form an independent coalition with more than one sets {v} ⊆ A′. Thus, |A′| ≤ |B′|. Using a similar argument, we deduce that |B′| ≤ |A′|, and so |A′| = |B′|. Consequently, the following statements hold in the graph G: • G[{x, y, z, t, e, f, g, h}] is a bipartite graph with partite sets V1 = {x, z, g, h} and V2 = {y, t, e, f}, which is isomorphic to K0, • [V1, A′] is full and [V2, B′] is full, • [V1, B′] is empty and [V2, A′] is empty, • G[A′∪B′] is a bipartite graph with partite sets A′ and B′ such that degG[A′∪B′](v) = |A′| − 1 = |B′| − 1, for each v ∈ A′ ∪B′. Hence, G ∈ K and the proof is complete. Using Observation 5.1, Corollary 5.6, Lemmas 5.7, 5.8 and 5.9, and Theorem 5.12, we infer the following result. Corollary 5.13. Let G be a triangle-free graph of order n. Then IC(G) = n if and only if G ∈ {C4, C5, C6, P1, P2, P3, P4,K2,K1 ∪K2,K2 ∪K2,K0} ∪ K. 5.3 Trees T with IC(T ) = n − 1 The following theorem characterizes all trees T of order n with IC(T ) = n− 1. Theorem 5.14. Let T be a tree of order n. Then IC(T ) = n − 1 if and only if T ∈ {P5, P6, S1,2,K1,3}. Proof. By Theorem 3.6, we have IC(P5) = 4 and IC(P6) = 5. Further, by Observa- tion 3.2, we have IC(K1,3) = 3 and by Observation 3.3, we have IC(S1,2) = 4. Con- versely, let T be a tree of order n with IC(T ) = n − 1, where x is a leaf, and y is the support vertex of x. Define A = N(y) \ {x} and B = V (G) \ ({x, y} ∪ A). Further, let π be an IC(T )-partition. Note that π contains a set of cardinality 2 (say V1 = {u, v}) and n − 2 singleton sets. Since x and y are adjacent, we have V1 ̸= {x, y}. Note as well that any set in π must be an ic-partner of the set containing x, or the set containing y, to dominate x. We consider two cases. Case 1. B = ∅. If A = ∅, then we have T ≃ K2, and so IC(T ) = 2 ̸= n− 1. Hence, A ̸= ∅, and so T ≃ K1,n−1, for some n ≥ 3. Now by Lemma 3.7, we have IC(T ) = 3. Hence, T ≃ K1,3. Case 2. B ̸= ∅. Since T is connected, we have A ̸= ∅. We divide this case into some subcases. Subcase 2.1. u ∈ A and v ∈ B. We first show that |A| = 1. Suppose, to the contrary, that |A| ≥ 2. Then there is a vertex z ∈ A such that z ̸= u. Since z and y are adjacent, M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 549 {z} cannot be an ic-partner of {y}, so it must be an ic-partner of {x}. Since {x} does not dominate u, u must be adjacent to z, which is a contradiction, since y, z and u induce a triangle. Now {y} cannot be an ic-partner of {x} or {u, v}, so it must have an ic-partner in B. This implies that |B| ≥ 2. Let {t} ⊂ B be an ic-partner of {y}. we show that B \ {v, t} = ∅. Suppose that B \ {v, t} ̸= ∅. Let z ∈ B \ {v, t}. Note that v and t are adjacent. Now {z} must be an ic-partner of {x} or {y}, so z must be adjacent to t and v, which is a contradiction, since z, t and v induce a triangle. Hence, B = {v, t} and so T ≃ P5. Subcase 2.2. {u, v} ⊆ B. An argument similar to the one presented above implies that |A| = 1. Now we show that B \ {u, v} = ∅. Suppose that B \ {u, v} ≠ ∅. Let z ∈ B \ {u, v}, and let A = {t}. Since t and y are adjacent, {t} cannot be an ic-partner of {y}, so it must be an ic-partner of {x}. Thus, t must be adjacent to u, v and z. Now {z} must be an ic-partner of {x} or {y}, so z must be adjacent to u and v, which is a contradiction, since z, u and t induce a triangle. Hence, T ≃ S1,2. Subcase 2.3. {u, v} ⊆ A. We first show that A\{u, v} = ∅. Suppose that A\{u, v} ≠ ∅. Let z ∈ A \ {u, v}. Since z is adjacent to y, {z} must be an ic-partner of {x}, so z must be adjacent to u and v, which is a contradiction, since z, u and y induce a triangle. Now we show that |B| = 1. Suppose that |B| ≠ 1. First assume |B| ≥ 3. Let z, t, w ∈ B. Now z, t and w induce a triangle, since the sets containing each of them, must be an ic-partner of {x} or {y}, a contradiction. Now assume |B| = 2. Let B = {z, t}. Each of the sets {z} and {t} must be an ic-partner of {x} or {y}. Thus, z must be adjacent to t. Now {u, v} must be an ic-partner of {x}, so z and t must be dominated by {u, v}. Now the induced subgraph T [{u, v, z, t}] contains at least one cycle, a contradiction. Hence, we have T ≃ S1,2. Subcase 2.4. u = y and v ∈ B. we first show that |A| = 1. Suppose that |A| ≥ 2. Let z, t ∈ A. Since z and t are adjacent to y, {z} and {t} cannot be an ic-partner of y, so each of them must be an ic-partner of {x}. Thus, z must be adjacent to t, which is a contradiction, since z, t and y induce a triangle. Now we show that |B| = 2. Suppose that |B| ̸= 2 . If |B| = 1, then T ≃ P4, a contradiction. Otherwise, let {v, z, t} ⊆ B and A = {w}. Now {w} cannot be an ic-partner of {u, v}, so it must be an ic-partner of {x}. Thus, w must be adjacent to v, z and t. Now observe that {u, v} must have an ic-partner in B. Assume, without loss of generality, that {u, v} and {z} are ic-partners. This implies that t is adjacent to z or v, which is impossible, since both cases lead to existence of an induced triangle. Hence, T ≃ S1,2. Subcase 2.5. u = x and v ∈ A. We first show that |B| ≤ 2. Suppose that |B| ≥ 3. Let {z, t, w} ⊆ B. Note that {y} must have an ic-partner in B. Assume, without loss of generality, that {y} and {z} are ic-partners. It follows that z is adjacent to t and w. Now if {y} is an ic-partner of {t} or {w}, then t must be adjacent to w, which is impossible, since z, t and w induce a triangle. Hence, both t and w must be ic-partners of {u, v}, which implies that t is adjacent to w. Now z, t and w induce a triangle, a contradiction. Now we show that |A| ≤ 2. Suppose that |A| ≥ 3. Let {z, t, v} ⊆ A. The sets {z} and {t} must be ic-partners of {u, v}. This implies that z is adjacent to t. Now z, t and y induce a triangle, a contradiction. Further, we observe that the case |A| = |B| = 2 is impossible. Hence, either |A| = 2 and |B| = 1, which implies that T ≃ S1,2, or |A| = 1 and |B| = 2, which implies that T ≃ P5. Subcase 2.6. u = x and v ∈ B. We first show that |A| = 1. Suppose that |A| ≥ 2. Let {z, t} ⊆ A. Now each of the sets {z} and {t} must be an ic-partner of {u, v}. This 550 Ars Math. Contemp. 