Volume 20, Number 2, Spring/Summer 2021, Pages 171–288 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 Some algebraic properties of Sierpiński-type graphs Mohammad Farrokhi Derakhshandeh Ghouchan, Ebrahim Ghorbani, Hamid Reza Maimani, Farhad Rahimi Mahid . . . . . . . . . . . . . . . . 171 The chromatic index of strongly regular graphs Sebastian M. Cioabă, Krystal Guo, Willem H. Haemers . . . . . . . . . . . 187 A simple and elementary proof of Whitney’s unique embedding theorem Gunnar Brinkmann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195 The average genus for bouquets of circles and dipoles Jinlian Zhang, Xuhui Peng, Yichao Chen . . . . . . . . . . . . . . . . . . . 199 Well-totally-dominated graphs Selim Bahadır, Tınaz Ekim, Didem Gözüpek . . . . . . . . . . . . . . . . 209 Nordhaus-Gaddum type inequalities for the distinguishing index Monika Pilśniak . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 Closed formulas for the total Roman domination number of lexicographic product graphs Abel Cabrera Martínez, Juan Alberto Rodríguez-Velázquez . . . . . . . . . 233 Wiener-type indices of Parikh word representable graphs Nobin Thomas, Lisa Mathew, Sastha Sriram, K. G. Subramanian . . . . . . 243 A double Sylvester determinant Darij Grinberg . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261 Boundary-type sets of strong product of directed graphs Bijo S. Anand, Manoj Changat, Prasanth G. Narasimha-Shenoi, Mary Shalet Thottungal Joseph . . . . . . . . . . . . . . . . . . . . . . . . 275 Volume 20, Number 2, Spring/Summer 2021, Pages 171–288 xiii ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 171–186 https://doi.org/10.26493/1855-3974.2199.97e (Also available at http://amc-journal.eu) Some algebraic properties of Sierpiński-type graphs* Mohammad Farrokhi Derakhshandeh Ghouchan Department of Mathematics, Institute for Advanced Studies in Basic Sciences (IASBS), and the Center for Research in Basic Sciences and Contemporary Technologies, IASBS, P. O. Box 45137-66731, Zanjan, Iran Ebrahim Ghorbani † Department of Mathematics, K. N. Toosi University of Technology, P. O. Box 16765-3381, Tehran, Iran Hamid Reza Maimani , Farhad Rahimi Mahid Department of Basic Sciences, Shahid Rajaee Teacher Training University, P. O. Box 16783-163, Tehran, Iran Received 16 December 2019, accepted 4 November 2020, published online 21 October 2021 Abstract This paper deals with some algebraic properties of Sierpiński graphs and a family of regular generalized Sierpiński graphs. For the family of regular generalized Sierpiński graphs, we obtain their spectrum and characterize those graphs that are Cayley graphs. As a by-product, a new family of non-Cayley vertex-transitive graphs, and consequently, a new set of non-Cayley numbers are introduced. We also obtain the Laplacian spectrum of Sierpiński graphs in some particular cases, and make a conjecture on the general case. Keywords: Sierpiński graph, spectrum, Laplacian, Cayley graph, non-Cayley number. Math. Subj. Class. (2020): 05C50, 05C25, 05C75 *The authors would like to thank anonymous referees for constructive comments which led to improvement of the presentation of the paper. The second author carried this work during a Humboldt Research Fellowship at the University of Hamburg. He thanks the Alexander von Humboldt-Stiftung for financial support. †Corresponding author. E-mail addresses: farrokhi@iasbs.ac.ir, m.farrokhi.d.g@gmail.com (Mohammad Farrokhi Derakhshandeh Ghouchan), ghorbani@kntu.ac.ir (Ebrahim Ghorbani), maimani@ipm.ir (Hamid Reza Maimani), farhad.rahimi@sru.ac.ir (Farhad Rahimi Mahid) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 172 Ars Math. Contemp. 20 (2021) 171–186 1 Introduction Sierpiński-type graphs show up in a wide range of areas; for instance, physics, dynamical systems, probability, and topology, to name a few. The Sierpiński gasket graphs form one of the most significant families of such graphs that are obtained by a finite number of iterations that give the Sierpiński gasket in the limit. Several more families of Sierpiński- type graphs have been introduced and studied in the literature (see Barrière, Comellas, and Dalfó [1] and Hinz, Klavžar, and Zemljič [6]). In this paper, we deal with two families of them, as described below. For positive integers n, k, the Sierpiński graph S(n, k) is defined with vertex set [k]n, where [k] := {1, . . . , k}, and two different vertices (u1, . . . , un) and (v1, . . . , vn) are ad- jacent if and only if there exists a t ∈ [n] such that • ui = vi for i = 1, . . . , t− 1, • ut ̸= vt, • uj = vt and vj = ut for j = t+ 1, . . . , n. For instance, the graphs S(3, 3) and S(2, 4) are depicted in Figure 1. 222 223 232 233 322 323 332 333 221 231 321 331 212 213 312 313 211 311 122 123 132 133 121 131 112 113 111 44 43 34 33 41 42 31 32 14 13 24 23 11 12 21 22 Figure 1: The Sierpiński graphs S(3, 3) (left) and S(2, 4) (right). Sierpinśki graphs S(n, k) were introduced in Klavžar and Milutinović [8]. The graph S(n, 3) is indeed isomorphic to the graph of the Tower of Hanoi with n disks. The graph S(n, k) has kn − k vertices of degree k and k vertices of degree k − 1 that are (i, . . . , i) for i ∈ [k]. These vertices are called the extreme vertices. In addition to S(n, k), we consider a ‘regularization’ of them as another family of Sierpiński-type graphs. The graphs S++(n, k), introduced in Klavžar and Mohar [9], are defined as follows. The graph S++(1, k) is the complete graph Kk+1. For n ≥ 2, S++(n, k) is the graph obtained from the disjoint union of k + 1 copies of S(n− 1, k) in which the extreme vertices in distinct copies of S(n− 1, k) are connected as the complete graph Kk+1. See Figure 2 for an illustration of S++(3, 3). Many properties of Sierpiński-type graphs, including those of S(n, k) and S++(n, k), have been studied in the literature, for a survey see Hinz, Klavžar, and Zemljič [6]. In M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 173 Figure 2: The graph S++(3, 3). this paper, we investigate some algebraic properties of the two families of graphs, namely the spectrum and the property of being a Cayley graph. More precisely, in Section 2, we determine the spectrum of the graphs S++(n, k). The Laplacian spectrum of S(n, k) is already known for k = 2, 3. We establish the case n = 2, in Section 3, and make a conjecture on the Laplacian spectrum of S(n, k) in general. We also characterize the graphs S++(n, k) that are Cayley graphs in Section 4. As a by-product, a new family of non-Calyley vertex-transitive graphs are obtained. From this result, we conclude a new set of square-free non-Cayley numbers in Section 5, and we discuss its distribution. 2 Spectrum of S++(n, k) Let Γ be a simple graph with vertex set V (Γ) = {v1, . . . , vn} and edge set E(Γ). Its adjacency matrix A(Γ) = [aij ] is an n × n symmetric matrix with aij = 1 if vi and vj are adjacent, and aij = 0 otherwise. The multi-set of the eigenvalues of A(Γ) is called the spectrum of Γ. In this section, we determine the spectrum of S++(n, k). As we will see, the recursive structure of these Sierpiński-type graphs also shows up in their spectrum. We first recall some basic facts. The incidence matrix of a graph Γ is a 0-1 matrix X(Γ) = [xve], with rows indexed by the vertices and columns indexed by the edges of Γ, where xve = 1 if the vertex v is an endpoint of the edge e. For a graph Γ, L(Γ) denotes the line graph of Γ, in which V (L(Γ)) corresponds with E(Γ), and two vertices of L(Γ) are adjacent if and only if they have a common vertex as edges of Γ. The subdivision graph S(Γ) of Γ is the graph obtained by inserting a new vertex into every edge of Γ. It is easy to verify that X(Γ)⊤X(Γ) = 2I +A(L(Γ)), (2.1) and, moreover, if Γ is k-regular, then X(Γ)X(Γ)⊤ = kI +A(Γ). (2.2) 174 Ars Math. Contemp. 20 (2021) 171–186 The following lemma gives a recursive relation for the graphs S++(n, k). Lemma 2.1. The graph S++(n+ 1, k) is isomorphic to L(S(S++(n, k))). Proof. Let k be fixed. The graph Γn := S++(n, k) can be obtained by the union of S(n, k) and S(n − 1, k) by adding a matching between the extreme vertices of the two graphs. If we consider {0} × [k]n−1 as the vertex set of S(n− 1, k) (to make them compatible with the length n of the vertices of S(n, k)), then ({0} ∪ [k])× [k]n−1 is the vertex set of Γn. It follows that any edge e = {u,v} of Γn is of one of the following types: (1) u = (u1, . . . , ur, u, v, . . . , v), v = (u1, . . . , ur, v, u, . . . , u) for some r ≤ n− 2 and u ̸= v; (2) u = (u1, . . . , un−1, u), v = (u1, . . . , un−1, v) with u ̸= v; (3) u = (0, u, . . . , u), v = (u, u, . . . , u). Each e = {u,v} ∈ E(Γn) is divided into two new edges eu and ev in S(Γn), where we assume that u ∈ eu and v ∈ ev. We define a map ψ : E(S(Γn)) → ({0} ∪ [k]) × [k]n based on the type of e as follows: (i) If e is of type (1) or (2), then ψ(eu) = (u, v) and ψ(ev) = (v, u); (ii) If e is of type (3), then ψ(eu) = (u, u) and ψ(ev) = (v, u). It is easily seen that ψ is a one-to-one map. We show that ψ is an isomorphism from L(S(Γn)) to Γn+1. Let e and e′ be two edges that share a vertex x of S(Γn). If x = (x1, . . . , xn) is an ‘old’ vertex of S(Γn), then ψ(e) = (x, y) and ψ(e′) = (x, z) for some y ̸= z. Then, it is clear that ψ(e) and ψ(e′) are adjacent in Γn+1. If x is a ‘new’ vertex of S(Γn), then from (i) and (ii), it is clear that ψ(e) and ψ(e′) are adjacent in Γn+1. This shows that ψ is indeed a one-to-one homomorphism. As L(S(Γn)) and Γn+1 have the same number of edges, it follows that ψ is an isomorphism. We recall that if A is a non-singular square matrix, then∣∣∣∣A BC D ∣∣∣∣ = |A| · ∣∣D − CA−1B∣∣ , (2.3) where | · | denotes the determinant of a matrix. Also, recall that if M is a p× q matrix, then |xI −MM⊤| = xp−q|xI −M⊤M |. (2.4) (Note that (2.4) might not be valid if p ≤ q and x = 0, but this has no effect in our argument since two polynomials that agree in all but finitely many points, agree everywhere.) Let f(x) = x2 + (2− k)x− k, (2.5) and let f j(x) denote the polynomial of degree 2j obtained by j times composition of f with itself. As a convention, we let f0(x) = x. We now give the main result of this section. M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 175 Theorem 2.2. Let k be an integer and Pn(x) denote the characteristic polynomial of the adjacency matrix of S++(n, k). Then, Pn satisfies the recursion relation Pn(x) = (x(x+ 2)) kn−2((k2)−1) Pn−1(f(x)), n ≥ 2, (2.6) with P1(x) = (x − k)(x + 1)k. Moreover, for n ≥ 2, the spectrum of S++(n, k) consists of the following eigenvalues: (i) k with multiplicity 1, (ii) the zeros of fn−1(x) + 1 each with multiplicity k, (iii) the zeros of f j(x) each with multiplicity kn−2−j (( k 2 ) − 1 ) for j = 0, 1, . . . , n− 2, (iv) the zeros of f j(x)+2 each with multiplicity kn−2−j (( k 2 ) − 1 ) +1 for j = 0, 1, . . . , n− 2. Proof. Let Γn := S++(n, k). Suppose that X and Y are the incidence matrices of Γn−1 and S(Γn−1), respectively. By Lemma 2.1, Γn is isomorphic to L(S(Γn−1)). It follows that Y Y ⊤ = [ kIp X X⊤ 2Iq ] , where the matrix is divided according to the partition of the vertices into p = kn−1+kn−2 ‘old’ vertices of Γn−1 and q = 12 (k n + kn−1) ‘new’ vertices (which have degree 2) added to Γn−1 to obtain S(Γn−1). Therefore, from (2.3),∣∣xI − Y Y ⊤∣∣ = |(x− k)Ip| · ∣∣∣(x− 2)Iq −X⊤ ((x− k)Ip)−1X∣∣∣ = (x− k)p ∣∣∣∣(x− 2)Iq − 1x− kX⊤X ∣∣∣∣ = (x− k)p−q ∣∣(x− 2)(x− k)Iq −X⊤X∣∣ = (x− k)p−q((x− 2)(x− k))q−p ∣∣(x− 2)(x− k)Ip −XX⊤∣∣ (by (2.4)) = (x− 2)q−p |((x− 2)(x− k)− k)Ip −A(Γn−1)| (by (2.2)) = (x− 2)q−pPn−1((x− 2)(x− k)− k). (2.7) On the other hand, by (2.1) and (2.2), we have Pn(x) = ∣∣(x+ 2)I2q − Y ⊤Y ∣∣ = (x+ 2)q−p ∣∣(x+ 2)Ip+q − Y Y ⊤∣∣ . Now, from (2.7) it follows that Pn(x) = (x(x+ 2)) q−pPn−1(x(x+ 2− k)− k), implying (2.6). To prove the second part of the theorem, note that as Γ1 = Kk+1, we have P1(x) = (x− k)(x+ 1)k. From (2.6), we conclude that P2(x) = (x(x+ 2)) (k2)−1(f(x)− k)(f(x) + 1)k, 176 Ars Math. Contemp. 20 (2021) 171–186 and since f(x)− k = (x+ 2)(x− k), (2.8) the assertion follows for n = 2. Now assume that n ≥ 3 and the assertion holds for n− 1. So, we have Pn−1(x) = (x− k) ( fn−2(x) + 1 )k n−3∏ j=0 ( f j(x) )mn−3−j ( f j(x) + 2 )1+mn−3−j , in which mi = ki (( k 2 ) − 1 ) . It follows that Pn−1(f(x)) = (f(x)− k) ( fn−1(x) + 1 )k n−2∏ j=1 ( f j(x) )mn−2−j ( f j(x) + 2 )1+mn−2−j . This, together with (2.6) and (2.8), implies the result. Remark 2.3. It is straightforward to see that the zeros of f j(x) and f j(x) + 2 for j = 1 are 12 (k − 2 ± √ k2 + 4) and 12 (k − 2 ± √ k2 − 4), respectively, and for j ≥ 2 are of the form 1 2 (k − 2)± 1 2 √ k(k + 2)± 2 √ k(k + 2)± 2 √ · · · ± 2 √ k2 + 4 and 1 2 (k − 2)± 1 2 √ k(k + 2)± 2 √ k(k + 2)± 2 √ · · · ± 2 √ k2 − 4, respectively, each of them consisting of j nested radicals in iterative forms. Moreover, the zeros of fn−1(x) + 1 are − 1, k − 1, 1 2 (k − 2)± 1 2 √ k2 + 4k, 1 2 (k − 2)± 1 2 √ k(k + 2)± 2 √ k2 + 4k, . . . , 1 2 (k − 2)± 1 2 √ k(k + 2)± 2 √ k(k + 2)± 2 √ · · · ± 2 √ k2 + 4k, where the last one consists of n− 2 nested radicals. 3 Laplacian spectrum of S(n, k) For a graph Γ, the matrix L(Γ) = D(Γ)−A(Γ) is the Laplacian matrix of Γ, where D(Γ) is the diagonal matrix of vertex degrees. The multi-set of eigenvalues of L(Γ) is called the Laplacian spectrum of Γ. In this section, we deal with the Laplacian spectrum of S(n, k). This is trivial for n = 1 or k = 1. For k = 2, 3, the Laplacian spectrum of S(n, k) is already known (see Remark 3.4 below). We establish the case n = 2, and put forward a conjecture explicitly describing the Laplacian spectrum of S(n, k) in general. Let Eij be a k × k matrix in which all entries are 0, except the (i, j) entry that is 1. Consider the k2 × k2 matrix C := k∑ i=1 k∑ j=1 (Eij ⊗ Eji), M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 177 where ‘⊗’ denotes the Kronecker product. The matrix C is called the commutation matrix. The main property of the commutation matrix (see Magnus and Neudecker [10]) is that it commutes the Kronecker product: for any k × k matrices M,N , C(M ⊗N)C = N ⊗M. Note that each row and each column of C corresponds with a pair (i, j) for 1 ≤ i, j ≤ k. Moreover, C is indeed a permutation matrix in which the only 1 entry in the row (i, j) is located at the column (j, i) for every 1 ≤ i, j ≤ k. For n = 1, the Laplacian spectrum of S(1, k) = Kk is { 0[1], k[k−1] } , where the superscripts indicate multiplicities. In the following theorem, we determine the Laplacian spectrum of S(2, k). Theorem 3.1. The Laplacian spectrum of S(2, k) is the following:{ 0[1], k[( k 2)], (k + 2)[( k−1 2 )], ( 1 2 (k + 2)± 1 2 √ k2 + 4 )[k−1]} . Proof. First, note that the graph S(2, k) consists of k copies ofKk together with a matching M of size ( k 2 ) ; exactly one edge for each pair of copies of Kk. Let L denote the Laplacian matrix of S(2, k), and L′ be the Laplacian matrix of the induced subgraph by the edges of M . It is seen that L = Q − B, where Q = L(kKk) + I and B = I − L′. Note that B is a permutation matrix with ( k 2 ) + k eigenvalues 1 and ( k 2 ) eigenvalues −1. Observe that Q has k eigenvalues 1 and k2 − k eigenvalues k + 1. We have the following bounds on the dimensions of intersections of the eigenspaces of B and Q: dim(E1(B) ∩ Ek+1(Q)) ≥ k2 − k + ( k 2 ) + k − k2 = ( k 2 ) , dim(E−1(B) ∩ Ek+1(Q)) ≥ k2 − k + ( k 2 ) − k2 = ( k 2 ) − k, in which Eλ denotes the eigenspace corresponding to the eigenvalue λ. For x ∈ E1(B) ∩ Ek+1(Q), we have Lx = kx and for x ∈ E−1(B)∩Ek+1(Q), Lx = (k+2)x. This means that L has eigenvalues k and k+2 with multiplicities at least ( k 2 ) and ( k 2 ) −k, respectively. We also have Q = Ik ⊗ ((k + 1)Ik − Jk), (3.1) and from the eigenvalues of Q, Q2 − (k + 2)Q+ (k + 1)I = O. (3.2) Coming back to B, for each of the extreme vertices (1, 1), . . . , (k, k) of S(2, k), there is a 1 on all the entries of the diagonal of B. The off-diagonal 1’s correspond with the edges of M . By the definition of S(2, k), the edges of M connect the vertices (i, j) and (j, i) for i ̸= j. It turns out that B is the commutation matrix, and thus BQB = ((k + 1)Ik − Jk)⊗ Ik. (3.3) The right sides of (3.1) and (3.3) commute, and so BQBQ = QBQB. 178 Ars Math. Contemp. 