24 (2024) #P3.09 / 535–551 implies that z is adjacent to t, which is a contradiction, since y, z and t induce a triangle. Now we show that |B| ≤ 3. Suppose that |B| ≥ 4. Let {v, t, w, h} ⊆ B and A = {z}. Note that {y} must have an ic-partner in B. Assume, without loss of generality, that {y} and {t} are ic-partners. It follows that t is adjacent to w, h and v. Now {w} must be an ic-partner of {y} or {u, v}. One can observe that both cases lead to contradiction. Hence, either |B| = 2, which implies that T ≃ P5, or |B| = 3, which implies that T ≃ P6. 6 Discussion and conclusions In Proposition 2.2, we introduced a family of graphs admitting no ic-partition. This result motivates the following problem: Problem 6.1. Characterize graphs admitting an ic-partition. In Observations 2.3 and 2.4, we presented the sharp inequalities IC(G) ≤ C(G) and IC(G) ≥ χ(G). This raises the following problems: Problem 6.2. Characterize graphs G in which the equality IC(G) = C(G) holds. Problem 6.3. Characterize graphs G in which the equality IC(G) = χ(G) holds. In Theorem 5.14, trees T of order n, with IC(T ) = n − 1 have been characterized. This raises the following problem: Problem 6.4. Characterize graphs G of order n with IC(G) = n− 1. ORCID iDs Mohammad Reza Samadzadeh https://orcid.org/0009-0004-4370-5905 Doost Ali Mojdeh https://orcid.org/0000-0001-9373-3390 References [1] A. Buğra Özer, E. Saygı and Z. Saygı, Domination type parameters of Pell graphs, Ars Math. Contemp. 23 (2023), Paper No. 1.03, 12 pp., doi:10.26493/1855-3974.2637.f61, https:// doi.org/10.26493/1855-3974.2637.f61. [2] A. Cabrera Martı́nez, A note on the k-tuple domination number of graphs, Ars Math. Contemp. 22 (2022), Paper No. 4.03, 5 pp., doi:10.26493/1855-3974.2600.dcc, https://doi.org/ 10.26493/1855-3974.2600.dcc. [3] A. Cabrera Martı́nez, A. Estrada-Moreno and J. A. Rodrı́guez-Velázquez, From Italian dom- ination in lexicographic product graphs to w-domination in graphs, Ars Math. Contemp. 22 (2022), Paper No. 1.04, 25 pp., doi:10.26493/1855-3974.2318.fb9, https://doi.org/ 10.26493/1855-3974.2318.fb9. [4] G. J. Chang, The domatic number problem, volume 125, pp. 115–122, 1994, doi:10.1016/ 0012-365X(94)90151-1, https://doi.org/10.1016/0012-365X(94)90151-1. [5] A. Gorzkowska, M. A. Henning, M. Pilśniak and E. Tumidajewicz, Paired domination stability in graphs, Ars Math. Contemp. 22 (2022), Paper No. 2.04, 18 pp., doi:10.26493/1855-3974. 2522.eb3, https://doi.org/10.26493/1855-3974.2522.eb3. [6] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. McRae and N. Phillips, The upper domatic number of a graph, AKCE Int. J. Graphs Comb. 17 (2020), 139–148, doi:10.1016/j.akcej.2018. 09.003, https://doi.org/10.1016/j.akcej.2018.09.003. M. R. Samadzadeh and D. A. Mojdeh: Independent coalition in graphs: existence and . . . 551 [7] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. A. McRae and R. Mohan, Introduction to coalitions in graphs, AKCE Int. J. Graphs Comb. 17 (2020), 653–659, doi:10.1080/09728600. 2020.1832874, https://doi.org/10.1080/09728600.2020.1832874. [8] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. A. McRae and R. Mohan, Upper bounds on the coalition number, Australas. J. Comb. 80 (2021), 442–453, https://ajc.maths. uq.edu.au/pdf/80/ajc_v80_p442.pdf. [9] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. A. McRae and R. Mohan, Coalition graphs, Commun. Comb. Optim. 8 (2023), 423–430, doi:10.22049/CCO.2022.27916.1394, https://doi.org/10.22049/CCO.2022.27916.1394. [10] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. A. McRae and R. Mohan, Coalition graphs of paths, cycles, and trees, Discuss. Math. Graph Theory 43 (2023), 931–946, doi: 10.7151/dmgt.2416, https://doi.org/10.7151/dmgt.2416. [11] T. W. Haynes, J. T. Hedetniemi, S. T. Hedetniemi, A. A. McRae and R. Mohan, Self-coalition graphs, Opuscula Math. 43 (2023), 173–183, doi:10.7494/opmath.2023.43.2.173, https:// doi.org/10.7494/opmath.2023.43.2.173. [12] T. W. Haynes, S. T. Hedetniemi and P. J. Slater, Fundamentals of Domination in Graphs, vol- ume 208 of Monographs and Textbooks in Pure and Applied Mathematics, Marcel Dekker, Inc., New York, 1998. [13] D. B. West, Introduction to Graph Theory, Prentice Hall, NJ, 2nd edition, 2001. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 24 (2024) #P3.10 / 553–566 https://doi.org/10.26493/1855-3974.3227.fd5 (Also available at http://amc-journal.eu) Spectra of signed graphs and related oriented graphs Zoran Stanić * Faculty of Mathematics, University of Belgrade, Studentski trg 16, 11 000 Belgrade, Serbia Received 14 September 2023, accepted 23 April 2024, published online 6 September 2024 Abstract For every oriented graph G′, there exists a bipartite signed graph Ḣ such that the spec- trum of Ḣ contains the full information about the spectrum of the skew adjacency ma- trix of G′. This allows us to transfer some problems concerning the skew eigenvalues of oriented graphs to the framework of signed graphs, where the theory of real symmetric matrices can be employed. In this paper, we continue the previous research by relating the characteristic polynomials, eigenspaces and the energy of G′ to those of Ḣ . Simul- taneously, we address some open problems concerning the skew eigenvalues