20 (2021) 171–186 Next, we see that (L2 − (k + 2)L+ kI)(QB −BQ) = ((Q−B)2 − (k + 2)(Q−B) + kI)(QB −BQ) = ( Q2 − (k + 2)Q+ kI +B2 + (k + 2)B −QB −BQ ) (QB −BQ) = ((k + 2)B −QB −BQ) (QB −BQ) (by (3.2) and since B2 = I) = Q2 − (k + 2)Q−B ( Q2 − (k + 2)Q ) B − (QB)2 + (BQ)2 = (BQ)2 − (QB)2 = O. The above equality shows that every vector in the column space of QB − BQ is an eigenvector forLwith eigenvalue λ, where λ2−(k+2)λ+k = 0. To obtain the multiplicity of such λ, we compute the rank of QB −BQ: rank(QB −BQ) = rank(Q−BQB) = rank(Ik ⊗ ((k + 1)Ik − Jk)− ((k + 1)Ik − Jk)⊗ Ik) = rank(Jk ⊗ Ik − Ik ⊗ Jk) = 2k − 2. (3.4) To show (3.4), suppose P is a k × k matrix whose first column is 1√ k (1, . . . , 1)⊤ and that PP⊤ = Ik. Then, Jk = P (kE11)P⊤, and so Jk ⊗ Ik − Ik ⊗ Jk = (P ⊗ P ) ( (kE11 ⊗ Ik)− (Ik ⊗ kE11) ) (P⊤ ⊗ P⊤). Since (kE11 ⊗ Ik) − (Ik ⊗ kE11) is a diagonal matrix having precisely 2k − 2 non-zero entries in the columns 2, 3, . . . , k, k + 1, 2k + 1, 3k + 1, . . . , (k − 1)k + 1, (3.4) follows. As x2 − (k + 2)x + k is an irreducible polynomial, each of its roots is an eigenvalue of L with multiplicity at least k − 1. The matrix L has a 0 eigenvalue. Thus, we have obtained so far ( k 2 ) + ( k 2 ) − k + 2(k − 1) + 1 = k2 − 1 eigenvalues of L. As the sum of the eigenvalues of L is twice the number of edges of S(2, k), it follows that the remaining eigenvalue is k + 2. So the proof is complete. Based on empirical evidence, we put forward the following conjecture. Conjecture 3.2. For n, k ≥ 2, the Laplacian spectrum of S(n, k) consists of the following eigenvalues: (i) 0 with multiplicity 1. (ii) The zeros of f j(k − x), each with multiplicity 12 (k n−j − 2kn−j−1 + k) for j = 0, 1, . . . , n− 1, where f is given in (2.5). (iii) The zeros of f j(k − x) + 2, each with multiplicity 12 (k n−j−1 − 1)(k − 2) for j = 0, 1, . . . , n− 2. Remark 3.3. The zeros of f j(k−x) and f j(k−x)+2 for j = 1 are 12 (k+2± √ k2 + 4) and 12 (k + 2± √ k2 − 4), respectively, and for j ≥ 2 are of the form 1 2 (k + 2)± 1 2 √ k(k + 2)± 2 √ k(k + 2)± 2 √ · · · ± 2 √ k2 + 4 M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 179 and 1 2 (k + 2)± 1 2 √ k(k + 2)± 2 √ k(k + 2)± 2 √ · · · ± 2 √ k2 − 4, respectively, each of them consisting of j nested radicals. Remark 3.4. The graph S(n, 2) is the path graph on 2n vertices. Proposition 3.5 below shows that Conjecture 3.2 holds for S(n, 2). In Grigorchuk and Šunić [5], the spectrum of the Schreier graph Γn was determined. The graph Γn is, in fact, the graph obtained from S(n, 3) by adding a loop on each extreme vertex. By the way the adjacency matrix of A(Γn) is defined in [5], for each loop a 1 entry on the diagonal is considered, so that each row and column of A(Γn) has constant sum 3. It is then observed that the Laplacian spec- trum of S(n, 3) can be deduced from the spectrum of Γn, which agrees with Conjecture 3.2. In summary, Conjecture 3.2 holds for n = 2 and for k = 2, 3. It is known in the literature that the characteristic polynomial of the Laplacian matrix of a path can be expressed in terms of the Chebyshev polynomials. From this fact, for a path with 2n vertices, we obtain the iterated form according to Conjecture 3.2. For the sake of completeness, we give its complete argument here. Proposition 3.5. The characteristic polynomial of the Laplacian matrix of the path graph on 2n vertices is equal to x ∏n−1 j=0 g j(2− x), where g(x) = x2 − 2. Proof. Let ϕm be the characteristic polynomial of the Laplacian matrix of the path graph on m vertices. Let Tm and Um be the Chebyshev polynomials of degree m of the first and the second kind, respectively. Then Tm is the only polynomial satisfying Tm(cos θ) = cosmθ and Um(x) = sin((m + 1) arccosx)/ sin(arccosx) (Snyder [18]). From the identities given in Cvetković, Doob, and Sachs [2, p. 220], it follows that ϕm(x) = xUm−1(x/2−1). By successive use of the identity U2k−1(x) = 2Tk(x)Uk−1(x) (see [18, p. 98]), we get U2n−1(x) = 2 n−1T2n−1(x)T2n−2(x) · · ·T2(x)U1(x). Note that U1(x) = 2x and T2(x) = 2x2 − 1. It is seen that 2T2(x/2 − 1) = x2 − 4x + 2 = g(2 − x). This, together with the identity T2k(x) = T2(Tk(x)), implies that 2T2j (x/2− 1) = gj(2− x). The proof is now complete. 4 What S++(n, k) are Cayley graphs? Recall that a graph Γ is vertex-transitive if for any two vertices u, v of Γ, there exists an automorphism σ of Γ such that σ(u) = v. Let G be a group and C ⊂ G such that 1 ̸∈ C and c ∈ C implies that c−1 ∈ C. The Cayley graph Cay(G,C) with the group G and the ‘connection set’ C is the graph with vertex set G in which vertex u is connected to v if and only if vu−1 ∈ C. It is known that any Cayley graph is vertex-transitive. In the other way around, at least for small orders, it seems that the great majority of vertex-transitive graphs are Cayley graphs, see McKay and Praeger [12]. It is expected to continue to be this way for larger or- ders. In fact, it is conjectured in Praeger, Li, and Niemeyer [15] that most vertex-transitive graphs are Cayley graphs. In this section, we first determine what S++(n, k) are vertex- transitive and, then, classify S++(n, k) that are Cayley graphs. 180 Ars Math. Contemp. 20 (2021) 171–186 Proposition 4.1. The graph S++(n, k) is vertex-transitive if and only if either n ≤ 2 or k ≤ 2. Proof. We have S++(n, 1) ∼= K2, S++(1, k) ∼= Kk+1, and S++(n, 2) is the cycle graph on 2n−1 · 3 vertices, which are all vertex-transitive graphs. By Lemma 2.1, S++(2, k) is isomorphic to L(S(Kk+1)). In the graph S(Kk+1), the ‘new’ vertices are in one-to-one correspondence with 2-subsets of [k+1]. It is then easy to see that that any permutation of [k+ 1] induces an automorphism of S(Kk+1). Now, for a given pair of edges of S(Kk+1) which can be represented as e = {i, {i, j}} and e′ = {i′, {i′, j′}}, the automorphism induced by a permutation σ that σ(i) = i′ and σ(j) = j′, maps e to e′. It follows that S(Kk+1) is edge-transitive, and so L(S(Kk+1)) ∼= S++(2, k) is vertex-transitive. Hence, assume that n ≥ 3 and k ≥ 3. We observe that an extreme vertex of a copy ∆ of S(n−1, k) in Γ = S++(n, k) cannot be mapped to a non-extreme vertex of ∆ by any automorphism of Γ. To be more precise, let u = (1, . . . , 1, 1) and v = (1, . . . , 1, 2). It can be seen that u is a cut vertex for the induced subgraph by the vertices at distance at most 3 form u, while v is not a cut vertex for the induced subgraph by the vertices at distance at most 3 form v. It follows that u cannot be mapped to v by any automorphism of Γ, and thus Γ is not vertex-transitive. From Proposition 4.1, it follows that the graphs S++(n, k) for n ≥ 3 and k ≥ 3 cannot be Cayley graphs. The graphs S++(n, 1), S++(n, 2), and S++(1, k) are all Cayley graphs. It remains to characterize what S++(2, k) are Cayley graphs for k ≥ 3. This is our goal in the rest of this section. Definition 4.2. Let Γ be a graph and ∆ a subgraph of Γ. We say that Γ is strongly ∆- partitioned if: (i) The vertex set of Γ is partitioned by the vertex sets of copies ∆0, . . . ,∆k of ∆. (ii) Apart from ∆0, . . . ,∆k, the graph Γ contains no further copies of ∆. By the way S++(n, k) is defined, it is constructed based on k+1 copies of S(n−1, k). The following proposition gives a structural property of S++(n, k) that it is indeed strongly S(n − 1, k)-partitioned for n ≥ 2 and k ≥ 3. Note that this is not the case for k = 2 because S++(n, 2), that is a cycle with 3 · 2n−1 vertices, contains more than three copies of S(n− 1, 2), which is a path on 2n−1 vertices. Although we only need the case n = 2 of the proposition, we state it in its full generality because it could be of independent interest. Proposition 4.3. Let n ≥ 2 and k ≥ 3. The graph S++(n, k) is strongly S(n − 1, k)- partitioned. Proof. Let Γ := S++(n, k) and Γ0, . . . ,Γk be the k + 1 copies of S(n − 1, k) used to construct Γ by its definition. Clearly, V (Γ0), . . . , V (Γk) is a partition of V (Γ). We show that Γ contains no more copies of S(n − 1, k). Let ∆ be a subgraph of Γ isomorphic to S(n− 1, k). First, assume that n = 2. Let u ∈ V (∆) ∩ V (Γt) for some t, with 0 ≤ t ≤ k. Since u has at most one neighbor in V (Γ) \ V (Γt) and k ≥ 3, there exists another vertex v ∈ V (∆)∩V (Γt) adjacent to u. Now, ifw is any vertex of ∆ other than u and v, then since w is adjacent to two vertices u and v of Γt it must belong to V (Γt). Hence V (∆) ⊆ V (Γt) and, consequently, ∆ = Γt. M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 181 Now, let n ≥ 3. Note that S(n − 1, k) is connected and has no bridges since every edge of S(n − 1, k) lies on a cycle (which can be seen by induction on n). If ∆ ̸= Γi for i = 0, . . . , k, then ∆ shares its vertices with at least two Γs and Γt. By the definition, exactly one extreme vertex, say u, of Γs is adjacent to exactly one extreme vertex, say v, of Γt. Because of the connectivity, ∆ must contain the edge uv. Note that for any vertex w outside Γs and Γt, the distance between w and either u or v is greater than the diameter of S(n−1, k), and so w ̸∈ V (∆). It follows that ∆ is a subgraph of Γ′ := Γ[V (Γs)∪V (Γt)]. However, uv is a bridge for Γ′ and thus a bridge for ∆, a contradiction. The following lemma reveals the structure of strongly ∆-partitioned Cayley graphs. Lemma 4.4. Let Γ be a Cayley graph with a subgraph ∆ such that Γ is strongly ∆- partitioned. Then, the vertex sets of the copies of ∆ are all the right cosets of a subgroup of the underlying group of Γ. Proof. Let Γ be a Cayley graph on a group G, and X ⊆ G be such that 1 ∈ X and Γ[X], the subgraph of Γ induced by X , is isomorphic to ∆. Since for any x ∈ X , Γ[Xx−1] is isomorphic to ∆ and 1 ∈ Xx−1, from the hypothesis of the lemma, it follows that Xx−1 = X . Thus XX−1 = X and, hence, X is a subgroup of G. As for any g ∈ G, Γ[Xg] is isomorphic to ∆ and the setsXg cover all elements ofG, it follows that the vertex set of every induced subgraph of Γ isomorphic to ∆ is a right coset of X , as required. Definition 4.5. Let Γ be a strongly ∆-partitioned graph. We say that Γ has connection constant c if there are exactly c edges between any two copies of ∆ in Γ. We denote the set of all strongly ∆-partitioned graphs with connection constant c by SPc(∆). Remark 4.6. The family SP1(Kd) contains only one regular graph. However, this is not the case in any regular graph ∆. If Γ ∈ SP1(∆) is a regular graph with ∆ being a d- regular graph on k vertices, then Γ necessarily contains k + 1 copies of ∆ and thus Γ is (d+1)-regular with k(k+1) vertices. For instance, in the case in which ∆ is C4, the cycle on 4 vertices, Γ is a cubic graph on 20 vertices. By a computer search, we found all the regular graphs in SP1(C4). It turned out that there are seven non-isomorphic such graphs, among which only one is a Cayley graph. Here we recall some notions from group theory that will be used in what follows. Let G be a finite group and H be a nontrivial proper subgroup of G. The conjugate of H by an element g of G is defined as Hg = {hg : h ∈ G}, where hg := g−1hg denotes the conjugate of h by g. The group G is called a Frobenius group with Frobenius complement H if H ∩ Hg = {1} for all g ∈ G \ H . A celebrated theorem of Frobenius states that N := G \ ⋃ g∈G(H \ {1})g is a normal subgroup of G, called the Frobenius kernel of G, satisfying G = NH and N ∩H = {1}, that is, G = N ⋊H is a semidirect product of N by H (see [17, 8.5.5]). The other concepts we use in the following are standard and can be found in Robinson [17]. Theorem 4.7. Suppose that Γ and ∆ are two regular graphs and Γ ∈ SP1(∆). If Γ is a Cayley graph Cay(G,C), then |∆|+1 = pm is a prime power, G = N ⋊H is a Frobenius group with minimal normal Frobenius kernel N ∼= Zmp and Frobenius complement H , C = C ′ ∪ {c} with ∆ ∼= Cay(H,C ′) and c2 = 1, and either (i) c ∈ N and H = ⟨C ′⟩, or 182 Ars Math. Contemp. 20 (2021) 171–186 (ii) c = hn for some h ∈ H \ {1} and n ∈ N \ {1}, and H = ⟨C ′, h⟩. Conversely, if ∆ satisfies the above conditions, then Cay(G,C) ∈ SP1(∆). Proof. Let Γ = Cay(G,C), and ∆0,∆1, . . . ,∆k be the copies of ∆ in Γ. As there is exactly one edge between any two copies of ∆, it is observed that |∆| = k and Γ is (d+1)- regular if ∆ is d-regular. Let H := V (∆0) and assume, without loss of generality, that 1 ∈ H . By Lemma 4.4, H is a subgroup of G. Let C ′ be the neighborhood of 1 in Γ[H]. Since Γ is (d + 1)-regular, besides the elements of C ′, the vertex 1 has exactly one other neighbor, say c ∈ G \ H . So C = C ′ ∪ {c}. Since H is a subgroup of G, C ′−1 ⊆ H , which implies that C ′−1 = C ′. Thus, c = c−1 is an involution. Clearly, Hc ̸= H so that Γ[Hc] = ∆i for some 1 ≤ i ≤ k. On the other hand, 1 = |E(∆0,∆i)| = |{{h, ch} : h ∈ H ∩Hc}| = |H ∩Hc|, from which it follows that H ∩Hc = {1}. Now, a simple verification shows that Hch ∩ Hch′ = ∅ for all distinct elements h, h′ ∈ H . Since Γ[H] = ∆0 and Γ[Hch] (h ∈ H) are equal to ∆1, . . . ,∆k in some order, we must have G = H ∪ ⋃ h∈H Hch, where the unions are disjoint. As a result, every element g ∈ G \ H can be written as g = hch′ for some h, h′ ∈ H , from which it follows that H ∩Hg = (Hh ′−1 ∩ (Hh)c)h ′ = (H ∩Hc)h ′ = {1}h ′ = {1}. Hence, G is a Frobenius group with complement H . Let N be the Frobenius kernel of G. By [17, 10.5.1(i)], N is nilpotent. Let N0 be a nontrivial characteristic subgroup of N with minimum order. Then N0 is a normal subgroup of G (see [17, 1.5.6(iii)]). Note that N0 is an elementary Abelian p-group for N0 is nilpotent and the subgroup of N0 generated by central elements of a given prime order p dividing |Z(N0)| is a characteristic subgroup of N0 and hence of N (see [17, 1.5.6(ii)]). If N ̸= N0, then N0H is a Frobenius group for N0H is a subgroup of G and H ∩ Hg = 1 for all g ∈ N0H \ H . Moreover, as a proper subgroup of N , |N0| ≤ |N |/2 ≤ (k + 1)/2 and hence |N0| − 1 is not divisible by |H| = k contradicting [17, Exercises 8.5(6)]. Thus N = N0 so that k + 1 = |N | = pm is a prime power for some m ≥ 1. Note that N is a minimal normal subgroup of G for if N contains a nontrivial normal subgroup N0 of G properly, then N0H would be a Frobenius group which leads us to the same contradiction as above. If c ∈ N , then since G ⊆ N⟨C ′⟩ it follows that H = ⟨C ′⟩. Now assume that c /∈ N . Then cn ∈ H \ {1} for some n ∈ N \ {1}. As G ⊆ N⟨C ′, cn⟩ it follows that H = ⟨C ′, cn⟩, as required. The converse is straightforward. We are now in a position to conclude the main result of this section. Theorem 4.8. The graph S++(n, k) is a Cayley graph if and only if either (i) n = 1, (ii) k ≤ 2, or M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 183 (iii) n = 2 and k + 1 = pm is a prime power. Furthermore, in the case (iii), we have S++(n, k) ∼= Cay(G, (H \ {1}) ∪ {c}), for every Frobenius group G with complement H of order pm − 1, elementary Abelian minimal normal Frobenius kernel of order pm, and involution c ∈ G \H . Proof. By Proposition 4.1, S++(n, k) for n ≥ 3 and k ≥ 3 is not a Cayley graph. As mentioned above, S++(n, 1), S++(n, 2), and S++(1, k) are all Cayley graphs. So, we may assume that n = 2 and k ≥ 3. First, we show that S++(2, q − 1) are Cayley graphs for all prime powers q. Let Fq denote the finite field with q elements. Then G := F∗q ×Fq together with the multiplication (x, a) · (y, b) = (xy, xb+ a), forms a group known as one dimensional affine group. We show that S++(2, q − 1) ∼= Cay(G,C), where C = { (x, 0) : 1 ̸= x ∈ F∗q } ∪ {(−1,−1)}. To this end, let H := {(x, 0) : x ∈ F∗q} be a subgroup of G of order q − 1. Then H has q right cosets each of which induces a complete subgraph in Cay(G,C) for h′g(hg)−1 = h′h−1 ∈ C for all distinct elements hg and h′g of a right coset Hg of H . Since (x, 0)(1, ax−1) = (x, a) covers all elements of G when x and a ranges over F∗q and Fq , respectively, it follows that every right coset of H has a representative of the form (1, b) for some b ∈ Fq . Let Hg and Hg′ be distinct right cosets of H with g = (1, a) and g′ = (1, a′). Then an element hg of Hg is adjacent to an element h′g′ of Hg′ if and only if h′g′g−1h−1 = (h′g′)(hg)−1 ∈ C or equivalently (h′g′)(hg)−1 = (−1,−1) as g′g−1 /∈ H . A simple verification shows that this equation has a unique solution for (h, h′) so that there is a unique edge between any two right cosets of H . Indeed, h = (x, 0) and h′ = (x′, 0) satisfy the equation if and only if −x′ = x = (a′ − a)−1. Hence, from the definition, it follows that S++(2, q − 1) ∼= Cay(G,C). Now, assume that Γ := S++(2, k) ∼= Cay(G,C) be a presentation of S++(2, k) as a Cayley graph. By Proposition 4.3, Γ is strongly Γ0-partitioned for some complete subgraph Γ0 of Γ of order k. Let H := V (Γ0) and assume that 1 ∈ H . We know from Lemma 4.4 that H is a subgroup of G. By Theorem 4.7, k + 1 = pm is a prime power, G = N ⋊H is a Frobenius group with Frobenius kernel N and Frobenius complement H such that N ∼= Zmp is a minimal normal subgroup of G, C = C ′ ∪ {c}, C ′−1 = C ′ ⊆ H , c2 = 1, and either (a) c ∈ N and H = ⟨C ′⟩, or (b) c = hn for some h ∈ H \ {1} and n ∈ N \ {1}, and H = ⟨C ′, h⟩. Since Γ0 is a complete graph, we must have H \ {1} ⊆ C. Then, (a) and (b) together are equivalent to say that c ∈ G \H . The proof is now complete. As a generalization of Theorem 4.7, we pose the following problem. Problem 4.9. Let ∆ be a regular graph. Classify all Cayley graphs in SPc(∆) for c ≥ 2. 184 Ars Math. Contemp. 20 (2021) 171–186 5 New non-Cayley numbers In this final section, we give a partial answer to a famous rather old open problem in alge- braic graph theory. A positive integer n is called a Cayley number if all vertex-transitive graphs of order n are Cayley graphs. Marušič [11] in 1983 posed the problem of charac- terizing the set NC of all non-Cayley numbers. Since disjoint unions of copies of vertex- transitive (non-Cayley) graphs are again vertex-transitive (non-Cayley) graphs, it follows that every multiple of a non-Cayley number is again a non-Cayley number. Hence the problem of determining NC reduces to finding ‘minimal’ non-Cayley numbers. It is well- known that all primes are Cayley numbers. Following a series of papers by various authors, McKay and Praeger [13] and Iranmanesh and Praeger [7] provided necessary and sufficient conditions under which the product of two and three distinct primes is a Cayley number, respectively. In the same paper, McKay and Praeger established the following remarkable result determining all non-square-free Cayley numbers. Theorem 5.1 (McKay and Praeger [13]). Let n be a positive integer that is divisible by the square of a prime p. Then n ∈ NC unless n = p2, n = p3, or n = 12. It follows that, for determining NC, it is enough to consider only square-free positive integers. While the problem is yet open for the products of at least four distinct primes, there are partial results worth to mention here. Theorem 5.2 (Dobson and Spiga [3]). There exists an infinite set of primes every finite product of its distinct elements is a Cayley number. As a consequence of Theorem 4.8, the graph S++(2, k) that has k(k + 1) vertices is not a Cayley graph if k + 1 is not a prime power. Therefore, we obtain a new infinite class of square-free non-Cayley numbers as follows. Theorem 5.3. Let k be any positive integer such that k(k+ 1) is square-free, and k+ 1 is not a prime. Then, k(k + 1) ∈ NC. As mentioned in Dobson and Spiga [3], it is straightforward by making use of the group-theoretic and the number-theoretic results already available in the literature to prove that Cayley numbers have density zero in the set of natural numbers, and hence the density of non-Cayley numbers is 1. In the light of this fact, one might wonder about the dis- tribution of the numbers k satisfying the conditions of Theorem 5.3 in the set of positive integers. The following theorem shows that for large enough N , more than one third of positive integers less than or equal to N satisfies the conditions of Theorem 5.3. Theorem 5.4. The density of the set {k : k(k + 1) is square-free, and k + 1 is not a prime} is about 0.3226. Proof. Let f ∈ Z[t] be a primitive polynomial (that is, the greatest common divisor of its coefficients is 1) without multiple roots such that its image on N has k-free greatest common divisor. Recall that a number that is not divisible by any proper k-th power is called k-free. Let Skf (x) denote the number of all positive integers n ≤ x such that f(n) is k-free, and consider δf,k := ∏ p prime ( 1− ϱ(p k) pk ) , M. Farrokhi D. G. et al.: Some algebraic properties of Sierpiński-type graphs 185 where ϱ(d) denotes the number of roots of f in Zd. Ricci [16] (see also Pappalardi [14]) proved that Skf (x) ∼ δf,kx provided that deg f ≤ k. Clearly, the function f(t) := t(t + 1) satisfies the above re- quirements of Ricci’s theorem for k = 2. Also, it is obvious that ϱ(p2) = 2 for all primes p. Thus, by Ricci’s theorem, the density of all positive integers k, for which k(k + 1) is square-free, in the set of all positive integers, is equal to δf,2 = ∏ p prime ( 1− 2 p2 ) = 2CFeller-Tornier − 1 ≈ 0.3226340989, where CFeller-Tornier is the Feller-Tornier constant (see Finch [4, §2.4.1]). Since primes have zero density in the set of all positive integers, the result follows. To date, all the numbers whose membership in NC is known are determined based on the results of [7, 12, 13]. Using a computer search, we see that the list of the numbers whose membership in NC is not yet determined begins with 9982, 12958, 18998, 19646, 20398, 21574, 24662, 25438, 25606, . . . . A simple computation reveals that among the numbers less than or equal to 108, there are 2763 square-free integers of the form k(k + 1), with k + 1 not a prime of which the following eight integers are new non-Cayley numbers: 1386506, 2668322, 15503906, 23985506, 38359442, 74261306, 89898842, 95912642. ORCID iDs Mohammad Farrokhi Derakhshandeh Ghouchan https://orcid.org/0000-0002-5850-969X Ebrahim Ghorbani https://orcid.org/0000-0001-7195-8601 Hamid Reza Maimani https://orcid.org/0000-0001-9020-0871 Farhad Rahimi Mahid https://orcid.org/0000-0001-8996-2078 References [1] L. Barrière, F. Comellas and C. Dalfó, Fractality and the small-world effect in Sierpinski graphs, J. Phys. A 39 (2006), 11739–11753, doi:10.1088/0305-4470/39/38/003. [2] D. M. Cvetković, M. Doob and H. Sachs, Spectra of Graphs: Theory and Applications, Johann Ambrosius Barth, Heidelberg, 3rd edition, 1995. [3] T. Dobson and P. Spiga, Cayley numbers with arbitrarily many distinct prime factors, J. Comb. Theory Ser. B 122 (2017), 301–310, doi:10.1016/j.jctb.2016.06.005. [4] S. R. Finch, Mathematical Constants, volume 94 of Encyclopedia of Mathematics and its Ap- plications, Cambridge University Press, Cambridge, 2003. [5] R. Grigorchuk and Z. Šunić, Schreier spectrum of the Hanoi Towers group on three pegs, in: P. Exner, J. P. Keating, P. Kuchment, T. Sunada and A. Teplyaev (eds.), Analysis on Graphs and Its Applications, American Mathematical Society, Providence, RI, volume 77 of Proceedings of Symposia in Pure Mathematics, pp. 183–198, 2008, doi:10.1090/pspum/077/2459869. [6] 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. 186 Ars Math. Contemp. 20 (2021) 171–186 [7] M. A. Iranmanesh and C. E. Praeger, On non-Cayley vertex-transitive graphs of order a product of three primes, J. Comb. Theory Ser. B 81 (2001), 1–19, doi:10.1006/jctb.2000.1985. [8] S. Klavžar and U. Milutinović, Graphs S(n, k) and a variant of the Tower of Hanoi problem, Czechoslovak Math. J. 47 (1997), 95–104, doi:10.1023/a:1022444205860. [9] S. Klavžar and B. Mohar, Crossing numbers of Sierpiński-like graphs, J. Graph Theory 50 (2005), 186–198, doi:10.1002/jgt.20107. [10] J. R. Magnus and H. Neudecker, The commutation matrix: some properties and applications, Ann. Statist. 7 (1979), 381–394. [11] D. Marušič, Cayley properties of vertex symmetric graphs, Ars Combin. 16 (1983), 297–302. [12] B. D. McKay and C. E. Praeger, Vertex-transitive graphs which are not Cayley graphs, I, J. Austral. Math. Soc. Ser. A 56 (1994), 53–63, doi:10.1017/s144678870003473x. [13] B. D. McKay and C. E. Praeger, Vertex-transitive graphs that are not Cayley graphs, II, J. Graph Theory 22 (1996), 321–334, doi:10.1002/(sici)1097-0118(199608)22:4⟨321::aid-jgt6⟩ 3.3.co;2-9. [14] F. Pappalardi, A survey on k-freeness, in: S. D. Adhikari, R. Balasubramanian and K. Srinivas (eds.), Number Theory, Ramanujan Mathematical Society, Mysore, volume 1 of Ramanujan Mathematical Society Lecture Notes Series, pp. 71–88, 2005, proceedings of the International Conference on Analytic Number Theory with Special Emphasis on L-functions held in Chen- nai, January 2002. [15] C. E. Praeger, C. H. Li and A. C. Niemeyer, Finite transitive permutation groups and finite vertex-transitive graphs, in: G. Hahn and G. Sabidussi (eds.), Graph Symmetry: Algebraic Methods and Applications, Kluwer Academic Publishers Group, Dordrecht, volume 497 of NATO Advanced Science Institutes Series C: Mathematical and Physical Sciences, pp. 277– 318, 1997, doi:10.1007/978-94-015-8937-6 7, proceedings of the NATO Advanced Study In- stitute and the Séminaire de Mathématiques Supérieures held in Montreal, PQ, July 1 – 12, 1996. [16] G. Ricci, Ricerche aritmetiche sui polinomi, Rend. Circ. Mat. Palermo 57 (1933), 433–475. [17] D. J. S. Robinson, A Course in the Theory of Groups, volume 80 of Graduate Texts in Mathe- matics, Springer-Verlag, New York, 2nd edition, 1996, doi:10.1007/978-1-4419-8594-1. [18] M. A. Snyder, Chebyshev Methods in Numerical Approximation, Prentice-Hall, Englewood Cliffs, NJ, 1966. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 187–194 https://doi.org/10.26493/1855-3974.2435.3db (Also available at http://amc-journal.eu) The chromatic index of strongly regular graphs Sebastian M. Cioabă * Department of Mathematical Sciences, University of Delaware, United States Krystal Guo † Korteweg-De Vries Institute, University of Amsterdam, The Netherlands Willem H. Haemers Department of Econometrics and Operations Research, Tilburg University, The Netherlands Received 16 September 2020, accepted 20 January 2021, published online 27 October 2021 Abstract We determine (partly by computer search) the chromatic index (edge-chromatic num- ber) of many strongly regular graphs (SRGs), including the SRGs of degree k ≤ 18 and their complements, the Latin square graphs and their complements, and the triangular graphs and their complements. Moreover, using a recent result of Ferber and Jain, we prove that an SRG of even order n, which is not the block graph of a Steiner 2-design or its complement, has chromatic index k, when n is big enough. Except for the Petersen graph, all investigated connected SRGs of even order have chromatic index equal to k, i.e., they are class 1, and we conjecture that this is the case for all connected SRGs of even order. Keywords: Strongly regular graph, chromatic index, edge coloring, 1-factorization. Math. Subj. Class. (2020): 05C15, 05E30 1 Introduction An edge-coloring of a graph G is a coloring of its edges such that intersecting edges have different colors. Thus a set of edges with the same colors (called a color class) is a match- ing. The edge-chromatic number χ′(G) (also known as the chromatic index) of G is the *Research supported by NSF grants DMS-1600768 and CIF-1815922. †This research was done while K. Guo was a postdoctoral fellow at Université Libre de Bruxelles, supported by ERC Consolidator Grant 615640-ForEFront. E-mail addresses: cioaba@udel.edu (Sebastian M. Cioabă), k.guo@uva.nl (Krystal Guo), haemers@uvt.nl (Willem H. Haemers) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 188 Ars Math. Contemp. 20 (2021) 187–194 minimum number of colors in an edge-coloring. By Vizing’s famous theorem [24], the chromatic index of a graph G of maximum degree ∆ is ∆ or ∆ + 1. A graph with maxi- mum degree ∆ is called class 1 if χ′(G) = ∆ and is called class 2 if χ′(G) = ∆ + 1. It is also known that determining whether a graph G is class 1 is an NP-complete problem [18]. If G is regular of degree k, then G is class 1 if and only if G has an edge coloring such that each color class is a perfect matching. A perfect matching is also called a 1-factor, and a partition of the edge set into perfect matchings is called a 1-factorization. So being regular and class 1 is the same as having a 1-factorization (being 1-factorable), and requires that the graph has even order. A graph G is called a strongly regular graph (SRG) with parameters (n, k, λ, µ) if it has n vertices, is k-regular (0 < k < n− 1), any two adjacent vertices of G have exactly λ common neighbors and any two distinct non-adjacent vertices of G have exactly µ common neighbors. The complement of a strongly regular graph with parameters (n, k, λ, µ) is again strongly regular, and has parameters (n, n − k − 1, n − 2k + µ − 2, n − 2k + λ). An SRG G is called imprimitive if G or its complement is disconnected, and primitive otherwise. A disconnected SRG is ℓKm (ℓ,m ≥ 2), the disjoint union of ℓ cliques of order m (indeed, µ = 0 implies that no two vertices have distance two, so every connected component is a clique). It is well-known that Km (m ≥ 2), and hence also ℓKm, is class 1 if and only if m is even. The complement of ℓKm is a regular complete multipartite graph which is known to be class 1 if and only if the order is even [17]. A vertex coloring of G is a coloring of the vertices of G such that adjacent vertices have different colors. The chromatic number χ(G) of G is the minimum number of colors in a vertex coloring. For the chromatic number there exist bounds in terms of the eigenvalues of the adjacency matrix, which turn out to be especially useful for strongly regular graphs (see for example [6]). These bounds imply that there exist only finitely many primitive SRGs with a given chromatic number, and made it possible to determine all SRGs with chromatic number at most four (see [15]). Motivated by these results, Alex Rosá asked the third author whether eigenvalue techniques can give information on the chromatic index of an SRG. There exist useful spectral conditions for the existence of a perfect matching (see [5, 9]), and Brouwer and Haemers [5] have shown that every regular graph of even order, degree k and second largest eigenvalue ϑ2 contains at least ⌊(k − ϑ2 + 1)/2⌋ edge disjoint perfect matchings. From this it follows that every connected SRG of even order has a perfect matching. Moreover, Cioabă and Li [10] proved that any matching of order k/4 of a primitive SRG of valency k and even order, is contained in a perfect matching. These authors conjectured that k/4 can be replaced by ⌈k/2⌉ − 1 which would be best possible. Unfortunately, we found no useful eigenvalue tools for determining the chromatic index. However, the following recent result of Ferber and Jain [13] gives an asymptotic condition for being class 1 in terms of the eigenvalues. Theorem 1.1. There exist universal constants n0 and k0, such that the following holds. If G is a connected k-regular graph of even order n with eigenvalues k = ϑ1 > ϑ2 ≥ · · · ≥ ϑn, and n > n0, k > k0 and max{ϑ2,−ϑn} < k0.9, then G is class 1. If the maximum distance in G is 2 (as is the case for a connected SRG), then n ≤ 1 + k + k(k − 1) = k2 + 1. This implies that for an SRG we do not need to require that k > k0 when we take n0 ≥ k20 + 1. Theorem 1.1 enables us to show that, except for one family of SRGs, all connected SRGs of even order n are class 1, provided n is large enough. In addition, we present a number of sufficient conditions for an SRG to be class 1. S. M. Cioabă, K. Guo and W. H. Haemers: The chromatic index of strongly regular graphs 189 By computer, using SageMath [22], we verified that all primitive SRGs of even order and degree k ≤ 18 and their complements are class 1, except for the Petersen graph, which has parameters (10, 3, 0, 1) and edge-chromatic number 4 (see [20, 25] for example). We also determine the chromatic index of several other primitive SRGs of even order, and all are class 1. Therefore we believe: Conjecture 1.2. Except for the Petersen graph, every connected SRG of even order is class 1. 2 Sufficient conditions for being class 1 A well known conjecture (first stated by Chetwynd and Hilton [8], but attributed to Dirac) states that every k-regular graph of even order n with k ≥ n/2 is 1-factorable. Cariolaro and Hilton [7] proved that the conclusion holds when k ≥ 0.823n, and Csaba, Kühn, Lo, Osthus, and Treglown [11], proved the following result. Theorem 2.1. There exists a universal constant n0, such that if n is even, n > n0 and if k ≥ 2⌈n/4⌉ − 1, then every k-regular graph of order n has chromatic index k. König [19] proved that every regular bipartite graph of positive degree has a 1-factor- ization. We need the following generalization of König’s result. Lemma 2.2. Let G = (V,E) be a connected regular graph of even order n, and let {V1, V2} be a partition of V such that |V1| = |V2| = n/2. (i) If the graphs induced by V1 and V2 are 1-factorable, then so is G. (ii) If V1 (and hence V2) is a clique or a coclique, then G is class 1. Proof. Partition the edge set E into two classes E1 and E2, where E1 contains all edges with both endpoints in the same vertex set V1 or V2, and the edges of E2 have one endpoint in V1 and the other endpoint in V2. (i): If the graphs induced by V1 and V2 are 1-factorable, then both have the same degree, and therefore also (V,E1) is 1-factorable. By König’s theorem (V,E2) is 1-factorable, therefore G is class 1. (ii): If V1 is a coclique, then so is V2 and we have the theorem of König. If V1 is a clique, then so is V2. If n/2 is even, then the result is proved in (i). If n/2 is odd, then we choose a 1-factor F in (V,E2) (here we use that G is connected), and define E′2 = E2 \ F and E′1 = E1 ∪ F . Then (V,E′2) is 1-factorable (or has no edges), and (V,E′1) consists of two cliques of order n/2 and the 1-factor F . Thus F gives a bijection ϕ (say) between V1 and V2. By Vizing’s theorem the edges of both cliques can be colored with n/2 colors. We do this coloring such that ϕ preserves colors, which means that {v, w} and {ϕ(v), ϕ(w)} get the same color. Then for each edge {v, ϕ(v)} of F , the two sets of colored edges that intersect at v and ϕ(v) use the same n/2 − 1 colors. So we can color {v, ϕ(v)} with the remaining color. There exist several SRGs that have the partition of case (i). The Gewirtz graph is the unique SRG with parameters (56, 10, 0, 2), and admits a partition into two Coxeter graphs (see [4]). The Coxeter graph is known to be 1-factorable (see [3]), therefore the Gewirtz graph is class 1. The same holds for the point graph of the generalized quadrangle GQ(3, 9) (the unique SRG(112, 30, 2, 10)), which admits a partition into two Gewirtz graphs, and for 190 Ars Math. Contemp. 