of oriented graphs. Keywords: Oriented graph, signed graph, eigenvalues, characteristic polynomial, eigenspaces, en- ergy. Math. Subj. Class. (2020): 05C20, 05C22, 05C50 1 Introduction For a finite simple undirected graph G = (V,E), an oriented graph G′ is a pair (G, σ′), where σ′ is the edge orientation satisfying σ′(ij) ∈ {i, j}, for every ij ∈ E. Similarly, a signed graph Ġ is a pair (G, σ̇), where σ̇ is the edge signature satisfying σ̇(ij) ∈ {+1,−1}, for every ij ∈ E. In both cases, G is referred to as the underlying graph. The order n is the number of vertices of G. The edge set of G′ consists of oriented edges, where the edge ij is oriented from i to j if σ′(ij) = j; this is designated by i → j (or j ← i). The edge set of Ġ consists of positive and negative edges. *The research is supported by the Science Fund of the Republic of Serbia; grant number 7749676: Spectrally Constrained Signed Graphs with Applications in Coding Theory and Control Theory – SCSG-ctct. E-mail address: zstanic@matf.bg.ac.rs (Zoran Stanić) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 554 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 The skew adjacency matrix SG′ = (sij) of G′ is the n × n matrix such that sij = 0 if ij is not an edge of G′, sij = 1 if i → j, and sij = −1 otherwise. This matrix is skew symmetric and differs from the adjacency matrix of G′ whose (i, j)-entry is 0 whenever i ← j. The characteristic polynomial, the eigenvalues and the spectrum of SG′ are known as the skew characteristic polynomial, the skew eigenvalues, the skew spectrum of G′, respectively. To easy language and be consistent with the forthcoming terminology for signed graphs, in this article we omit the prefix ‘skew’. The spectrum of G′ consists of purely imaginary numbers, the non-zero eigenvalues come as complex conjugates, and thus the rank of SG′ is even. Also, S ⊺ G′SG′ = −S2G′ . The adjacency matrix AĠ of Ġ is obtained from the standard (0, 1)-adjacency matrix of G by reversing the sign of all edges mapped to −1 by σ̇. By the characteristic poly- nomial, the eigenvalues and the spectrum of Ġ we mean the characteristic polynomial, the eigenvalues and the spectrum of AĠ, respectively. Since AĠ is symmetric, its eigenvalues are real. An r-cube or a hypercube Qr is the r-regular graph of order 2r with vertex set {0, 1}r (all possible binary r-tuples) in which two vertices are adjacent if and only if they differ in exactly one coordinate. Accordingly, an oriented (resp. signed) r-cube is an oriented (signed) graph underlined by Qr. If Γ is either an oriented graph or a signed graph, then its energy E(Γ) is the sum of modulus of its eigenvalues. It follows from [19] that every oriented graph G′ is related to a bipartite signed graph Ḣ in such a way that the spectrum of Ḣ contains the full information about the spectrum of G′. All necessary details about this relation are given in the next section. This means that the theory of spectra of oriented graphs is strongly connected to the theory of spectra of signed graphs, and that many problems concerning spectral parameters of oriented graphs can be transferred to the framework of signed graphs, where the entire theory of real symmetric matrices can be employed. In this way, some known results on oriented graphs are proved in an elementary way [19], some open problems are resolved [15] and some known results concerning signed graphs are transferred to the context of oriented graphs [18]. In this article we continue the research by expressing the coefficients of the character- istic polynomial of G′ in terms of the coefficients of the characteristic polynomial of Ḣ . We also generate the eigenspaces of G′ on the basis of the eigenvectors of Ḣ and relate the energies of both graphs. The results on characteristic polynomials and energies are of particular interest since they address some open problems posed in literature. The results on eigenspaces are followed by an immediate application in the engineering domain. Section 2 contains additional terminology and notation, along with a short review of re- sults of [19]. In particular, oriented graphs are related to signed graphs in the forthcoming Theorem 2.1. Some comments and results that arise directly from this theorem are given in Section 3. Characteristic polynomials, eigenspaces and energies are considered in Sec- tions 4–6, respectively. Some notes on particular (oriented or signed) hypercubes are given in Section 7. 2 Preliminaries We say that an oriented or a signed graph is connected, regular, or bipartite if the same holds for its underlying graph. Similarly, a matching (or a perfect matching) refers to the matching in the underlying graph. Two oriented (resp. signed) graphs with the same underlying graph are switching equiv- Z. Stanić: Spectra of signed graphs and related oriented graphs 555 alent if there is a subset U of the vertex set V , such that one of them is obtained by reversing the orientation (sign) of every edge located between U and V \ U . In matrix terminology, G′1 and G ′ 2 are switching equivalent if there exists a diagonal matrix S with±1 on the main diagonal, such that SG′2 = S −1SG′1S, and similarly for signed graphs. S is referred to as the switching matrix. Observe that the spectrum remains unchanged under the switching operation. We say that an oriented even cycle C ′2ℓ is oriented uniformly if by traversing along the cycle we pass through an odd (resp. even) number of edges oriented in the route direction for ℓ odd (even), where the ‘route direction’ refers to any of two possible directions: clock- wise or counterclockwise. A canonical orientation in a bipartite