20 (2021) 187–194 the Higman-Sims graph (the unique SRG(100, 22, 0, 6)), which can be partitioned into two copies of the Hoffman-Singleton graph (the unique strongly regular graph with parameters (50, 7, 0, 1)), which has chromatic index 7 (see Section 4). Suppose ϑ1 ≥ ϑ2 ≥ · · · ≥ ϑn are the eigenvalues of a graph G of order n. Hoffman (see [6, Theorem 3.6.2] for example) proved that the chromatic number of G is at least 1 − ϑ1/ϑn. A vertex coloring that meets this bound is called a Hoffman coloring. For k-regular graphs, the color classes of a Hoffman coloring are cocliques of which the size meets Hoffman’s coclique bound nϑn/(ϑn−k). This implies (see [6] for example) that all the color classes have equal size, and any vertex v of G has exactly −ϑn neighbors in each color class different from the color class of v. Theorem 2.3. Suppose G = (V,E) is a k-regular graph with an even chromatic number 2t (say) that meets Hoffman’s bound. Then both G and its complement G are class 1, or G is a disjoint union of cliques of odd order. Proof. Let S1, . . . , S2t be the color classes in a Hoffman coloring of G. Since G is regular, this implies that each Si is a coclique attaining equality in the Hoffman ratio bound, which means that each vertex outside Si has exactly −ϑn neighbors in Si. Hence, each subgraph induced by two distinct cocliques Si and Sj is a bipartite regular graph of valency −ϑn. A 1-factorization of K2t corresponds to a partition E1, . . . , E2t−1 of E, such that each (V,Ei) consists of t disjoint regular bipartite graphs of degree −ϑn = k/(2t − 1). By König’s theorem it follows that each (V,Ei) is 1-factorable, and therefore G is class 1. For the complement G = (V, F ) of G, a similar approach works. We can partition the edge set F into the subsets F0, F1, . . . , F2t−1, such that for i = 1, . . . , 2t − 1 the graph (V, Fi) is the disjoint union of t regular bipartite graphs of the same positive degree (if the degree is 0, then G is a disjoint union of cliques). But now there is an additional graph (V, F0) consisting of 2t disjoint cliques. We combine F0 and F1. Then (V, F0 ∪ F1) is the disjoint union of t complements of regular incomplete bipartite graphs with the same positive degree, and therefore has a 1-factorization by Lemma 2.2. Since (V, Fi) has a 1-factorization for i = 2, . . . , 2t− 1, it follows that G is 1-factorable. For an SRG the color partition of a Hoffman coloring corresponds to a so-called spread in the complement (see [16]). As a consequence of this result, it follows that any primitive strongly regular graph with a spread with an even number of cliques, or a Hoffman coloring with an even number of colors is class 1. Among such SRGs are the Latin square graphs. Consider a set of t (t ≥ 0) mutually orthogonal Latin squares of order m (m ≥ 2). The vertices of the Latin square graph are the m2 entries of the Latin squares, and two distinct entries are adjacent if they lie in the same row, the same column, or have the same symbol in one of the squares. If t = m− 1 we obtain the complete graph Km2 , and if t = m− 2 we have a complete multipartite graph. Otherwise the Latin square graph is a primitive SRG with parameters (m2, (t+ 2)(m− 1),m− 2 + t(t+ 1), (t+ 1)(t+ 2)). If t = 0 we only have the rows and columns, then the Latin square graph is better known as the Lattice graph L(m). If m ̸= 4, the Lattice graph is determined by the parameters. The m rows of a Latin square give a partition of the vertex set of the Latin square graph into cliques, which is a spread. Thus we have: Corollary 2.4. If G is a Latin square graph of even order, then both G and its complement are 1-factorable. S. M. Cioabă, K. Guo and W. H. Haemers: The chromatic index of strongly regular graphs 191 3 Asymptotic results A Steiner 2-design (or 2-(m, ℓ, 1) design) consists of a point set P of cardinality m, to- gether with a collection of subsets of P of size ℓ (ℓ ≥ 2), called blocks, such that every pair of points from P is contained in exactly one of the blocks. The block graph of a Steiner 2-design is defined as follows. The blocks are the vertices, and two vertices are adjacent if the blocks intersect in one point. If m = ℓ2 − ℓ + 1, the Steiner 2-design is a projective plane, and the block graph is Km. Otherwise the block graph is an SRG with parameters (m(m− 1)/ℓ(ℓ− 1), ℓ(m− ℓ)/(ℓ− 1), (ℓ− 1)2 + (m− 2ℓ+ 1)/(ℓ− 1), ℓ2). Theorem 3.1. There exists an integer n0, such that every primitive strongly regular graph of even order n > n0, which is not the block graph of a Steiner 2-design or its complement, is class 1. Proof. Suppose G is a primitive (n, k, λ, µ)-SRG of even order n. Then it is well-known (see for example [6, Chapter 9]) that G has exactly three distinct eigenvalues k = ϑ1, ϑ2 and ϑn. Moreover, the eigenvalues are nonzero integers and satisfy k + ϑ2ϑn = µ. Assume that G nor its complement G is the block graph of a Steiner 2-design or a Latin square graph. Using a result of Neumaier [21] (known as the claw bound), we get that ϑ2 ≤ ϑn(ϑn + 1)(µ+ 1)/2− 1. Another result of Neumaier [21] (the µ-bound) gives µ ≤ ϑ3n(2ϑn + 3). Combining these inequalities, after some straightforward calculations, we obtain that ϑ2 < (−ϑn)6. Since k + ϑ2ϑn = µ > 0, we deduce that k6 > (−ϑ2ϑn)6 > ϑ62ϑ2 = ϑ72, so ϑ2 < k6/7. Next we apply the same result to G, and obtain −1 − ϑn < (1 + ϑ2)6, which yields −ϑn ≤ (2ϑ2)6 (since ϑ2 is a positive integer). Hence k6 > (−ϑ2ϑn)6 > 2−6(−ϑn)(−ϑn)6 = 2−6(−ϑn)7, so − ϑn < (2k)6/7 < k0.9, when k is large enough. Thus we get max{ϑ2,−ϑn} ≤ k0.9. Now we apply the result of Ferber and Jain and conclude that G is class 1 when n is large enough. If G is a Latin square graph of even order then by Corollary 2.4 both G and its comple- ment G are class 1. In many cases the complement of the block graph of a Steiner 2-design has k > n/2, so it will have a 1-factorization by Theorem 2.1, provided n is even and large enough. The following result follows straightforwardly from the mentioned result of Cariolaro and Hilton [7]. Proposition 3.2. If G is the complement of the block graph of a 2-(m, ℓ, 1) design with 6ℓ2 ≤ m, then G is class 1, provided G has even order. For every m ≥ 2 there is a unique 2-(m, 2, 1) design, and its block graph is the triangu- lar graph T (m). It is isomorphic to the line graph of the complete graph Km, and if m ≥ 4 T (m) is an SRG with parameters (m(m− 1)/2, 2(m− 2),m− 2, 4). The triangular graph is uniquely determined by its parameters if m ̸= 8. Alspach [1] has proved that T (m) has 192 Ars Math. Contemp. 20 (2021) 187–194 a 1-factorization if the order is even, which is the case if m ≡ 0, or 1 (mod 4). Proposi- tion 3.2 implies that the complement of T (m) is class 1 if m ≥ 24 and the order is even. The complement of T (5) is the Petersen graph, which is class 2. For 5 < m < 24 and m ≡ 0, or 1 (mod 4) we found a 1-factorization in the complement of T (m) by computer (see next section for more about the computer search). Thus we can conclude: Theorem 3.3. For m ≡ 0, or 1 (mod 4) the triangular graph T (m) is class 1, and if m ̸= 5 so is its complement. If the block size equals 3, the design is better known as a Steiner triple system. The chromatic index of the block graph of a Steiner triple system is investigated in [12]. The paper contains several sufficient conditions for such a graph to be class 1, and the authors conjecture that all these graphs are class 1 when the order is even. One of the results from [12] (Theorem 2.2) can be generalized to arbitrary Steiner 2-designs. A set of m/ℓ disjoint blocks of a 2-(m, ℓ, 1) design is called a parallel class, and a partition of the block graph into parallel classes is a parallelism. A parallelism of a Steiner 2-design gives a Hoffman coloring in the block graph, so we have: Proposition 3.4. If a Steiner 2-design has a parallelism with an even number of parallel classes, then the block graph and its complement are class 1. 4 SRGs of degree at most 18 According to the list of Brouwer [2] all primitive SRGs of even order and degree at most 18 are known (one only has to check the parameter sets up to n = 182 + 1 = 325). The parameters together with the number of nonisomorphic SRGs with k < n/2 are given in Table 1 (the ones with k ≥ n/2 are the complements of a to e). The graph with parameter Table 1: Primitive SRGs with n even and k ≤ 18, k < n/2. a (10, 3, 0, 1) 1 f (36, 10, 4, 2) 1 k (50, 7, 0, 1) 1 b (16, 5, 0, 2) 1 g (36, 14, 4, 6) 180 l (56, 10, 0, 2) 1 c (16, 6, 2, 2) 2 h (36, 14, 7, 4) 1 m (64, 14, 6, 2) 1 d (26, 10, 3, 4) 10 i (36, 15, 6, 6) 32548 n (64, 18, 2, 6) 167 e (28, 12, 6, 4) 4 j (40, 12, 2, 4) 28 o (100, 18, 8, 2) 1 set a is the Petersen graph, which is class 2. The complement of the Petersen graph is the triangular graph T (5) which is class 1 by Alspach’s result [1]. Also Case h and one of the graphs of Case e is a triangular graph and therefore class 1. For the parameter sets f , m and o there is a unique SRG, the so called Lattice graph. This SRG belongs to the Latin square graphs, and by Corollary 2.4 the graph is class 1, and so is its complement. Case l is the Gewirtz graph, which is class 1 by Lemma 2.2, as we saw in Section 2. All other graphs are tested by computer (we actually tested all graphs in Table 1 and their complements). Using SageMath [22], we wrote a computer program that searches for an edge coloring in a k-regular graph with k colors. In each step we look (randomly) for a perfect matching, remove all its edges and continue until the remaining graph has no perfect matching. If there are still edges left we start again. We run this algorithm repeatedly until an edge coloring is found. The code for this project is made freely available in a public GitHub repository which can be found at [14]. By use of this approach we found a 1-factorization S. M. Cioabă, K. Guo and W. H. Haemers: The chromatic index of strongly regular graphs 193 in all graphs of Table 1, and in their complements, except for the Petersen graph. Thus we found: Theorem 4.1. With the single exception of the Petersen graph, a primitive SRG of even order and degree at most 18 is class 1 and so is its complement. For the description of the graphs we used the website of Spence [23]. This website also contains several SRGs with parameters (50, 21, 8, 9). We also ran the search for these graphs. All are class 1. It is surprising that in all cases our straightforward heuristic finds a 1-factorization. The heuristic is fast. It took about one hour to find a 1-factorization in each of the 32548 SRGs with parameter set i. ORCID iDs Sebastian M. Cioabă https://orcid.org/0000-0001-9983-0212 Krystal Guo https://orcid.org/0000-0001-8776-537X Willem H. Haemers https://orcid.org/0000-0001-7308-8355 References [1] B. Alspach, A 1-factorization of the line graphs of complete graphs, J. Graph Theory 6 (1982), 441–445, doi:10.1002/jgt.3190060408. [2] A. E. Brouwer, Parameters of strongly regular graphs, http://www.win.tue.nl/˜aeb/ graphs/srg/srgtab.html. [3] A. E. Brouwer, A slowly growing collection of graph descriptions, http://www.win.tue. nl/˜aeb/graphs/index.html. [4] A. E. Brouwer and W. H. Haemers, The gewirtz graph – an exercise in the theory of graph spectra, European J. Combin. 14 (1993), 397–407, doi:10.1006/eujc.1993.1044. [5] A. E. Brouwer and W. H. Haemers, Eigenvalues and perfect matchings, Linear Algebra Appl. 395 (2005), 155–162, doi:10.1016/j.laa.2004.08.014. [6] A. E. Brouwer and W. H. Haemers, Spectra of Graphs, Springer, New York, NY, 2012, doi: 10.1007/978-1-4614-1939-6. [7] D. Cariolaro and A. J. W. Hilton, An application of Tutte’s theorem to 1-factorization of regular graphs of high degree, Discrete Math. 309 (2009), 4736–4745, doi:10.1016/j.disc.2008.05.046. [8] A. G. Chetwynd and A. J. W. Hilton, Regular graphs of high degree are 1-factorizable, Proc. London Math. Soc. 50 (1985), 193–206, doi:10.1112/plms/s3-50.2.193. [9] S. M. Cioabă, D. A. Gregory and W. H. Haemers, Matchings in regular graphs from eigenval- ues, J. Comb. Theory Ser. B 99 (2009), 287–297, doi:10.1016/j.jctb.2008.06.008. [10] S. M. Cioabă and W. Li, The extendability of matchings in strongly regular graphs, Electron. J. Combin. 21 (2014), #P2.34 (23 pages), doi:10.37236/4142. [11] B. Csaba, D. Kühn, A. Lo, D. Osthus and A. Treglown, Proof of the 1-factorization and Hamil- ton Decomposition Conjectures, volume 244 of Memoirs of the American Mathematical Soci- ety, American Mathematical Society, Providence, RI, 2016, doi:10.1090/memo/1154. [12] I. Darijani, D. A. Pike and J. Poulin, The chromatic index of block intersection graphs of Kirkman triple systems and cyclic Steiner triple systems, Australas. J. Combin. 69 (2017), 145–158, https://ajc.maths.uq.edu.au/pdf/69/ajc_v69_p145.pdf. 194 Ars Math. Contemp. 20 (2021) 187–194 [13] A. Ferber and V. Jain, 1-factorizations of pseudorandom graphs, Random Structures Algorithms 57 (2020), 259–278, doi:10.1002/rsa.20927. [14] K. Guo, kguo-sagecode/srg-edge-coloring: Determining if a k-regular graph is k-edge- colorable, Version v1.0.0, Zenodo, 2020, doi:10.5281/zenodo.4031224. [15] W. H. Haemers, Eigenvalue techniques in design and graph theory, Ph.D. thesis, Eindhoven University of Technology, The Netherlands, 1979, doi:10.6100/ir41103. [16] W. H. Haemers and V. Tonchev, Spreads in strongly regular graphs, Des. Codes Cryptogr. 8 (1996), 145–157, doi:10.1007/bf00130574. [17] D. G. Hoffman and C. A. Rodger, The chromatic index of complete multipartite graphs, J. Graph Theory 16 (1992), 159–163, doi:10.1002/jgt.3190160207. [18] I. Holyer, The NP-completeness of edge-colouring, SIAM J. Comput. 10 (1981), 718–720, doi: 10.1137/0210055. [19] D. König, Über Graphen und ihre Anwendung auf Determinantentheorie und Mengenlehre, Math. Ann. 77 (1916), 453–465, doi:10.1007/bf01456961. [20] R. Naserasr and R. Škrekovski, The Petersen graph is not 3-edge-colorable – a new proof, Discrete Math. 268 (2003), 325–326, doi:10.1016/s0012-365x(03)00138-9. [21] A. Neumaier, Strongly regular graphs with smallest eigenvalue −m, Arch. Math. 33 (1979), 392–400, doi:10.1007/bf01222774. [22] The Sage Developers, SageMath, the Sage Mathematics Software System (Version 8.3), http://www.sagemath.org. [23] E. Spence, Strongly regular graphs on at most 64 vertices, http:/www.maths.gla.ac. uk/˜es/srgraphs.php. [24] V. G. Vizing, Critical graphs with given chromatic class, Diskret. Analiz 5 (1965), 9–17. [25] L. Volkmann, The Petersen graph is not 1-factorable: postscript to ‘The Petersen graph is not 3-edge-colorable – a new proof’ [Discrete Math. 268 (2003), 325–326], Discrete Math. 287 (2004), 193–194, doi:10.1016/j.disc.2004.07.008. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 195–197 https://doi.org/10.26493/1855-3974.2334.331 (Also available at http://amc-journal.eu) A simple and elementary proof of Whitney’s unique embedding theorem Gunnar Brinkmann * Applied Mathematics, Computer Science and Statistics, Ghent University, Belgium Received 13 May 2020, accepted 12 February 2021, published online 27 October 2021 Abstract In this note we give a short and elementary proof of a more general version of Whit- ney’s theorem that 3-connected planar graphs have a unique embedding in the plane. A consequence of the theorem is also that cubic plane graphs cannot be embedded in a higher genus with a simple dual. The aim of this paper is to promote a simple and elementary proof, which is especially well suited for lectures presenting Whitney’s theorem. Keywords: Polyhedra, graph, embedding. Math. Subj. Class. (2020): 05C10, 57M60, 57M15 1 Introduction Whitney’s famous unique embedding theorem has been formulated in various equivalent forms. One form is that the facial cycles of 3-connected graphs embedded in the plane are well determined, so that for any two embeddings there is a graph isomorphism between the duals. Another is (implied by the Jordan-Schönflies Theorem) that any two topological embeddings of a graph on the sphere can be mapped onto each other by a homeomorphism of the sphere that maps the two images of a vertex onto each other. We will formulate the theorem and describe the proof in the language of combinatorial embeddings in oriented closed surfaces. For the translation to the language of topological 2-cell embeddings, methods from standard books like [1] or [3] can be used. We interpret each edge {u, v} of an undirected embedded graph G as two directed edges: e = (u, v) and its inverse e−1 = (v, u). An embedded graph in an oriented closed surface is a graph where for every vertex u there is a cyclic order of all edges (u, .) (usually called a rotation). The cyclic ordering defines the orientation around the vertex. We write *I would like to thank Bojan Mohar for pointing me to the earlier uses of the crossing Jordan curves argument! E-mail address: gunnar.brinkmann@ugent.be (Gunnar Brinkmann) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 196 Ars Math. Contemp. 20 (2021) 195–197 nx(e) for the next edge in the order around the starting point of a directed edge e. The inverse graph or mirror image is the graph G−1 with all cyclic orders reversed. A face in an embedded graph G is a directed cyclic walk e0, . . . , en−1, so that for 0 ≤ i < n we have that nx(e−1i ) = e(i+1) (mod n). We say that the set {e,nx(e)} forms an angle of G and G−1 if one of them has a face containing e−1,nx(e) as a subsequence. In this case the other has a face containing nx(e)−1, e. If a face is a simple cyclic walk, we call the corresponding undirected cycle also a (simple) facial cycle. We consider an embedded graph G and its mirror image G−1 as equivalent, as the faces have the same sequences of underlying undirected edges – only in reversed order. The genus of an embedded graph can be computed by the Euler formula using the number v of vertices, e of (undirected) edges, and f of faces as γ(G) = 2−(v−e+f)2 . We refer to a (not necessarily embedded) graph that can be embedded with genus 0 as planar and to an embedded graph with genus 0 as plane. With this notation and concept of equivalence Whitney’s famous theorem [5] can be shortly stated as: A 3-connected planar graph has an – up to equivalence – unique embedding in the plane. We will prove a stronger theorem using the concept of polyhedral embedding that re- quires some important properties of polyhedra – that is plane 3-connected graphs – but allows higher genera. It is an easy consequence of the Jordan Curve Theorem that polyhe- dra are polyhedral embeddings. Definition 1.1. A polyhedral embedding of a graph G = (V,E) in an oriented closed surface is an embedding so that each facial walk is a simple cycle and the intersection of any two faces is either empty, a single vertex or a single edge. For cubic embedded graphs this is equivalent to the dual graph being simple. The argument of crossing Jordan curves that we will use in the proof was first published by Thomassen in [4], but also known to Robertson and later used by Mohar and Robertson in [2]. See also Theorem 5.7.1 in [3]. In fact, in [4] the argument was used to prove that 3-connected planar graphs embedded with genus g > 0 have facewidth at most 2. Together with Whitney’s theorem, this implies Theorem 1.2. We will give every detail of the proof in order to make it well suited for lectures presenting Whitney’s theorem, but the arguments are exactly the same arguments of crossing Jordan curves that Thomassen used – only that here the planar case, that is: Whitney’s theorem – is included too. Theorem 1.2. A 3-connected planar graph has an – up to equivalence – unique polyhedral embedding. Proof. Let G be a plane embedding of a 3-connected planar graph with mirror image G−1 and let G′ be an embedding different from these two. We say that a vertex of G′ has type 1 if the order is the same as in G, type −1 if it is the same as in G−1 and type 2 otherwise. As G′ is neither G nor G−1, G′ has a vertex of type 2 or an edge with one vertex of type 1 and one vertex of type −1. Assume first that there is a vertex v of type 2. Let e0, . . . , ed−1 be the order of edges around v in G′. If {e0, e1} is not an angle of G, we take this set of edges. Otherwise assume w.l.o.g. that e1 = nx(e0) in G and let j be minimal so that in G we have nx(ej) ̸= e(j+1) (mod d). As in G−1 we have nx(ej) = ej−1, the edge e(j+1) (mod d) follows ej neither in G nor in G′, so {ej , e(j+1) (mod d)} is not an angle in G. W.l.o.g. assume j = 0. G. Brinkmann: A simple and elementary proof of Whitney’s unique embedding theorem 197 So the order around v in G is e0, ei1 , . . . , eij , e1, eij+1 , . . . , eid−2 with 1 ≤ j < d − 2 and assume w.l.o.g. that ed−1 ∈ {eij+1 , . . . , eid−2}. Let y = max{i1, . . . , ij}, so y < d−1 and (y+1) ∈ {ij+1, . . . , id−2}, which implies that {ey, ey+1} is an angle of G′ with ey ∈ {ei1 , . . . , eij} and ey+1 ∈ {eij+1 , . . . , eid−2}. Let F be the facial cycle in G′ containing the angle {e0, e1} and F ′ be the facial cycle containing {ey, ey+1}. We have F ̸= F ′ as otherwise the faces would not be simple cycles. In G these cycles are not facial cycles, but two Jordan curves crossing each other in v. Due to the Jordan curve theorem, there must be a second crossing, so F, F ′ are two facial cycles that have at least two vertices in common that are not endpoints of a common edge – a contradiction to G′ being polyhedral. Assume now that all vertices are of type 1 or type −1. Then there is an edge e0 with one vertex of type 1 and one of type −1. Assume that in G the orientation around the type 1 vertex of e0 is e0, e1, . . . , ed and around the type −1 vertex it is e−10 , e′1, . . . , e′d′ , so in G′ it is e0, e1, . . . , ed resp. e′d′ , e ′ d′−1, . . . , e −1 0 . In G ′ there is a face F containing e−1d , e0, e ′ d′ and another face F ′ containing e′1 −1 , e−10 , e1. In G the corresponding cycles are again no facial cycles but Jordan curves crossing each other (with one common edge), so like in the first case we get a contradiction from the fact that there must be a second intersection between F and F ′. As plane embeddings of 3-connected graphs are all polyhedral, this also implies Whit- ney’s theorem, but there are also other consequences that are worth mentioning. They follow already from Theorem 8.1 in [4]. Note that for graphs with 1- or 2-cut there are no polyhedral embeddings in any closed orientable surface. Corollary 1.3. • There are no polyhedral embeddings of planar graphs in any orientable surface but the plane. • There are no embeddings of cubic planar graphs with a simple dual in any orientable surface but the plane. ORCID iD Gunnar Brinkmann https://orcid.org/0000-0003-4168-0877 References [1] J. L. Gross and T. W. Tucker, Topological Graph Theory, Wiley-Interscience Series in Discrete Mathematics and Optimization, John Wiley & Sons, New York, 1987. [2] B. Mohar and N. Robertson, Planar graphs on nonplanar surfaces, J. Comb. Theory Ser. B 68 (1996), 87–111, doi:10.1006/jctb.1996.0058. [3] B. Mohar and C. Thomassen, Graphs on Surfaces, Johns Hopkins Studies in the Mathematical Sciences, Johns Hopkins University Press, Baltimore, Maryland, 2001. [4] C. Thomassen, Embeddings of graphs with no short noncontractible cycles, J. Comb. Theory Ser. B 48 (1990), 155–177, doi:10.1016/0095-8956(90)90115-g. [5] H. Whitney, 2-isomorphic graphs, Amer. J. Math. 55 (1933), 245–254, doi:10.2307/2371127. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 199–208 https://doi.org/10.26493/1855-3974.2043.fbb (Also available at http://amc-journal.eu) The average genus for bouquets of circles and dipoles Jinlian Zhang School of Mathematics and Statistics, Hunan University of Finance and Economics, Changsha, P. R. China Xuhui Peng * MOE-LCSM, School of Mathematics and Statistics, Hunan Normal University, Changsha, P. R. China Yichao Chen † Department of Mathematics and Physics, Suzhou University of Science and Technology, Suzhou, P. R. China Received 9 July 2019, accepted 6 February 2021, published online 29 October 2021 Abstract The bouquet of circles Bn and dipole graph Dn are two important classes of graphs in topological graph theory. For n ≥ 1, we give an explicit formula for the average genus γavg(Bn) of Bn. By this expression, one easily sees γavg(Bn) = n−lnn−γ+1−ln 22 + o(1), where γ is the Euler-Mascheroni constant. Similar results are obtained for Dn. Our method mainly depends on the technique of generating series and the knowledge in ordinary differ- ential equations. Keywords: Average genus, bouquet of circles, dipole, ordinary differential equation. Math. Subj. Class. (2020): 05C10 *Corresponding author. Xuhui Peng was supported by NNSFC (No. 12071123), the Scientific Research Project of Hunan Province Education Department (No. 20A329) and the Construct Program of the Key Disci- pline in Hunan Province. †Yichao Chen was supported by NNSFC under Grant No. 11471106. E-mail addresses: jinlian916@hnu.edu.cn (Jinlian Zhang), xhpeng@hunnu.edu.cn (Xuhui Peng), ycchen@hnu.edu.cn (Yichao Chen) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 200 Ars Math. Contemp. 20 (2021) 199–208 1 Introduction and main results A graph G = (V (G), E(G)) is permitted to have both loops and multiple edges. An embedding of a graph G into an orientable surface Ok is a cellular embedding, i.e., the interior of every face is homeomorphic to an open disc. The subscript in Ok is the genus of the orientable surface Ok, for k ≥ 0. We denote the number of cellular embeddings of G on the surface Ok by gk(G), where, by the number of embeddings, we mean the number of equivalence classes under ambient isotopy. The genus polynomial of a graph G is given by ΓG(x) = ∑ k≥0 gk(G)x k. This sequence {gk(G), k = 0, 1, 2, . . .} is called the genus distribution of the graph G. For a graph G, it is well known that the total number of cellular embeddings is∏ v∈V (G)(dG(v) −1)!, where dG(v) is the degree of the vertex v in G. For example, see [13, Chapter 3]. Hence, ΓG(1) = ∑ k≥0 gk(G) = ∏ v∈V (G) (dG(v)− 1)!. (1.1) The average genus γavg(G) of the graph G is the expected value of the genus random variable, over all labeled 2-cell orientable embeddings of G, using the uniform distribution. In other words, the average genus of G is γavg(G) = Γ′G(1) ΓG(1) = ∞∑ k=0 k · gk(G) ΓG(1) . The study of the average genus of a graph began by Gross and Furst [9], and was much further developed by Chen and Gross [1, 2, 3]. Two lower bounds were obtained in [4] for the average genus of two kinds of graphs. In [19], Stahl gave the asymptotic result for the average genus of linear graph families. The exact values for the average genus of small- order complete graphs, closed-end ladders, and cobblestone paths were derived by White [22]. More references are the following: [5, 10, 15, 17, 20] etc. For a general background in topological graph theory, we refer the reader to see Gross and Tucker [13] or White [21]. One of the purposes of the paper is to give an explicit expression of the average genus for a bouquet of circles. By a bouquet of circles, or more briefly, a bouquet, we mean a graph with one vertex and some self-loops. In particular, the bouquet with n self-loops is denoted by Bn. Figure 1 shows the graphs B1, B2, B3. The bouquets {Bn, n ≥ 1} are very important graphs in topological graph theory. First, since any connected graph can be reduced to a bouquet by contracting a spanning tree to a point, bouquets are fundamental building blocks of topological graph theory. Second, as shown in [8, 12], Cayley graphs and many other regular graphs are covering spaces of bouquets. For the genus distribution of Bn, Gross, Robbins and Tucker [11] proved that the num- bers gk(Bn) of embeddings of the Bn in an oriented surface of genus k satisfy the following recurrence for n > 2, (n+ 1)gk(Bn) = 4(2n− 1)(2n− 3)(n− 1)2(n− 2)gk−1(Bn−2) + 4(2n− 1)(n− 1)gk(Bn−1) (1.2) J. Zhang et al.: The average genus for bouquets of circles and dipoles 201 1B 2B 3B Figure 1: The bouquets B1, B2, and B3. with initial conditions gk(B0) = 1 for k = 0 and gk(B0) = 1 for k > 0, gk(B1) = 1 for k = 0 and gk(B1) = 1 for k > 0. (1.3) With the aid of an edge-attaching surgery technique, the total embedding polynomial of Bn was computed in [14]. Stahl [18] also did some research on the average genus of Bn. By [18, Theorem 2.5] and the definition of Euler-Mascheroni constant, one easily sees that lim n→∞ ( γavg(Bn)− ( n+ 1 2 − 1 2 2n∑ k=1 1 k )) = 0. (1.4) To achieve this, Stahl made many accurate estimates on the unsigned Stirling numbers s(n, k) of the first kind. In this paper, using knowledge in ordinary differential equations and Taylor’s formula, we derive an explicit expression of γavg(Bn). By this expression, (1.4) follows immediately. Our methods are totally different from that in [18] and we do not need to make estimates on s(n, k). In Section 2, we will give the computation of γavg(Bn) in detail. A dipole with n edges, denoted by Dn, has two vertices joined by n edges. Figure 2 shows the graphs D1, D2, D3. 1D 2D 3D Figure 2: The dipoles D1, D2, and D3. Another purpose of this paper is to give an explicit expression of the average genus for dipole Dn. The dipole, like the bouquet, is useful as a voltage graph. See [21] for example. Moreover, hypermaps correspond with the 2-cell embeddings of the dipole. The genus distribution of Dn is given by [14] and [16]. In Lemma 2.1 below, we obtain the following recurrence relation for γavg(Bn) (n+ 1)γavg(Bn) = 2γavg(Bn−1) + (n− 1) ( γavg(Bn−2) + 1 ) . (1.5) The most popular way to deal with sequences of numbers is to manipulate infinite series that “generate” those sequences. For instance, see [6, 7]. We apply this method to 202 Ars Math. Contemp. 20 (2021) 199–208 the calculation of γavg(Bn). Multiplying both sides of (1.5) by tn and summing on n ≥ 1, the generating function u(t) = ∑ n≥1 γavg(Bn)t n will satisfy an ordinary differential equation. We solve this differential equation with the aid of a computer system and find an explicit expression for γavg(Bn) by expanding u(t) as a power series in t. The calculation of γavg(Dn) is similar to that in γavg(Bn). But the processes are more complicated, so we still give their details in Section 3. 2 The average genus of Bn We begin by proving the following lemma. Lemma 2.1. The following recurrence relation holds for the average genus γavg(Bn) of Bn (n+ 1)γavg(Bn) = 2γavg(Bn−1) + (n− 1) ( γavg(Bn−2) + 1 ) (2.1) with initial conditions γavg(B1) = 0, γavg(B2) = 13 . Proof. Multiplying both sides of (1.2) by xk and summing on k ≥ 0, it holds that∑ k≥0 (n+ 1)gk(Bn)x k = ∑ k≥0 4(2n− 1)(2n− 3)(n− 1)2(n− 2)gk−1(Bn−2)xk + ∑ k≥0 4(2n− 1)(n− 1)gk(Bn−1)xk. (2.2) Hence, the genus polynomial ΓBn(x) satisfies the following recurrence (n+ 1)ΓBn(x) = 4(2n− 1)(2n− 3)(n− 1)2(n− 2) · x · ΓBn−2(x) + 4(2n− 1)(n− 1)ΓBn−1(x). (2.3) Differentiating both sides of (2.3) and taking x = 1 lead to (n+ 1)Γ′Bn(1) = 4(2n− 1)(2n− 3)(n− 1) 2(n− 2) · Γ′Bn−2(1) + 4(2n− 1)(2n− 3)(n− 1)2(n− 2) · ΓBn−2(1) + 4(2n− 1)(n− 1)Γ′Bn−1(1). Applying (1.1) to the graph Bn yields ΓBn(1) = (2n − 1)!. Dividing both sides of the above equality by ΓBn(1), by the definition of average genus, one arrives at (n+ 1)γavg(Bn) = 2γavg(Bn−1) + (n− 1) ( γavg(Bn−2) + 1 ) . A direct calculation gives rise to γavg(B1) = 0 and γavg(B2) = 13 . The proof is com- pleted. The main purpose of this section is to prove the following theorem. Theorem 2.2. The average genus of Bn is given by γavg(Bn) = n+ 1 2 − n−1∑ m=0 1 + (−1)m 2(m+ 1) − 1 + (−1) n 4(n+ 1) . (2.4) In particular, we have γavg(Bn) = n− lnn− γ + 1− ln 2 2 + o(1), where γ ≈ 0.5772 is the Euler-Mascheroni constant. J. Zhang et al.: The average genus for bouquets of circles and dipoles 203 Proof. For n ≤ 0, we define γavg(Bn) = 0 so that (2.1) holds for any integer n ≥ 1. For the simplicity of writing, we use an to denote γavg(Bn) in the proof. Multiplying both sides of (2.1) by tn and summing on n ≥ 1, we obtain∑ n≥1 (n+ 1)ant n = 2 ∑ n≥1 an−1t n + ∑ n≥1 (n− 1)(an−2 + 1)tn. (2.5) Let u(t) = ∑ n≥1 ant n. Then, with the help of (2.5), we obtain( t · ∑ n≥1 ant n )′ = 2t · ∑ n≥1 an−1t n−1 + ∑ n≥1 (n− 2)an−2tn + ∑ n≥1 an−2t n + ∑ n≥1 (n− 1)tn = 2tu(t) + t3 ∑ n≥1 (n− 2)an−2tn−3 + t2u(t) + t2 · (∑ n≥2 tn−1 )′ , that is (tu(t))′ = 2tu(t) + t3 ∑ n≥3 (n− 2)an−2tn−3 + t2u(t) + t2 ( t 1− t )′ = 2tu(t) + t3 ∑ n≥1 nant n−1 + t2u(t) + t2 ( t 1− t )′ = 2tu(t) + t3u′(t) + t2u(t) + t2 ( t 1− t )′ , which implies that u(t) satisfies the following equation (t− t3)u′(t) + (1− 2t− t2)u(t) = t 2 (1− t)2 (2.6) with initial condition u(0) = 0. Since the above equation is a first order linear differential equation, we can solve it directly and obtain its solution: u(t) = − ( t2 − 1 ) ln(1− t) + ( t2 − 1 ) ln(t+ 1) + 2t 4(t− 1)2t . Denote u1(t) = 1 2(t− 1)2 , u2(t) = − (t+ 1) ln(1− t) 4(t− 1)t , u3(t) = (t+ 1) ln(t+ 1) 4(t− 1)t . Then, clearly, u(t) = u1(t) + u2(t) + u3(t). Using Taylor’s formula, we get u1(t) = ∑ n≥0 n+ 1 2 tn (2.7) and u2(t) = 1 4 (1 + t) · 1 1− t · ln(1− t) t = 1 4 (1 + t) · ∑ ℓ≥0 tℓ · ∑ m≥0 ( − 1 m+ 1 tm ) 204 Ars Math. Contemp. 