graph G is the orientation which orients all the edges from one colour class to the other. Clearly, in this orientation every cycle is oriented uniformly. A cycle Ċ in a signed graph is positive if the product of its edge signs σ̇(Ċ) is 1. Otherwise, it is negative. A signed graph is said to be homogeneous if all edges have the same sign, e.g., if its edge set is empty. It is balanced if it switches to its underlying graph; equivalently, it does not contain negative cycles. The negation −Ġ of a signed graph is obtained by reversing the sign of every edge of Ġ. Observe that if Ġ is bipartite, then it is switching equivalent to −Ġ and they share the same spectrum. We proceed with results of [19]. The signature σ̇ is associated with the orientation σ′ (and also Ġ is associated with G′) if σ̇(ik)σ̇(jk) = siksjk holds for every pair of edges ik and jk. (2.1) Being associated is a symmetric relation. We write σ̇ ∼ σ′ to indicate that σ̇ and σ′ are mutually associated. The following results hold: If σ̇ ∼ σ′, then −S2G′ = A2Ġ. For a graph G and an orientation σ′, there exists a signature σ̇ associated with σ′ if and only if G is bipartite. The next result gives a crucial relation between the spectrum of an oriented graph and the spectrum of a related signed graph. If G′ is an oriented graph, then its bipartite dou- ble bd(G′) is the oriented graph whose skew adjacency matrix is the Kronecker product Sbd(G′) = AK2 ⊗ SG′ , where K2 is the complete graph with 2 vertices. This defini- tion extends the definition of a bipartite double of an ordinary graph, where bd(G) has Abd(G) = AK2 ⊗ AG as the adjacency matrix. Evidently, if G underlies G′, then bd(G) underlies bd(G′). In our notation, exponents denote multiplicities of the eigenvalues. Theorem 2.1. Let G′ = (G, σ′) be an oriented graph with rank(SG′) = 2k. The following statements hold true: (i) If G′ is bipartite and σ̇ ∼ σ′, then ±iλ1,±iλ2, . . . ,±iλk, 0(n−2k) are the eigen- values of G′ if and only if ±λ1,±λ2, . . . ,±λk, 0(n−2k) are the eigenvalues of Ġ = (G, σ̇). (ii) If G′ is non-bipartite, H ′ = (bd(G), σ′′) is a bipartite double of G′ and σ̇ ∼ σ′′, then ±iλ1,±iλ2, . . . ,±iλk, 0(n−2k) are the eigenvalues of G′ if and only if (±λ1)(2), (±λ2)(2), . . . , (±λk)(2), 0(2n−4k) are the eigenvalues of Ḣ = (bd(G), σ̇). Therefore, G′ is related to a bipartite signed graph whose spectrum gives the full infor- mation on the spectrum of G′. If G′ is bipartite, a required signed graph is its associate. If G′ is non-bipartite, then a required signed graph is associated with bd(G′). One may 556 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 Figure 1: An infinite family of oriented graphs G′ and signed graphs Ġ such that −S2G′ = A2 Ġ . Negative edges are dashed. notice that item (ii) of the previous theorem covers the bipartite case in the sense that, if G′ is bipartite, then bd(G′) consists of two copies of G′ and Ḣ also has two identical copies, each associated with G′. However, this would lead to unnecessary complicating, as there is no need to deal with a bipartite double if G′ is already bipartite. Thus, the bipartite case is separated in item (i) of the same theorem. 3 Comments to Theorem 2.1 In the bipartite case, G′ is associated with Ġ if and only if it is associated with −Ġ. Hence, G′ actually has two associates (with the same spectrum). Similarly, Ġ is associ- ated with G′ if and only if it is associated with the oriented graph obtained by reversing the orientation of every edge of G′. Again, Ġ has two associates which share the same spec- trum. Henceforth, when we say ‘let Ġ be a signed graph associated with G′’ (or something similar), we always mean that Ġ is allowed to take any of the two options (that differ up to negation). We have pointed out in the previous section that−S2G′ = A2Ġ holds whenever G ′ and Ġ are associated. However, this identity is not exclusively reserved for associated graphs. For example, Figure 1 illustrates infinite families of graphs that are not mutually associated in the sense of the equality (2.1) (since they are non-bipartite), but satisfy the previous matrix identity. In this case, they share the same underlying graph, but the identity can occur even if they do not. In fact, the identity occurs if and only if for every pair i, j of vertices, −s(2)ij = a (2) ij holds, where an exponent indicates that we deal with the entry of matrix square. This leads to the following result. Theorem 3.1. If an oriented graph G′ and a signed graph Ġ are defined on the same vertex set, then −S2G′ = A2Ġ holds if and only if for every pair of vertices i, j∣∣{k : (i→ k∧j → k)∨(i← k∧j ← k)}∣∣−∣∣{k : (i→ k∧j ← k)∨(i← k∧j → k)}∣∣ in G′ is equal to ∣∣{k : σ̇(ik)σ̇(jk) = 1}∣∣− ∣∣{k : σ̇(ik)σ̇(jk) = −1}∣∣ in Ġ. Z. Stanić: Spectra of signed graphs and related oriented graphs 557 To demonstrate an application of Theorem 2.1, we deduce a known result. Observe that a bipartite canonically oriented graph is associated with a homogeneous signed graph, necessarily switching equivalent to its underlying graph, and so sharing the spectrum with it. Since bipartite oriented graphs are switching equivalent if and only if associated signed graphs are switching equivalent (where the equivalence is realized by the same switch- ing matrix), we deduce the following result (conjectured in [7], and proved in [3]): If ±iλ1,±iλ2, . . . ,±iλn are the skew eigenvalues of a bipartite oriented graph G′ = (G, σ′), then the eigenvalues of G are ±λ1,±λ2, . . . ,±λn if and only if G′ switches to a canoni- cally oriented graph. Here are more details on the cycle structure of Ḣ of Theorem 2.1(ii). It will be used in the forthcoming