20 (2021) 199–208 = 1 4 (1 + t) · ∑ n≥0 n∑ m=0 ( − 1 m+ 1 ) tn = ∑ n≥0 bnt n, (2.8) where b0 = − 14 and bn = 1 4 [∑n m=0(− 1 m+1 ) + ∑n−1 m=0(− 1 m+1 ) ] , n ≥ 1. Also by the Taylor’s formula, u3(t) = − 1 4 (1 + t) · 1 1− t · ln(1 + t) t = −1 4 (1 + t) · ∑ ℓ≥0 tℓ · ∑ m≥0 (−1)m m+ 1 tm = −1 4 (1 + t) · ∑ n≥0 n∑ m=0 (−1)m m+ 1 tn = ∑ n≥0 cnt n, (2.9) where c0 = − 14 and cn = − 1 4 [ n∑ m=0 (−1)m m+ 1 + n−1∑ m=0 (−1)m m+ 1 ] , n ≥ 1. It follows from (2.7) – (2.9) that an = n+ 1 2 + bn + cn = n+ 1 2 + 1 4 [ n∑ m=0 ( − 1 m+ 1 ) + n−1∑ m=0 ( − 1 m+ 1 )] − 1 4 [ n∑ m=0 (−1)m m+ 1 + n−1∑ m=0 (−1)m m+ 1 ] , which yields (2.4). In view of γ = lim n→+∞ [ n∑ m=0 1 m+ 1 − lnn ] and lim n→+∞ n−1∑ m=0 (−1)m m+ 1 = ln 2, (2.10) we complete the proof of (2.2). 3 The average genus of Dn Our first purpose is to show the following lemma. Lemma 3.1. The following recurrence relation holds for the average genus γavg(Dn) of Dn n(n+ 2)γavg(Dn+1) = (2n+ 1)γavg(Dn) + (n 2 − 1) · γavg(Dn−1) + n2 (3.1) with initial conditions γavg(D1) = γavg(D2) = 0. Proof. By [16, Theorem 5.2], we obtain (n+2)gk(Dn+1) = n(2n+1)gk(Dn) +n 3(n− 1)2gk−1(Dn−1)−n(n− 1)2gk(Dn−1). Applying (1.1) to the graph Dn+1 yields ΓDn+1(1) = (n!) 2. Following the lines in the proof of Lemma 2.1, we derive the recurrence relation (3.1). The initial conditions γavg(D1) = γavg(D2) = 0 are due to a direct calculation. The proof is finished. J. Zhang et al.: The average genus for bouquets of circles and dipoles 205 The main purpose of this section is to prove the following theorem. Theorem 3.2. γavg(D1) = γavg(D2) = 0 and for n ≥ 3, we have γavg(Dn) = n [ 1 2 n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 2)(m− 1)m + 1 6 ] − 1 2 n+1∑ m=1 1 m − n+1∑ m=4 (−1)m(2m2 − 6m+ 3) (m− 3)(m− 1)m + 7 12 . (3.2) In particular, we have γavg(Dn) = n− lnn− γ 2 + o(1), (3.3) where γ ≈ 0.5772 is the Euler-Mascheroni constant. Proof. First, we give a proof of (3.2). For the simplicity of writing, we use an to denote γavg(Dn) in the proof. Let u(t) = ∑ n≥1 ant n−3 = ∑ n≥2 an+1t n−2. Multiplying both sides of (3.1) by tn−2 and summing on n ≥ 2, we obtain∑ n≥2 n(n+ 2)an+1t n−2 = ∑ n≥2 (2n+ 1)ant n−2 + ∑ n≥2 (n2 − 1)an−1tn−2 + ∑ n≥2 n2tn−2. (3.4) Since u′(t) = ∑ n≥2 (n− 2)an+1tn−3, u′′(t) = ∑ n≥2 (n− 2)(n− 3)an+1tn−4, it follows that∑ n≥2 n(n+ 2)an+1t n−2 = ∑ n≥2 [ (n− 2)(n− 3) + 7(n− 2) + 8 ] an+1t n−2 = t2u′′(t) + 7tu′(t) + 8u(t),∑ n≥2 (2n+ 1)ant n−2 = ∑ n≥2 (2n+ 3)an+1t n−1 = ∑ n≥2 ( 2(n− 2) + 7 ) an+1t n−1 = 2t2u′(t) + 7tu(t),∑ n≥2 (n2 − 1)an−1tn−2 = ∑ n≥4 (n2 − 1)an−1tn−2 = ∑ n≥2 (n2 + 4n+ 3)an+1t n = ∑ n≥2 [ (n− 2)(n− 3) + 9(n− 2) + 15 ] an+1t n = t4u′′(t) + 9t3u′(t) + 15t2u(t),∑ n≥2 n2tn−2 = ∑ n≥2 n(n− 1)tn−2 + ∑ n≥2 ntn−2 = v′′(t) + ∑ n≥0 ntn−2 − t−1 = v′′(t) + v′(t) t − t−1 = 3t− 4− t 2 (t− 1)3 , 206 Ars Math. Contemp. 20 (2021) 199–208 where v(t) = ∑ n≥0 t n, v′(t) = ∑ n≥0 nt n−1, v′′(t) = ∑ n≥0 n(n− 1)tn−2. Substituting the above equalities into (3.4), u(t) satisfies the following second order linear differential equation (t2 − t4)u′′(t) + (7t− 2t2 − 9t3)u′(t) + (8− 7t− 15t2)u(t) = 3t− 4− t 2 (t− 1)3 with initial conditions u(0) = a3 = γavg(D3) = 12 , u ′(0) = a4 = γavg(D4) = 5 6 . With the help of a computer algebra systems, the solution of the above equation is u(t) = 1 4(t− 1)t2 + w(t) 4(t− 1)2t4 , (3.5) where w(t) = −t3 + 2t3 ln(t+ 1) + 3t2 − 2t2 ln(t+ 1) − 2t ln(1− t)− 2t ln(t+ 1) + 2 ln(1− t) + 2 ln(t+ 1). By Taylor’s formula, we get 1 4(t− 1)t2 = ∑ m≥−2 ( − 1 4 ) tm, w(t) = t2 − t3 + ∑ m≥4 2 ( 4(−1)mm2 +m2 − 12(−1)mm− 5m+ 6(−1)m + 6 ) (m− 3)(m− 2)(m− 1)m tm, 1 4(t− 1)2t4 = ∑ m≥−4 m+ 5 4 tm. Therefore, comparing the coefficients of tn−3 of the both sides of (3.5) gives an = − 1 4 + n 4 − n− 1 4 + n+1∑ m=4 2 ( 4(−1)mm2 +m2 − 12(−1)mm− 5m+ 6(−1)m + 6 ) (m− 3)(m− 2)(m− 1)m · n−m+ 2 4 = n 2 n+1∑ m=4 [ (−1)m(4m2 − 12m+ 6) (m− 3)(m− 2)(m− 1)m + (m2 − 5m+ 6) (m− 3)(m− 2)(m− 1)m ] − n+1∑ m=4 (−1)m(4m2 − 12m+ 6) + (m2 − 3m) + (−2m+ 6) (m− 3)(m− 2)(m− 1)m · m− 2 2 = n 2 n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 2)(m− 1)m + n 2 n+1∑ m=4 1 (m− 1)m − 1 2 n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 1)m − 1 2 n+1∑ m=4 1 m− 1 + n+1∑ m=4 1 m(m− 1) J. Zhang et al.: The average genus for bouquets of circles and dipoles 207 = n 2 n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 2)(m− 1)m + n 2 (1 3 − 1 n+ 1 ) − 1 2 n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 1)m − 1 2 n+1∑ m=1 1 m + 3 4 + 1 2(n+ 1) + (1 3 − 1 n+ 1 ) which yields the desired result (3.2). Now we are in a position to prove (3.3). Using the software Mathematica or series theory, one has n+1∑ m=4 (−1)m(4m2 − 12m+ 6) (m− 3)(m− 2)(m− 1)m = 2 3 + o ( 1 n ) (3.6) and n+1∑ m=4 (−1)m(2m2 − 6m+ 3) (m− 3)(m− 1)m = 7 12 + o(1). (3.7) Combining (3.6) – (3.7), (2.10) and (3.2), we complete the proof of (3.3). 4 Some remarks Bouquets and dipoles are two important classes of graphs in topological graph theory. Their average genera are of independent interest. In this paper, we obtain explicit formulas for γavg(Bn) and γavg(Dn). By Theorems 2.2 and 3.2, we have the following relation between γavg(Bn) and γavg(Dn), γavg(Bn) = γavg(Dn) + 1− ln 2 2 + o(1). It follows that the difference of γavg(Bn) and γavg(Dn) tends to the constant 1−ln 22 when n tends to infinity. Since both Bn and Dn are upper-embeddable, the maximum genera of Bn and Dn are⌊ n 2 ⌋ and ⌊ n−1 2 ⌋ , respectively. Recall that the minimum genera of Bn and Dn equal 0. Therefore, also by Theorems 2.2 and 3.2, we have lim n→∞ γavg(Bn) ⌊n2 ⌋ = 1 and lim n→∞ γavg(Dn) ⌊n−12 ⌋ = 1. This implies that the average genus of Bn (Dn) is closer to the maximum genus than to the minimum genus. ORCID iDs Xuhui Peng https://orcid.org/0000-0002-2443-4896 References [1] J. Chen, A linear-time algorithm for isomorphism of graphs of bounded average genus, SIAM J. Discrete Math. 7 (1994), 614–631, doi:10.1137/s0895480191196769. 208 Ars Math. Contemp. 20 (2021) 199–208 [2] J. Chen and J. L. Gross, Limit points for average genus. I. 3-connected and 2-connected sim- plicial graphs, J. Comb. Theory Ser. B 55 (1992), 83–103, doi:10.1016/0095-8956(92)90033-t. [3] J. Chen and J. L. Gross, Kuratowski-type theorems for average genus, J. Comb. Theory Ser. B 57 (1993), 100–121, doi:10.1006/jctb.1993.1009. [4] J. Chen, J. L. Gross and R. G. Rieper, Lower bounds for the average genus, J. Graph Theory 19 (1995), 281–296, doi:10.1002/jgt.3190190302. [5] Y. Chen, Lower bounds for the average genus of a CF-graph, Electron. J. Combin. 17 (2010), #R150 (14 pages), doi:10.37236/422. [6] L. Comtet, Advanced Combinatorics, Springer Netherlands, 1974, doi:10.1007/ 978-94-010-2196-8. [7] R. L. Graham, D. E. Knuth and O. Patashnik, Concrete Mathematics: A Foundation for Com- puter Science, Addison-Wesley, Reading, MA, 1989. [8] J. L. Gross, Every connected regular graph of even degree is a Schreier coset graph, J. Comb. Theory Ser. B 22 (1977), 227–232, doi:10.1016/0095-8956(77)90068-5. [9] J. L. Gross and M. L. Furst, Hierarchy for imbedding-distribution invariants of a graph, J. Graph Theory 11 (1987), 205–220, doi:10.1002/jgt.3190110211. [10] J. L. Gross, E. W. Klein and R. G. Rieper, On the average genus of a graph, Graphs Combin. 9 (1993), 153–162, doi:10.1007/bf02988301. [11] J. L. Gross, D. P. Robbins and T. W. Tucker, Genus distributions for bouquets of circles, J. Comb. Theory Ser. B 47 (1989), 292–306, doi:10.1016/0095-8956(89)90030-0. [12] J. L. Gross and T. W. Tucker, Generating all graph coverings by permutation voltage assign- ments, Discrete Math. 18 (1977), 273–283, doi:10.1016/0012-365x(77)90131-5. [13] J. L. Gross and T. W. Tucker, Topological Graph Theory, Wiley-Interscience Series in Discrete Mathematics and Optimization, John Wiley & Sons, New York, 1987. [14] J. H. Kwak and S. H. Shim, Total embedding distributions for bouquets of circles, Discrete Math. 248 (2002), 93–108, doi:10.1016/s0012-365x(01)00187-x. [15] K. McGown and A. Tucker, Statistics of genus numbers of cubic fields, 2016, arXiv:611.07088 [math.NT]. [16] R. Riper, The enumeration of graph embeddings, Ph.D. thesis, Western Michigan University, 1990, https://scholarworks.wmich.edu/dissertations/2105. [17] S. Stahl, The average genus of classes of graph embeddings, Congr. Numer. 40 (1983), 375– 388, proceedings of the Fourteenth Southeastern Conference on Combinatorics, Graph Theory and Computing (Boca Raton, Florida, 1983). [18] S. Stahl, Region distributions of graph embeddings and Stirling numbers, Discrete Math. 82 (1990), 57–78, doi:10.1016/0012-365x(90)90045-j. [19] S. Stahl, Permutation-partition pairs. III. Embedding distributions of linear families of graphs, J. Comb. Theory Ser. B 52 (1991), 191–218, doi:10.1016/0095-8956(91)90062-o. [20] S. Stahl, Bounds for the average genus of the vertex-amalgamation of graphs, Discrete Math. 142 (1995), 235–245, doi:10.1016/0012-365x(93)e0221-o. [21] A. T. White, Graphs, Groups and Surfaces, volume 8 of North-Holland Mathematics Studies, North-Holland Publishing Company, Amsterdam, 2nd edition, 1984. [22] A. T. White, An introduction to random topological graph theory, Combin. Probab. Comput. 3 (1994), 545–555, doi:10.1017/s0963548300001395. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 209–222 https://doi.org/10.26493/1855-3974.2465.571 (Also available at http://amc-journal.eu) Well-totally-dominated graphs* Selim Bahadır Department of Mathematics, Ankara Yıldırım Beyazıt University, Ankara, Turkey Tınaz Ekim Department of Industrial Engineering, Boğaziçi University, İstanbul, Turkey Didem Gözüpek Department of Computer Engineering, Gebze Technical University, Kocaeli, Turkey Received 22 October 2020, accepted 19 February 2021, published online 29 October 2021 Abstract A subset of vertices in a graph is called a total dominating set if every vertex of the graph is adjacent to at least one vertex of this set. A total dominating set is called minimal if it does not properly contain another total dominating set. In this paper, we study graphs whose all minimal total dominating sets have the same size, referred to as well-totally- dominated (WTD) graphs. We first show that WTD graphs with bounded total domination number can be recognized in polynomial time. Then we focus on WTD graphs with total domination number two. In this case, we characterize triangle-free WTD graphs and WTD graphs with packing number two, and we show that there are only finitely many planar WTD graphs with minimum degree at least three. Lastly, we show that if the minimum degree is at least three then the girth of a WTD graph is at most 12. We conclude with several open questions. Keywords: Total domination, well-totally-dominated graphs, minimal total dominating sets. Math. Subj. Class. (2020): 05C69 *This work has been supported by the Scientific and Technological Research Council of Turkey (TUBITAK) under grant no. 118E799. The work of Didem Gözüpek was also supported by the BAGEP Award of the Science Academy of Turkey. E-mail addresses: sbahadir@ybu.edu.tr (Selim Bahadır), tinaz.ekim@boun.edu.tr (Tınaz Ekim), didem.gozupek@gtu.edu.tr (Didem Gözüpek) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 210 Ars Math. Contemp. 20 (2021) 209–222 1 Introduction Total domination in graphs has been extensively studied in the literature (see [15]) and has numerous applications. For instance, consider a computer network where a core group of file servers has the ability to communicate directly with every computer outside the core group. Moreover, each file server is directly linked to at least one other backup file server where duplicate information is stored. This core group of servers corresponds to a total dominating set in the graph representing the computer network. Another application area is a specific committee selection mechanism such that every non-member of the committee knows at least one member of the committee and every member of the committee knows at least one other member of the committee to avoid feelings of isolation and thus enhance cooperation (see [14]). Let G be a graph with no isolated vertices. A subset S of V (G) is called a total dom- inating set (TDS) of G if every vertex in G is adjacent to at least one element in S. A total dominating set is minimal if it contains no other TDS of G. The minimum size of a total dominating set of a graph G is called the total domination number and denoted by γt(G), while the maximum size of a minimal total dominating set is called the upper total domination number and denoted by Γt(G). G is called well-totally-dominated (WTD) if every minimal TDS of G is of the same size, that is, γt(G) = Γt(G). WTD graphs with γt = k are denoted by WTD(k). Given a graph, computing its total domination number and its upper total domination number are NP-hard in general [6, 18] and already NP-hard even in specific graph classes such as bipartite graphs, comparability graphs and claw-free graphs [15]. One way to deal with such a problem is to consider “trivial” instances where these two paramaters have the same value. Examples of graph classes defined in this way in the literature include well- covered graphs (whose all maximal independent sets have the same size), well-dominated graphs (whose all minimal dominating sets have the same size), and equimatchable graphs (whose all maximal matchings have the same size). Structural properties of each one of these graph classes have been studied extensively in the literature. In this paper, we take the same approach for the total dominating sets. Works on total domination in the literature mostly focused on the relation of the total domination number with other graph parame- ters and characterized graphs with total domination number being equal to an upper bound (e.g. [3, 4]). Inequalities relating the total domination number to other domination param- eters and characterization of graphs that tightly attain these bounds have also been studied (see [1, 16]). Clearly, if the total domination number and the upper total domination number are polynomial time solvable for a given class of graphs, then the recognition of WTD graphs belonging to this class of graphs is polynomial. However, the complexity of recognizing WTD graphs in general is unknown. In such a situation, a classical approach consists in studying the structure of WTD graphs in restricted graph classes and providing structural characterizations along with efficient recognition algorithms whenever possible. WTD graphs were initially introduced in [12], where WTD cycles and paths are char- acterized and several constructions of WTD trees are given. They also proved that a WTD graph with minimum degree at least two has girth at most 14. The work in [7] focused on the composition and decomposition of WTD trees and proved that any WTD tree can be constructed from a family of three small trees. To the best of our knowledge, [12] and [7] are the only work on WTD graphs. A graph class resembling WTD graphs is well- dominated graphs, which are graphs whose minimal dominating sets have the same size. It S. Bahadır et al.: Well-totally-dominated graphs 211 is known that well-dominated graphs form a proper subset of well-covered graphs [8]. We note that well-covered graphs are graphs whose maximal independent sets have the same size and there is a rich literature about them (see [13, 19]). Well-dominated graphs were in- troduced by Finbow et al. [8], who provided a characterization of bipartite well-dominated graphs and well-dominated graphs with girth at least 5. Characterizations of these graphs within other graph classes were also obtained [9, 10, 17, 20]. Although their definitions re- semble each other, there is not a containment relationship between WTD graphs and well- dominated graphs. For instance, a cycle on six vertices is WTD but not well-dominated, whereas the graph T10 described in [17] is well-dominated but not WTD. It follows from the previous studies on WTD graphs that we do not know much about their structure. In this paper, we investigate the study of WTD graphs from a structural point of view. We first study WTD graphs with bounded total domination number. We prove in Section 2 that the recognition of WTD graphs with total domination number k is solvable in polynomial time for every positive integer k. We then focus on WTD graphs with total domination number 2, referred to as WTD(2) graphs in Section 3. We char- acterize triangle-free WTD(2) graphs and WTD(2) graphs with packing number 2 (or equivalently of diameter 3). We also show that there is a finite number of planar WTD(2) graphs with minimum degree at least 3. Subsequently, we study the girth of WTD graphs in Section 4. In particular, building on a result in [12], we prove that WTD graphs with mini- mum degree at least three have girth at most 12. Finally, we discuss several open research directions. 