sections. Theorem 3.2. Let G′ and Ḣ be as in Theorem 2.1(ii). For every odd cycle C ′ of G′, the signed cycle Ċ of Ḣ associated with bd(C ′) is negative. Proof. If the vertices of C ′, labelled in the natural order, are 1, 2, . . . , ℓ, then the vertices of its bipartite double bd(C ′) are divided into the colour classes, say A = {a1, a2, . . . , aℓ} and B{b1, b2, . . . , bℓ}, the edges of bd(C ′) are a1b2, b2a3, a3b4, . . . , bℓ−1aℓ, aℓb1, b1a2, . . . , aℓ−1bℓ, bℓa1, and these edges inherit the orientation from G′ in the sense that i→ j implies ai → bj and bi → aj . Now, if the edges of G′ are oriented in the route direction, so are the edges of bd(G′). In this case, the edges of the associated signed cycle Ċ alternate in sign, which in particular means that Ċ is negative (as it counts exactly ℓ negative edges and ℓ is odd) and the edges aibj and biaj differ in sign (again, since ℓ is odd). If the edges of G′ are not oriented in the route direction, then C ′ is obtained from the previous cycle by reversing the orientation of some edges. The desired conclusion follows since changing the orientation of a single edge ij changes the orientation of both aibj and biaj (in bd(C ′)) and changes the sign of both aibj and biaj (in Ċ), so it does not change the signature of Ċ. Corollary 3.3. Let G′ and Ḣ be as in Theorem 2.1(ii). Then Ḣ is unbalanced and bd(G′) contains a cycle that is not oriented uniformly. Proof. The first statement follows from the previous theorem. The second one follows from an easy observation that a negative cycle in Ḣ corresponds to a cycle in bd(G′) that is not oriented uniformly. 4 Characteristic polynomials Here are relations between the coefficients of the characteristic polynomials. Theorem 4.1. If ∑n i=0 six n−i is the characteristic polynomial of an oriented graph G′, then si = 0 for i odd and si ≥ 0 for i even. If G′ is bipartite and ∑⌊n/2⌋ i=0 a2ix n−2i is the characteristic polynomial of an associated signed graph, then a2i = (−1)is2i. (4.1) If G′ is non-bipartite and ∑n i=0 a2ix n−2i is the characteristic polynomial of Ḣ (where Ḣ is as in the formulation of Theorem 2.1(ii)), then 558 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 a2i = (−1)i min{i,n−i}∑ ℓ = −min{i, n − i} i ≡ ℓ (mod 2) si−ℓsi+ℓ. (4.2) Proof. Under the assumptions of Theorem 2.1, the characteristic polynomial of the skew adjacency matrix of G′ is xn + s1x n−1 + · · ·+ sn−1x+ sn = xn−2k(x2 + λ21)(x2 + λ22) · · · (x2 + λ2k). (4.3) We immediately obtain si = 0 for i odd (since si is the ith elementary symmetric polyno- mial in eigenvalues), and si ≥ 0 for i even. If G′ is bipartite, then the characteristic polynomial of an associated signed graph reads∑⌊n/2⌋ i=0 a2ix n−2i = xn−2k(x2 − λ21)(x2 − λ22) · · · (x2 − λ2k), which yields that even coef- ficients alternate in sign, and |a2i| = s2i. This implies the equality (4.1). If G′ is non-bipartite, then the characteristic polynomial of Ḣ is n∑ i=0 a2ix n−2i = ( xn−2k(x2 − λ21)(x2 − λ22) · · · (x2 − λ2k) )2 . Comparing it with (4.3), we arrive at a2i = (−1)i min{i,n−i}∑ ℓ=−min{i,n−i} si−ℓsi+ℓ, but we know that odd coefficients under the sum are zero, so the summation reduces to even coefficients, which leads to (4.2). For 2i ≤ n, we note that the formula (4.2) is simplified to a2i = (−1)i i∑ ℓ=0 s2ℓs2(i−ℓ). We visualize this in the following example. Example 4.2. Let G′ be the oriented graph with 10 vertices illustrated in Figure 1. The coefficients of its characteristic polynomial are (s0, s2, . . . , s10) = (1, 15, 60, 92, 48, 0). The coefficients of the characteristic polynomial of Ḣ of Theorem 2.1(ii) are (a0, a2, . . . , a20) = (1,−30, 345,−1984, 6456,−12480, 14224,−8832, 2304, 0, 0). Say, 345 = a4 = s0s4 + s22 + s4s0 = 60 + 225 + 60 or −8832 = a14 = (−1)7(s10s4 + s8s6 + s6s8 + s4s10) = 0− 2 · 92 · 48 + 0, as in Theorem 4.1. We recall that a basic figure in a graph is a disjoint union of edges and cycles. If∑n i=0 aix i is the characteristic polynomial of the adjacency matrix of a signed graph Ġ = (G, σ̇), then from [4] ai = ∑ B∈Bi (−1)p(B)σ̇(B)2|c(B)|, Z. Stanić: Spectra of signed graphs and related oriented graphs 559 where Bi is the set of basic figures on i vertices in G, p(B) is the number of components of B, c(B) is the set of cycles in B and σ̇(B) = ∏ Ċ∈c(B) σ̇(Ċ). This result can be extended to oriented graphs on the basis of Theorems 2.1 and 4.1. However, this has already been performed in [5, pages 4516–4517] and [9, Theorem 2.3], and so we just refer the reader to these references. It is worth mentioning that these results address the research problem of [1, Section 6], asking for an interpretation of the coefficients si in terms of G′. We conclude this section by the following observation. If the characteristic polynomials of G (the underlying graph), Ġ = (G, σ̇) (a signed graph) and G′ = (G, σ′) (an oriented graph) are ∑n i=0 ai(G)x n−i, ∑n i=0 ai(Ġ)x n−i and ∑n i=0 si(G ′)xn−i, respectively, then ai(G) = ai(Ġ) = si(G ′) (mod 2), for 1 ≤ i ≤ n, as AG = AĠ = SG′ (mod 2). Since si(G ′) = 0 for i odd, this in particular means that ai(G) and ai(Ġ) are even, whenever i is odd. Moreover, we have the following consequence. Theorem 4.3. For a graph G, let Ġ (resp. G′) consist of all signed graphs (oriented graphs) having G as the underlying graph. If there is at least one Γ ∈ {G} ∪ Ġ ∪ G′, such that the determinant of Γ is odd, then G, all signed graphs of Ġ and all oriented graphs of G′ are non-singular (i.e. their determinant is non-zero). Proof. This result follows from the previous observation applied to a0(G), a0(Ġ) and s0(G ′). Indeed, if for example a0(Ġ) is odd, then a0(Ġ) = a0(G) = 1 (mod 2), and the