2 WTD graphs with bounded total domination number Recall that the complexity of recognizing WTD graphs is unknown. In this section, we show that for any positive integer k, WTD(k) graphs can be recognized in polynomial time. To this end, we will use an equivalent description of WTD(k) graphs using transversal hypergraphs. Let us first introduce necessary definitions. A hypergraph H is a pair H = (X,E) where X is a set of elements called vertices, and E is a set of nonempty subsets of X called hyperedges. Therefore, a hypergraph might have a vertex which belongs to none of the hyperedges, but cannot have multiple hyperedges. A transversal (or hitting set) of a hypergraph H = (X,E) is a set T ⊆ X that has nonempty intersection with every hyperedge of H . A transversal of a collection of sets is a transversal of the hypergraph whose hyperedges are the given collection. A transversal T is called minimal if no proper subset of T is a transversal. The transversal hypergraph of H = (X,E) is the hypergraph H∗ = (X,F ) whose hyperedge set F consists of all minimal transversals of H . Let G be a graph with no isolated vertex. Let HG be the hypergraph whose vertex set is V (G) and hyperedges are open neighborhoods of the vertices of G. Let also MTDS(G) denote the set of all minimal total dominating sets of G. Lemma 2.1. MTDS(G) consists of hyperedges of the transversal hypergraph of HG. Proof. Let T be a hyperedge of H∗G, that is a minimal transversal of the set of open neigh- borhoods of G. This means that T contains a neighbor of every vertex in G, thus it is a total dominating set. By minimality of the transversal T , it is also a minimal total dominating set of G. Conversely, let S be a minimal total dominating set of G. Then, every vertex in G is adjacent to at least one vertex in S. That is, S has a nonempty intersection with every open neighborhood in G. Therefore, S is a transversal of the hypergraph HG and minimality of S implies that it is a minimal transversal. Thus, S is a hyperedge of H∗G. 212 Ars Math. Contemp. 20 (2021) 209–222 Proposition 2.2. Let G be a graph. Then, for any minimal transversal T of MTDS(G), there exists a vertex v in G such that N(v) = T . Proof. Let MTDS(G) = {A1, . . . , Am}. Since T has nonempty intersection with each Ai, V (G) \ T contains none of the minimal total dominating sets A1, . . . , Am. Therefore, V (G) \ T is not a TDS of G, and hence there exists at least one vertex v ∈ V (G) such that N(v) ∩ (V (G) \ T ) = ∅. Thus, we see that N(v) ⊆ T . Suppose that N(v) ̸= T . Then T \N(v) ̸= ∅ and let u ∈ T \N(v). Since T is a minimal transversal, T \ {u} is disjoint with at least one of A1, . . . , Am, say A1. As u ∈ T \ N(v), we have N(v) ⊆ T \ {u}, and hence N(v) ∩ A1 = ∅. That is, v is not dominated by A1, which is a contradiction. Therefore, N(v) = T . A hypergraph H is said to be Sperner if no hyperedge of H contains another hyperedge. The following result shows that any finite collection of finite sets which forms a Sperner hypergraph corresponds to the set of all minimal total dominating sets of a graph. Proposition 2.3. Let H be a Sperner hypergraph. Then there exists a graph G such that E(H) = MTDS(G). Proof. Let E(H) = {A1, . . . , Am} and A = ∪mi=1Ai. Consider a graph with vertex set A and draw edges between its vertices such that each vertex is adjacent to at least one vertex in Ai for all i = 1, . . . ,m (for example, draw all possible edges). Then, in accordance with Proposition 2.2, for each minimal transversal T of H , add a vertex vT to the graph such that N(vT ) = T . Let G be the resulting graph. We first show that each Ai is a TDS of G. By construction, every vertex of A is adjacent to at least one vertex in Ai. Moreover, for every minimal transversal T of A1, . . . , Am we have T ∩ Ai ̸= ∅, and hence, each vT is dominated by Ai. Therefore, Ai is a TDS for i = 1, . . . ,m. We next show that every TDS of G contains at least one of A1, . . . , Am. Let S be a TDS of G and suppose that Ai ⊈ S for i = 1, . . . ,m. Then, V (G) \ S is a transversal of A1, . . . , Am, and hence, there exists a minimal transversal T of A1, . . . , Am such that T ⊆ V (G) \ S. On the other hand, we have N(vT ) = T and thus, we get N(vT )∩ S = ∅, which contradicts that S is a TDS of G. Consequently, a set other than A1, . . . , Am can not be a minimal TDS of G. We finally show that each Ai is a minimal TDS of G. Suppose that Ai is not minimal for some i. Then, Ai \{x} is still a TDS of G for some x ∈ Ai, and therefore, Aj ⊆ Ai \{x} for some j, which implies Aj ⊆ Ai contradicting that H is Sperner. Therefore, minimal TDSs of G are exactly A1, . . . , Am. Remark 2.4. One can extend G to another graph whose minimal TDSs are A1, . . . , Am as follows: Let G′ be a graph disjoint from G. Draw edges between the vertices of G′ and A in such a way that every vertex of G′ is adjacent to at least one vertex of Ai for i = 1, . . . ,m. By following the same arguments, it is easy to check that minimal TDSs of the resulting graph are A1, . . . , Am. Notice that any finite collection consisting of distinct sets of size k corresponds to a Sperner hypergraph and therefore, Proposition 2.3 implies the following result. Corollary 2.5. For every integer k ≥ 2, WTD(k) is an infinite graph family. S. Bahadır et al.: Well-totally-dominated graphs 213 The HYPERGRAPH TRANSVERSAL problem is the decision problem that takes as input two Sperner hypergraphs H and H ′ and asks whether H ′ is the transversal hypergraph H∗ of H . Theorem 2.6 ([2, 5]). For every positive integer k, the HYPERGRAPH TRANSVERSAL problem is solvable in polynomial time if all hyperedges of one of the two hypergraphs H and H ′ are of size at most k. Theorem 2.6 has the following consequence: Corollary 2.7 ([11]). For every positive integer k, the following problem is solvable in polynomial time: Given a Sperner hypergraph H , determine whether all minimal transver- sals of H are of size k. The complexity of recognition of WTD graphs with bounded total domination number can now be derived from Corollary 2.7. Theorem 2.8. For every positive integer k, the problem of recognizing WTD(k) graphs can be solved in polynomial time. Proof. Let G be a graph with no isolated vertices. Consider the hypergraph HG = (V, E), where E contains the inclusion-minimal elements of {N(v) : v ∈ V }. Observe that HG is Sperner and that the minimal transversals of HG are exactly the minimal total dominating sets of G by Lemma 2.1. It follows that G is WTD if and only if all minimal transversals of HG are of size k. By Corollary 2.7, this condition can be tested in polynomial time. 3 WTD graphs with total domination number two In this section, we study WTD graphs whose total domination number is 2. We give complete characterizations of WTD(2) graphs with packing number 2 and triangle-free WTD(2) graphs. We also show that planar WTD(2) graphs with minimum degree at least 3 have at most 16 vertices. Let G be a WTD(2) graph. Note that every minimal TDS of G is a pair consisting of endpoints of an edge of G. Consequently, every WTD(2) graph is connected. We will call an edge of G whose endpoints is a TDS of G a dominating edge of G. Let Gde be the graph with vertex set ∪S∈MTDS(G)S (i.e., vertices of G serve as an endpoint of a dominating edge) and edge set which consists of dominating edges of G. In other words, Gde is the edge-induced subgraph of G obtained by the dominating edges. See Figure 1 for an example. G x y z t w y z t w Gde Figure 1: A WTD(2) graph G and the graph Gde obtained by the dominating edges of G. 214 Ars Math. Contemp. 20 (2021) 209–222 Remark 3.1. Notice that the graph Gde and the subgraph of G induced by V (Gde) are not necessarily the same. In general, Gde is a subgraph of G but not necessarily an induced subgraph of G with respect to a set of vertices. A set S is a vertex cover of a graph G if every edge of G has an endpoint from S. Let MVC(G) denote the set of all minimal vertex covers of the graph G. Proposition 3.2. Let G be a WTD(2) graph. For every minimal vertex cover S of Gde there exists a vertex vS in G such that N(vS) = S. Proof. We notice that every minimal vertex cover S of Gde is a minimal transversal of MTDS(G). Therefore, by Proposition 2.2 there exists a vertex in G whose neighborhood is exactly S. 3.1 Characterization of WTD(2) graphs with packing number 2 A set S ⊆ V (G) is called a packing of G if N [u] ∩N [v] = ∅ for every distinct u, v ∈ S. The packing number ρ(G) is the maximum size of a packing of G. It is well-known that for any graph G we have ρ(G) ≤ γ(G) ≤ γt(G). Therefore, if γt(G) = 2, then ρ(G) is either 1 or 2. In this subsection, we provide a characterization of WTD(2) graphs G with ρ(G) = 2. In particular, this characterization allows us to construct any WTD(2) graph with ρ(G) = 2. Let W2 be the set of graphs obtained as follows: Step 1: Choose a bipartite graph H with no isolated vertices. Step 2: For every S ∈ MVC(H), choose a new vertex vS and draw edges from vS to every vertex in S. Step 3: For each edge uv in H and every w ∈ V (H) \ {u, v}, add the edges wu and/or wv if needed to make sure w is adjacent to at least one of u and v. Step 4: Choose a new graph H ′ (might be the empty graph) which is disjoint from the current graph. Then for each edge uv in H and every w ∈ V (H ′), draw at least one of the edges wu and wv. A graph in W2 is given in Figure 2. H Step 2 v{x,z} v{x,t} v{y,z} v{y,t} x y z t x y z t Step 3 v{x,z} v{x,t} v{y,z} v{y,t} x y z t Step 4 v{x,z} v{x,t} v{y,z} v{y,t} x y z t u1 u2 u3 Figure 2: A graph in W2 obtained by the given process. Bold edges represent the dominat- ing edges. S. Bahadır et al.: Well-totally-dominated graphs 215 Lemma 3.3. If a graph G is in W2, then G is a WTD(2) graph with ρ(G) = 2. Proof. Let G ∈ W2 and H = (U, V,E) be the bipartite graph in the first step of the construction of G. We first show that the packing number of G is 2. As H has no isolated vertices, both U and V are minimal vertex covers of H . Thus, the vertices vU and vV have disjoint closed neighborhoods since N(vU ) = U and N(vV ) = V and hence, we get ρ(G) ≥ 2. Clearly, by construction, every edge of H is a dominating edge of G. Therefore, we get γt(G) = 2. Since ρ(G) ≤ γt(G), we obtain ρ(G) ≤ 2 and hence, ρ(G) = 2. Now let T be a minimal TDS of G other than the edges of H . Then T contains at most one endpoint of an edge of H because otherwise T contains a TDS, which contradicts that T is minimal. Therefore, V (H) \ T is a vertex cover of H and hence, it contains a minimal vertex cover S of H . By construction there exists a vertex vS with N(vS) = S. As S ⊆ V (H) \ T , we obtain N(vS) ∩ T = ∅, which contradicts that T is a TDS of G. Consequently, edges of H are the only minimal TDSs of G and hence, G is a WTD(2) graph and Gde = H . Lemma 3.4. Let G be a WTD(2) graph with ρ(G) = 2. Then, G is in W2. Proof. Let {x, y} be a packing with minimum |N [x]|+ |N [y]|. Note that every dominating edge of G has one endpoint from N(x) and one from N(y) and hence, Gde is a bipartite graph, say with parts X and Y where X ⊆ N(x) and Y ⊆ N(y). We next show that X = N(x) and Y = N(y). By symmetry, it suffices to prove X = N(x). Notice that Gde has no isolated vertices and therefore, X is a minimal vertex cover of Gde. By Proposition 3.2 there exists a vertex vX satisfying N(vX) = X . Suppose that X ̸= N(x). Then, we get X ⊂ N(x). Clearly vX ̸= y. Moreover, vX /∈ N(y) since y /∈ X = N(vX). Thus, we get N [vX ] ∩N [y] = ∅ and hence {vX , y} is a packing of G. However, we obtain |N [vX ]| + |N [y]| < |N [x]| + |N [y]| since X ⊂ N(x), which contradicts the definition of the packing {x, y}. Consequently, we get X = N(x) and hence, we may take vX = x. Similarly, we have Y = N(y) and we may assume vY = y. Now let S be a minimal vertex cover of Gde. By Proposition 3.2 there exists a vertex vS satisfying N(vS) = S. If S = X or S = Y , we can take vS to be x or y, respectively, and in both cases, we have vS /∈ V (Gde). Otherwise, suppose that vS ∈ V (Gde) = X ∪ Y . Without loss of generality, let vS ∈ X . Then, as X = N(x), we get x ∈ N(vS) = S ⊆ N(x) ∪N(y), which is a contradiction. Therefore, vS is not a vertex of Gde, that is, vS ∈ V (G) \ V (Gde). Finally, we see that one can obtain the graph G by following the procedure in the definition of W2 with the initial bipartite graph H = Gde. Combining the results in Lemma 3.3 and Lemma 3.4 gives the following structural characterization of WTD(2) graphs with ρ(G) = 2. Moreover, by definition of the class W2, this provides us with a procedure to construct any WTD(2) graph with ρ(G) = 2. Theorem 3.5. A graph G is WTD(2) with ρ(G) = 2 if and only if G ∈ W2. Given a graph G, the diameter of G, denoted by diam(G) is the maximum length of a shortest path between any pair of vertices of G. Let G be a graph such that γt(G) = 2. Then, it is easy to see that diam(G) ≤ 3. Moreover, whenever γt(G) = 2, we have diam(G) = 3 if and only if ρ(G) = 2 and therefore, in all the statements in Lemma 3.3, Lemma 3.4 and Theorem 3.5, the condition ρ(G) = 2 can be replaced with diam(G) = 3. 216 Ars Math. Contemp. 20 (2021) 209–222 Corollary 3.6. A graph G is WTD(2) with diam(G) = 3 if and only if G ∈ W2. One may attempt to modify the description of W2 graphs in order to describe all WTD(2) graphs with ρ(G) = 1. In the first step of the process of building a graph in W2, if one starts with a non-bipartite graph H with no isolated vertices, then the resulting graph is still WTD(2) but has packing number 1. However, not every WTD(2) graph G with ρ(G) = 1 can be obtained in this way. For example, consider the graph presented in Figure 1. To obtain this graph G, in Step 1 one should definitely choose H to be the graph with vertex set {z, y, t, w} and edge set {zy, yt, tw} which is indeed Gde. However, in Step 2 if one chooses a new vertex vS for S = {y, w} (which is a minimal vertex cover of Gde), then the graph G can not be obtained. So, the complete characterization of WTD(2) graphs with ρ(G) = 1 is left as an open question. 