same holds for the elements of Ġ ∪ G′ in the role of G. 5 Eigenspaces Let E(λ) and E(−λ) be the eigenspaces of eigenvalues λ ( ̸= 0) and−λ of a bipartite signed graph Ġ, and E(iλ), E(−iλ) the eigenspaces of iλ and −iλ belonging to the spectrum of an associated oriented graph G′. Then E(iλ) ∪ E(−iλ) is spanned (in Cn) by the union of bases of E(λ) and E(−λ). Indeed, since −S2G′ = A2Ġ, the eigenspace of λ 2 for A2 Ġ coincides with the eigenspace of−λ2 for S2, and thus the former eigenspace is spanned by the union of eigenbases of λ and −λ (for AĠ), and the latter one is spanned by the union of eigenbases of iλ and −iλ (for SG′ ). By the same reasoning, if 0 is an eigenvalue of Ġ, then it has the same eigenspace for Ġ and G′. However, we can say more. We first note the following, probably known, result. If x is an eigenvector associated with the eigenvalue λ (̸= 0) of a bipartite signed graph Ġ, then by negating the entries on one colour class, we get an eigenvector associated with −λ. Indeed, with an appropriate vertex labelling, the adjacency matrix has the form AĠ = ( O B B⊺ O ) (5.1) If x = ( x1 x2 ) , then AĠ ( −x1 x2 ) = ( Bx2 −B⊺x1 ) = −λ ( −x1 x2 ) , as desired. If AĠ has no the previous form, the eigenvectors are permutationally equal, and the result follows. Henceforth, we assume that the adjacency matrix of a bipartite signed graph is as in (5.1). Theorem 5.1. Let xj = ( x1 x2 ) and x−j = ( −x1 x2 ) be eigenvectors associated with dis- tinct eigenvalues λj and −λj of a bipartite signed graph Ġ, and let, with a consistent vertex labelling, G′ be an associated oriented graph. Then xj + ix−j and xj − ix−j are eigenvectors associated with iλj and −iλj . 560 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 Proof. If the adjacency matrix of Ġ has the form (5.1), then the skew adjacency matrix of G′ is SG′ = ( O B −B⊺ O ) . Indeed, in the adjacency matrix of the underlying graph G, B has no negative entries, and reversing the sign of an edge uv changes the sign of both auv, avu in AĠ and both suv, svu in SG′ . We compute SG′(xj ± ix−j) = ( Bx2 −B⊺x1 ) ± i ( Bx2 B⊺x1 ) = ( λjx1 −λjx2 ) ± i ( λjx1 λjx2 ) = ± iλj (( x1 x2 ) ± i ( −x1 x2 )) = ±iλj(xj ± ix−j), as desired. The previous theorem tells us that, in the bipartite case, the eigenspaces of iλj and−iλj of G′ contain vectors of the form yij = ( x1 x2 ) + i ( −x1 x2 ) and y−ij = yij = ( x1 x2 ) − i ( −x1 x2 ) , respectively. Conversely, if these eigenvectors are given, then the eigenvectors associated with λj and −λj of Ġ are easily computed. We proceed with the non-bipartite case. Theorem 5.2. Let G′ be a non-bipartite oriented graph and y and y eigenvectors as- sociated with the eigenvalues iλj and −iλj , respectively. Then the pair ( y y ) , ( y −y ) (resp. ( y y ) , ( y −y ) ) is associated with iλj (resp. −iλj) in the bipartite double bd(G′), where Sbd(G′) = ( 0 1 1 0 ) ⊗ SG′ . Let Ḣ be a signed graph associated with bd(G′), and k the multiplicity of λj in the spectrum of Ḣ . The first (resp. second) pair is spanned by xjℓ + ix−jℓ (resp. xjℓ − ix−jℓ), for 1 ≤ ℓ ≤ k, where xjℓ and x−jℓ span the eigenspaces of λj and −λj , respectively. Proof. Since Sbd(G′) is the tensor product, the eiegnvectors of iλj for bd(G′) are (1, 1)⊺⊗ y and (1,−1)⊺ ⊗ y, and similarly for −iλj . This establishes the proof of the first part of the statement. By Theorem 5.1, the eigenvectors of iλj for bd(G′) are xjℓ + ix−jℓ, for 1 ≤ ℓ ≤ k. They are obviously linearly independent (because xjℓ and x−jℓ are). Since the multiplicity of iλj in the spectrum of bd(G′) is k (cf. Theorem 2.1(ii)), its geometric multiplicity is at most k and the desired conclusion follows. We proceed with an application in control theory. The following differential equation is the standard model of multi-agent single-input linear control systems: dx dt = Mx+ bu, (5.2) where the scalar u = u(t) is the control input, M ∈ Rn × Rn and x,b ∈ Rn. The system is controllable if for any x⋆ and time t⋆, there exists a control function u(t), 0 ≤ t ≤ t⋆, Z. Stanić: Spectra of signed graphs and related oriented graphs 561 such that the solution of the differential equation gives x⋆ = x(t⋆) irrespective of x(0). There are many controllable criteria, one of which is the Popov-Belevitch-Hautus Test to be found in any of [6, 11]. Accordingly, the system is controllable if and only if there is no z ∈ Cn \ {0} such that z∗M = λz∗ and z∗b = 0, where ∗ denotes the complex conjugate. Theorem 5.3. Let A be the adjacency matrix of a bipartite signed graph and S the skew adjacency matrix of an associated oriented graph. If the system (5.2) with M = A is controllable, then it is controllable with M = S. Proof. Since the system with M = A is controllable, we have x⊺b ̸= 0 for every eigen- vector of A. Moreover A has no repeated eigenvalues. Indeed, if x1 and x2 are linearly independent eigenvectors associated with the same eigenvalue, with x⊺1b = b1 ̸= 0 and x⊺2b = b2 ̸= 0, then (b2x1 − b1x2)⊺b = 0. As z∗S = λz∗ is equivalent to Sz = λz, and z is an eigenvector for S if and only if z∗ is an eigenvector for the same matrix, we have to show that for every eigenvector y of S, y∗b = 0 holds. Since A and S share the same eigenspace for 0, the claim holds for eigenvectors asso- ciated with 0 (if any). Every eigenvector associated with iλj has the form z(xj + ix−j), where z = z1 + iz2 ̸= 0, and xj and x−j are as in the formulation of Theorem 5.1. As before, we may take x⊺jb = b1 ̸= 0 and x ⊺ −jb = b2 ̸= 0. Then (z(xj + ix−j)) ∗b = ( z1xj − z2x−j + i(z1x−j + z2xj) )∗ b =(z1xj − z2x−j)⊺b− i(z1x−j + z2xj)⊺b = z1b1 − z2b2 + i(z1b2 + z2b1). Equating with 0, we obtain z1 = b2b1 z2 and z2 = − b2 b1 z1, which leads to the impossible scenario z1 = z2 = 0. Hence, the system with M = S is controllable. 6 Energy We start with the following result concerning signed graphs. There is a similar result for oriented graphs reported in [1]. Theorem 6.1. Let E(Ġ) be the energy of a signed graph Ġ having n vertices, m edges, average vertex degree d = 2mn and eigenvalues λ1, λ2, . . . , λn. Then√√√√2(m+ (n 2 )( n∏ i=1 |λi| ) 2 n ) ≤ E(Ġ) ≤ n √ d. (6.1) Both equalities hold if and only if Ġ is either edgeless or has exactly two distinct eigenval- ues and these eigenvalues are equal in absolute value. Proof. The proof of inequalities of (6.1) is an imitation of the proof of [1, Theorem 2.5] (concerning oriented graphs). Namely, both follow from the next chain of inequalities and 562 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 equalities: 2 ( m+ ( n 2 )( n∏ i=1 |λi| ) 2 n ) ≤ n∑ i=1 |λi|2 + 2 ∑ 1≤i E(Ġ) is illustrated in Figure 2. Motivated by these experiments, we formulate the following problems. Problem 6.2. Determine graphs G such that E(G) ≤ E((G, σ̇)) holds for every signature σ̇ defined on the edge set of G. Problem 6.3. Determine (or, at least, characterize) signed graphs Ġ with E(Ġ) < E(G). We proceed with oriented graphs. If G′ is associated with Ġ, then E(G′) = E(Ġ), by Theorem 2.1(i). If Ḣ is as in Theorem 2.1(ii), then E(G′) = E(Ḣ)2 . Accordingly, if the underlying graph G is regular of degree r, then E(G′) = n √ r (the upper bound (6.1)) if and only if G′ has exactly two eigenvalues (complex conjugates); this is established in [1], as well. Some examples are constructed in the mentioned reference. Here we note that the search on regular oriented graphs attaining the upper bound (6.1) is reduced to the search on signed graphs with the same spectral property: An associate of a bipartite signed graph with E(Ġ) = n √ r has the required spectral property, and if it figures as a bipartite double, then a corresponding constituent has the same property. In this context it is worth mentioning that, on the basis of the results of [10, 16], all oriented graphs with n vertices and spectrum [i2n/2, (−i2)n/2] are determined in [15] (their energy attains 2n); they include infinite families of both bipartite and non-bipartite oriented graphs. More examples can be extracted from known signed graphs with two eigenvalues obtained in the foregoing references. We also observe that an oriented graph shares the energy with its underlying graph G whenever it is associated with a balanced signed graph. In this context, we point out that, due to Corollary 3.3, Ḣ of Theorem 2.1(ii) is never balanced. It also shares the energy with G whenever it is associated with a signed graph that switches to −G. This section is concluded with the following result. A conference matrix A is an n× n matrix with diagonal entries 0 and off-diagonal entries ±1, satisfying A⊺A = (n− 1)I . Theorem 6.4. An oriented graph G′ (resp. a signed graph Ġ) with n vertices attains E(G′) = n √ n− 1 (E(Ġ) = n √ n− 1) if and only if its adjacency matrix is the skew- symmetric conference matrix (symmetric conference matrix). Proof. If E(G′) = n √ n− 1, by Theorems 2.1(ii) and 6.1, G′ is regular of degree n − 1 and its spectrum is [ i √ n− 1n/2,−(i √ n− 1)n/2 ] . Considering the minimal polynomial of AG′ , we obtain A2G′ = (n−1)I , which means that AG′ is a skew-symmetric conference 564 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 matrix. Conversely, if AG′ is a skew-symmetric conference matrix then, by definition, we have A2G′ = (n− 1)I , which gives E(G′) = n √ n− 1. The proof for Ġ is analogous. The previous result addresses the open problem (6) of [1, Section 6] related to the existence of skew-symmetric conference matrices that do not give the maximum energy of the corresponding oriented graph. It also partially addresses the problem (2) of the same reference related to determination of oriented graphs with maximum energy, as it gives their characterization via the matrix structure. However, their determination remains a difficult research problem, since it is equivalent to the complete determination of skew-symmetric conference matrices. 7 Notes on hypercubes Due to [22, Theorem 2.1(iv)] (see also [21, Theorem 2]) a signed graph is balanced if its cycle basis has all positive cycles. Since the induced quadrangles of a signed r-cube contain a cycle basis, we arrive at the following result. Lemma 7.1 (cf. [22, Theorem 2.1(iv)]). For r ≥ 2, a signed r-cube Q̇r is balanced if and only if it has no negative quadrangles. In [2, 3, 20], the authors gave several algorithms which iteratively construct an oriented r-cube such that its energy is either equal to the energy of the underlying graph or attains the upper bound (6.1). (This upper bound reduces to n √ r.) These algorithms are useful since they give explicit orientations for which the previous settings occur. In this context we offer the following contribution (see also the text below the theorem). Theorem 7.2. The following statements hold true: (i) A signed r-cube without negative quadrangle is associated with an oriented r-cube whose energy is equal to the energy of its underlying r-cube. (ii) A signed r-cube is associated with an oriented r-cube whose energy attains the upper bound of (6.1) if and only if it has no positive quadrangle. Proof. (i): If every quadrangle (if any) in Q̇r is positive then Q̇r is balanced, by Lemma 7.1, and therefore it switches to the underlying graph Qr. Consequently, Q̇r and Qr have the same energy. By Theorem 2.1, Q̇r shares the energy with an associated oriented cube, and the result follows. (ii): Since a signed graph and an associated oriented graph share the same energy, it is sufficient to show that the energy n √ r is exclusive to a signed cube without positive quadrangles. First, such a cube has this energy (as mentioned in the previous section). Conversely, assume that, for r ≥ 3, a signed r-cube with the adjacency matrix A has a positive quadrangle, say uvwz (where the vertices are in the natural order). It follows that the (u,w)-entry of A2 is 2, and so A2 is not a multiple of the identity matrix, but then the spectrum of A deviates the equality condition of Theorem 6.1, as follows by considering the minimal polynomial. In other words, to construct an oriented r-cube whose energy is E(Qr) (resp. n √ r), it is sufficient to take any signed r-cube without negative quadrangle (without positive quadrangle), and construct an oriented associate. Z. Stanić: Spectra of signed graphs and related oriented graphs 565 ORCID iDs Zoran Stanić https://orcid.org/0000-0002-4949-4203 References [1] C. Adiga, R. Balakrishnan and W. So, The skew energy of a digraph, Linear Algebra Appl. 432 (2010), 1825–1835, doi:10.1016/j.laa.2009.11.034, https://doi.org/10.1016/ j.laa.2009.11.034. [2] A. Anuradha and R. Balakrishnan, Skew spectrum of the Cartesian product of an oriented graph with an oriented hypercube, in: R. B. Bapat, S. J. Kirkland, K. M. Prasad and S. Pun- tanen (eds.), Combinatorial Matrix Theory and Generalized Inverses of Matrices, Springer, New Delhi, pp. 1–12, 2013, doi:10.1007/978-81-322-1053-5\ 1, https://doi.org/10. 1007/978-81-322-1053-5_1. [3] A. Anuradha, R. Balakrishnan, X. Chen, X. Li, H. Lian and W. So, Skew spectra of oriented bipartite graphs, Electron. J. Comb. 20 (2013), Paper 18, 12 pp., doi:10.37236/3331, https: //doi.org/10.37236/3331. [4] F. Belardo and S. K. Simić, On the Laplacian coefficients of signed graphs, Linear Algebra Appl. 475 (2015), 94–113, doi:10.1016/j.laa.2015.02.007, https://doi.org/10.1016/ j.laa.2015.02.007. [5] M. Cavers, S. M. Cioabă, S. Fallat, D. A. Gregory, W. H. Haemers, S. J. Kirkland, J. J. McDonald and M. Tsatsomeros, Skew-adjacency matrices of graphs, Linear Algebra Appl. 436 (2012), 4512–4529, doi:10.1016/j.laa.2012.01.019, https://doi.org/10.1016/ j.laa.2012.01.019. [6] C.-T. Chen, Linear System Theory and Design, Oxford University Press, New York, 1998. [7] D. Cui and Y. Hou, On the skew spectra of Cartesian products of graphs, Electron. J. Comb. 20 (2013), Paper 19, 13 pp., doi:10.37236/2864, https://doi.org/10.37236/2864. [8] G. R. W. Greaves and Z. Stanić, Signed (0, 2)-graphs with few eigenvalues and a symmetric spectrum, J. Comb. Des. 30 (2022), 332–353, doi:10.1002/jcd.21828, https://doi.org/ 10.1002/jcd.21828. [9] Y. Hou and T. Lei, Characteristic polynomials of skew-adjacency matrices of oriented graphs, Electron. J. Comb. 18 (2011), Paper 156, 12 pp., doi:10.37236/643, https://doi.org/ 10.37236/643. [10] Y. Hou, Z. Tang and D. Wang, On signed graphs with just two distinct adjacency eigenvalues, Discrete Math. 342 (2019), 111615, 8 pp., doi:10.1016/j.disc.2019.111615, https://doi. org/10.1016/j.disc.2019.111615. [11] J. Klamka, Controllability of Dynamical Systems, volume 48 of Mathematics and its Applica- tions (East European Series), Kluwer Academic Publishers Group, Dordrecht; PWN—Polish Scientific Publishers, Warsaw, 1991. [12] J. McKee and C. Smyth, Integer symmetric matrices having all their eigenvalues in the in- terval [−2, 2], J. Algebra 317 (2007), 260–290, doi:10.1016/j.jalgebra.2007.05.019, https: //doi.org/10.1016/j.jalgebra.2007.05.019. [13] Z. Stanić, On strongly regular signed graphs, Discrete Appl. Math. 271 (2019), 184–190, doi: 10.1016/j.dam.2019.06.017, https://doi.org/10.1016/j.dam.2019.06.017. [14] Z. Stanić, A decomposition of signed graphs with two eigenvalues, Filomat 34 (2020), 1949– 1957, doi:10.2298/FIL2006949S, https://doi.org/10.2298/FIL2006949S. 566 Ars Math. Contemp. 24 (2024) #P3.10 / 553–566 [15] Z. Stanić, Oriented graphs whose skew spectral radius does not exceed 2, Linear Algebra Appl. 603 (2020), 359–367, doi:10.1016/j.laa.2020.06.018, https://doi.org/10.1016/j. laa.2020.06.018. [16] Z. Stanić, Spectra of signed graphs with two eigenvalues, Appl. Math. Comput. 364 (2020), 124627, 9 pp., doi:10.1016/j.amc.2019.124627, https://doi.org/10.1016/j.amc. 2019.124627. [17] Z. Stanić, Signed graphs with two eigenvalues and vertex degree five, Ars Math. Contemp. 22 (2022), Paper No. 10, 13 pp., doi:10.26493/1855-3974.2329.97a, https://doi.org/10. 26493/1855-3974.2329.97a. [18] Z. Stanić, Some relations between the skew spectrum of an oriented graph and the spectrum of certain closely associated signed graphs, Rev. Un. Mat. Argentina 63 (2022), 41–50, doi: 10.33044/revuma.1914, https://doi.org/10.33044/revuma.1914. [19] Z. Stanić, Relations between the skew spectrum of an oriented graph and the spectrum of an associated signed graph, Linear Algebra Appl. 676 (2023), 241–250, doi:10.1016/j.laa.2023. 07.015, https://doi.org/10.1016/j.laa.2023.07.015. [20] G.-X. Tian, On the skew energy of orientations of hypercubes, Linear Algebra Appl. 435 (2011), 2140–2149, doi:10.1016/j.laa.2011.04.007, https://doi.org/10.1016/ j.laa.2011.04.007. [21] T. Zaslavsky, Characterizations of signed graphs, J. Graph Theory 5 (1981), 401–406, doi: 10.1002/jgt.3190050409, https://doi.org/10.1002/jgt.3190050409. [22] T. Zaslavsky, Signed graphs, Discrete Appl. Math. 4 (1982), 47–74, doi:10.1016/ 0166-218X(82)90033-6, https://doi.org/10.1016/0166-218X(82)90033-6. 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). xxi 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 xxii Printed in Slovenia by IME TISKARNE