3.2 Triangle-free WTD(2) graphs In this subsection, we provide characterization of triangle-free WTD(2) graphs. Lemma 3.7. If G is a triangle-free graph with γt(G) = 2, then G is a bipartite graph and we have ρ(G) = { 1, if G is complete bipartite; 2, otherwise. Proof. Let uv be a dominating edge of G. Then we have N(u) ∪ N(v) = V (G). As G is triangle-free, none of two adjacent vertices have a common neighbor. Therefore, we have N(u) ∩ N(v) = ∅ and also see that both N(u) and N(v) are independent sets. We consequently obtain that G is a bipartite graph with parts N(u) and N(v). Since ρ(G) ≤ γt(G) = 2, we have ρ(G) ∈ {1, 2}. Moreover, it is clear that ρ(G) = 1 if and only if each vertex in N(u) is adjacent to all the vertices in N(v), i.e., G is a complete bipartite graph. For a bipartite graph with parts X and Y , define Xu = {x ∈ X : N(x) = Y } and Yu = {y ∈ Y : N(y) = X}. In other words, Xu (resp. Yu) is the set of vertices in X (resp. Y ) which are adjacent to every vertex in Y (resp. X). The following result characterizes all triangle-free WTD(2) graphs. Theorem 3.8. The following three statements are equivalent: (i) G is a triangle-free WTD(2) graph. (ii) G is a bipartite WTD(2) graph. (iii) G is complete bipartite graph or G is a bipartite graph with parts X and Y such that there exist vertices a ∈ X \Xu and b ∈ Y \ Yu satisfying N(a) = Yu ̸= ∅ and N(b) = Xu ̸= ∅. Proof. By Lemma 3.7 we see that (i) implies (ii). On the other hand, the implication (iii) ⇒ (i) can be easily verified and hence, the proof finishes if we show that (ii) implies (iii). Now let G be a bipartite WTD(2) graph, say with parts X and Y . Clearly we will only consider the case when G is not a complete bipartite graph. By definition of Xu and Yu, note that every dominating edge of G has one endpoint in Xu ̸= ∅ and one endpoint in Yu ̸= ∅. Moreover, any edge xy where x ∈ Xu and y ∈ Yu is a dominating edge of G. S. Bahadır et al.: Well-totally-dominated graphs 217 Therefore, Gde is the subgraph of G induced by Xu∪Yu and it is complete bipartite. Thus, Gde has only two minimal vertex covers, namely Xu and Yu. Then, definition of a graph in W2 and Theorem 3.5 imply the existence of the vertices a ∈ X \ Xu and b ∈ Y \ Yu with N(a) = Yu and N(b) = Xu. Although a polynomial time recognition algorithm for WTD(2) graphs follows from Theorem 2.8, the characterization in Theorem 3.8 provides us with a simple linear time recognition algorithm. Corollary 3.9. Triangle-free WTD(2) graphs can be recognized in linear time. Proof. Given a graph G, one can check whether it is a connected bipartite graph and if so, find its unique bipartition (X,Y ) in linear time (in the number of vertices and edges of G). Then, sets Xu and Yu can be identified simply by assigning every vertex x ∈ X such that d(x) = |Y | into Xu, and y ∈ Y such that d(y) = |X| into Yu. According to Theorem 3.8, G is triangle-free WTD(2) if and only if either Xu = X and Yu = Y (thus, G is complete bipartite), or the removal of Xu and Yu leaves at least one isolated vertex in each one of X and Y . Clearly, all these checks take only linear time. 3.3 Planar WTD(2) graphs In this subsection, we study planar WTD(2) graphs whose minimum degree is at least three and show that such graphs can have at most sixteen vertices. Throughout this section, we frequently use the fact that a graph obtained by an edge contraction of a planar graph is also planar. Recall also that a planar graph contains no K5 or K3,3. Observation 3.10. Let G be a WTD(2) graph. The vertex obtained by edge contraction of a dominating edge is a universal vertex in the new graph. Let ν(G) denote the matching number of a graph G. Lemma 3.11. Let G be a planar WTD(2) graph. If ν(Gde) ≥ 3, then |V (G)| ≤ 8. Proof. Suppose that ν(Gde) ≥ 3 and G has at least 9 vertices. Then, G has three inde- pendent dominating edges, say u1v1, u2v2 and u3v3, and three vertices other than u1, u2, u3, v1, v2, v3, say w1, w2 and w3. Now contract the edges u1v1, u2v2 and u3v3. In the resulting graph, new three vertices and w1, w2, w3 contain a K3,3, which contradicts the planarity. Lemma 3.12. If G is a WTD(2) graph with δ(G) ≥ 3, then ν(Gde) ≥ 2. Proof. Let G be a WTD(2) graph with δ(G) ≥ 3. It suffices to show that G has two independent dominating edges. Let xy be a dominating edge of G. Since the minimum degree is at least three, each vertex of G has at least one neighbor in V (G) \ {x, y}. There- fore, V (G) \ {x, y} is a TDS of G and hence, it contains a dominating edge ab since G is WTD(2). As the dominating edges xy and ab share no vertex, we get ν(Gde) ≥ 2. Combining the results in Lemmas 3.11 and 3.12 gives the following result. Proposition 3.13. If G is a planar WTD(2) graph with δ(G) ≥ 3, then ν(Gde) = 2 or |V (G)| ≤ 8. 218 Ars Math. Contemp. 20 (2021) 209–222 We next study planar WTD(2) graphs whose minimum degree is at least 3 and match- ing number is 2. Proposition 3.14. If G is a planar WTD(2) graph with δ(G) ≥ 3 and ν(Gde) = 2, then |V (G)| ≤ 16. Proof. Let ab and xy be two independent dominating edges of G and H = G−{a, b, x, y}. Let H1, . . . ,Hm be the connected components of H and order of Hi be hi for i = 1, . . . ,m. Note that it suffices to show that h1 + · · ·+ hm ≤ 12. We first prove that each Hi is a path or a singleton. Note that it suffices to show that maximum degree of H is at most 2 and H contains no cycle. Suppose that a vertex v of H has three neighbors, say v1, v2, v3, in H . Then contraction of the edges ab and xy gives rise to a K3,3 with parts {ab, xy, v} and {v1, v2, v3}, which is a contradiction. Therefore, every vertex in H has at most two neighbors in H . Suppose that H has a cycle, say v1, v2, . . . , vk. Contract the edge vkvk−1 and denote the new point by vk−1. Then contract the edge vk−1vk−2 and denote the new point by vk−2 and so on. Follow this process until we get a triangle v1, v2, v3. Then contracting the edges ab and xy yields a K5 with vertices ab, xy, v1, v2, v3, which is a contradiction. Thus, H has no cycle and hence, H is a disjoint union of paths and singletons. We next show that for every vertex u ∈ H we have |N(u) ∩ {a, b, x, y}| ≥ 3 or |(N(u) ∪N(v)) ∩ {a, b, x, y}| ≥ 3 for some neighbor v ∈ V (H) of u. Since both ab and xy are dominating edges, the intersection N(u) ∩ {a, b, x, y} has at least two elements: one from {a, b} and one from {x, y}. Consider the case when |N(u) ∩ {a, b, x, y}| = 2. Without loss of generality, let N(u) ∩ {a, b, x, y} = {a, x}. Since the minimum degree of G is at least 3, there is no vertex v ∈ G such that N(v) = {a, x}. Hence, by Proposition 3.2 the set {a, x} is not a vertex cover of Gde. Then, there exists an edge wv of Gde such that {w, v} ∩ {a, x} = ∅. Thus, as ν(Gde) = 2 and ab, xy ∈ Gde, we have wv = by or w ∈ {b, y} and v ∈ V (H). Recall that wv is a dominating edge in G and hence, u is adjacent to w or v. Therefore, the case wv = by is impossible and we see that v is adjacent to u. Consequently, we get |(N(u) ∪ N(v)) ∩ {a, b, x, y}| ≥ 3 since w ∈ {b, y} is a neighbor of v. Note that this result implies that if {u} is a component of H , then u has at least three neighbors among a, b, x, y; otherwise, contraction of the edge uv gives rise to a vertex adjacent to at least three of a, b, x, y. We then apply the following process for each i = 1, . . . ,m: If hi ≤ 3, contract the edges of Hi and obtain a singleton. If hi ≥ 4, let Hi be the path v1, v2, . . . , vk where k = hi. First, contract v1v2 and vk−1vk. Then contract the paths v3v4v5, v6v7v8, . . . and so on. Note that for every i we obtain at least 2 + ⌊(hi − 4)/3⌋ = ⌊(hi + 2)/3⌋ vertices adjacent to at least three of a, b, x, y. Therefore, each such vertex is adjacent to both a and b or adjacent to both x and y. Assume that the number of vertices having at least three neighbors among a, b, x, y in the resulting graph is more than 4. Then, by pigeonhole principle, there will be three distinct vertices u1, u2 and u3 each of which is adjacent, without loss of generality, to both a and b. Then, contraction of the edge xy gives a K3,3 with parts {a, b, xy} and {u1, u2, u3}, contradicting the planarity of G. Thus, there are at most 4 vertices having at least three neighbors among a, b, x, y once the contraction process is terminated, that is, ∑m i=1⌊(hi +2)/3⌋ ≤ 4. Since hi is an integer, the inequality hi/3 ≤ ⌊(hi + 2)/3⌋ holds, implying that ∑m i=1 hi/3 ≤ 4 which yields ∑m i=1 hi ≤ 12 as desired. S. Bahadır et al.: Well-totally-dominated graphs 219 Propositions 3.13 and 3.14 imply that, unlike the general case stated in Corollary 2.5, there is a finite number of planar WTD(2) graphs with δ(G) ≥ 3. Theorem 3.15. If G is a planar WTD(2) graph with δ(G) ≥ 3, then |V (G)| ≤ 16. In contrast, there is no upper bound on the number of vertices for planar WTD(2) graphs with minimum degree 1 or 2. For example, consider a star with arbitrarily many leaves and a graph with arbitrarily many triangles sharing a common edge, respectively. 4 Girth of WTD graphs In this section, we provide a relation between the minimum degree and the girth for WTD graphs. We show that if the minimum degree is more than two in a WTD graph, then the graph contains a cycle of length at most twelve. It is shown in [12] that if G is a WTD graph with δ(G) ≥ 2, then the girth of G, g(G), is at most 14. Theorem 4.1 ([12, Theorem 4.1]). Suppose G is a connected graph with no leaves such that G has girth at least fifteen. Then γt(G) < Γt(G). By following the idea in the proof of Theorem 4.1 in [12], one can find other relations between δ(G) and g(G) of a WTD graph G. Before presenting such extensions, we need the following useful lemma, which is also given in [12]: Lemma 4.2. Let G be a WTD graph, u1v1, . . . , umvm be a subset of the edges of G and A = ∪mi=1{ui, vi}. If the subgraph of G induced by A is disjoint union of m complete graphs of order 2 and G−N [A] has no isolated vertices, then G−N [A] is also WTD. Proof. Let S be a minimal TDS of G −N [A]. We claim that S ∪ A is a minimal TDS of G. It is easy to see that it is a TDS of G. Suppose that S ∪ A contains another TDS of G, say T . Then T ∩S is a TDS of G−N [A] and hence, since S is minimal we get T ∩S = S. Therefore, we obtain T = S ∪A′ where A′ ⊆ A. If A \A′ is nonempty, then without loss of generality we assume that u1 ∈ A \ A′. But then, v1 is not dominated by T , which is a contradiction. Therefore, we have A′ = A, which implies that T = S ∪ A, that is, S ∪ A is minimal. As every minimal TDS of G has the same size, |S|+2m is independent of S and hence, G−N [A] is a WTD graph as well. Theorem 4.3. If G is a WTD graph with δ(G) ≥ 3, then g(G) ≤ 12. Proof. Assume that G is a WTD graph with δ(G) ≥ 3 and g(G) ≥ 13. Let P = v1, v2, v3, v4, v5 be a path in G. For any vertex v in G, let dP (v) = min1≤i≤5 dist(v, vi). Define Nk to be the set of vertices v with dP (v) = k for k = 1, 2, . . . . First note that every vertex in Nk has a neighbor in Nk−1 for every k ≥ 2. Moreover, for k = 1, 2, 3, Nk is an independent set since otherwise we obtain a cycle of length at most 11. We will now show that for k = 1, . . . , 4, any vertex in Nk has at least one neighbor in Nk+1. Suppose that there exist k ≤ 4 and v ∈ Nk such that v is adjacent to no vertex in Nk+1. By definition, it is clear that v has no neighbor in Nl for any l ≥ k + 2. Therefore, all the neighbors of v are in ∪1≤i≤kNi. Thus, as v has at least three neighbors, there exist three paths from v to P such that one of them has length k and two of them have length at most k + 1. By a simple case analysis, considering the vertices of these paths on P gives that there exist a cycle of length at most 2k + 3 ≤ 11, which is a contradiction. 220 Ars Math. Contemp. 20 (2021) 209–222 Now, let N2 = {w1, . . . , wm}. For every i = 1, . . . ,m, choose a neighbor of wi in N3, say ui. Let A = ∪mi=1{wi, ui}. For any i ̸= j, wi is not adjacent to uj because otherwise we obtain a cycle of length at most 10. Therefore, the induced subgraph of G induced by A is a disjoint union of m complete graphs of order 2. Next, consider the graph H = G −N [A]. Note that N [A] consists of N1, N2, N3 and some vertices in N4. Therefore, P is a connected component of H . As any vertex in N4 has a neighbor in N5, no vertex v ∈ N4 ∩ V (H) is isolated in H . Clearly, no vertex in Nk with k ≥ 6 is isolated in H since it has a neighbor in Nk−1. Suppose to the contrary that a vertex v in N5 is isolated in H . Then v has no neighbor in N5 and N6, and thus, all its neighbors are in N4. Therefore, since there exist three paths from v to P , this yields a cycle of length at most 12, which is a contradiction. Consequently, H has no isolated vertices and we can apply Lemma 4.2 and conclude that H is a WTD graph. However, P is a component of H and hence, it should be WTD as well. Nevertheless, a path of length 4 is not a WTD graph (both {v1, v2, v4, v5} and {v2, v3, v4} are minimal TDSs of P ), which is a contradiction. 5 Conclusion In this work, we studied graphs whose all minimal total dominating sets have the same size. We say these graphs are well-totally-dominated. We proved that well-totally-dominated graphs with bounded total domination number can be recognized in polynomial time. We then analyzed well-totally-dominated graphs with total domination number two for the special cases of triangle-free graphs and planar graphs. Finally, we focused on the girth of well-totally-dominated graphs. In particular, we proved that a well-totally-dominated graph with minimum degree at least three has girth at most 12. We now conclude with several future research directions. Although we proved in this paper that the problem of recognizing well-totally-dominated graphs with bounded total domination number can be solved in polynomial time, the com- plexity of the general case is an open research problem. Hence, we pose the following question: Problem 5.1. What is the computational complexity of recognizing well-totally-dominated graphs? We have already characterized WTD(2) graphs with packing number ρ(G) = 2 in The- orem 3.5. Since WTD(2) graphs have ρ(G) ≤ 2, in order to complete the characterization of all WTD(2) graphs, it remains to answer the following question: Problem 5.2. What are WTD(2) graphs with ρ(G) = 1? Along the same line, one may consider to generalize our result in Theorem 3.5. It is well known that ρ(G) ≤ γt(G) ≤ Γt(G); hence graphs with ρ(G) = Γt(G) form a subclass of WTD graphs. This suggests our next open problem: Problem 5.3. What are WTD(k) graphs with ρ(G) = k? Lastly, we have shown in Theorem 3.15 that planar WTD(2) graphs with δ(G) ≥ 3 have at most 16 vertices. Our intuition is that 16 is not a tight bound. Thus, we pose the following question: Problem 5.4. Is the upper bound of 16 for the number of vertices of a planar WTD(2) graph with δ(G) ≥ 3 tight? Can we determine all (finitely many) planar WTD(2) graphs? S. Bahadır et al.: Well-totally-dominated graphs 221 ORCID iDs Selim Bahadır https://orcid.org/0000-0003-1533-7194 Tınaz Ekim https://orcid.org/0000-0002-1171-9294 Didem Gözüpek https://orcid.org/0000-0001-8450-1897 References [1] S. Bahadır and D. Gözüpek, On a class of graphs with large total domination number, Discrete Math. Theor. Comput. Sci. 20 (2018), #23, doi:10.23638/dmtcs-20-1-23. [2] E. Boros, V. Gurvich and P. L. Hammer, Dual subimplicants of positive boolean functions, Optim. Methods Softw. 10 (1998), 147–156, doi:10.1080/10556789808805708. [3] R. C. Brigham, J. R. Carrington and R. P. Vitray, Connected graphs with maximum total dom- ination number, J. Combin. Math. Combin. Comput. 34 (2000), 81–96. [4] E. J. Cockayne, R. M. Dawes and S. T. Hedetniemi, Total domination in graphs, Networks 10 (1980), 211–219, doi:10.1002/net.3230100304. [5] T. Eiter and G. Gottlob, Identifying the minimal transversals of a hypergraph and related prob- lems, SIAM J. Comput. 24 (1995), 1278–1304, doi:10.1137/s0097539793250299. [6] Q. Fang, On the computational complexity of upper total domination, Discrete Appl. Math. 136 (2004), 13–22, doi:10.1016/s0166-218x(03)00195-1. [7] A. Finbow, A. Frendrup and P. D. Vestergaard, Total well dominated trees, Research Report R- 2009-14, Department of Mathematical Sciences, Aalborg University, 2009, https://vbn. aau.dk/en/publications/total-well-dominated-trees. [8] A. Finbow, B. Hartnell and R. Nowakowski, Well-dominated graphs: a collection of well- covered ones, Ars Combin. 25 (1988), 5–10. [9] S. Finbow and C. M. van Bommel, Triangulations and equality in the domination chain, Dis- crete Appl. Math. 194 (2015), 81–92, doi:10.1016/j.dam.2015.05.025. [10] T. J. Gionet, E. L. C. King and Y. Sha, A revision and extension of results on 4-regular, 4- connected, claw-free graphs, Discrete Appl. Math. 159 (2011), 1225–1230, doi:10.1016/j.dam. 2011.04.013. [11] D. Gözüpek, A. Hujdurović and M. Milanič, Characterizations of minimal dominating sets and the well-dominated property in lexicographic product graphs, Discrete Math. Theor. Comput. Sci. 19 (2017), #25, doi:10.23638/dmtcs-19-1-25. [12] B. Hartnell and D. F. Rall, On graphs in which every minimal total dominating set is minimum, Congr. Numer. 123 (1997), 109–117, proceedings of the Twenty-eighth Southeastern Inter- national Conference on Combinatorics, Graph Theory and Computing (Boca Raton, Florida, 1997). [13] B. L. Hartnell, Well-covered graphs, J. Combin. Math. Combin. Comput. 29 (1999), 107–116. [14] T. W. Haynes, S. Hedetniemi and P. Slater, Fundamentals of Domination in Graphs, volume 208 of Monographs and Textbooks in Pure and Applied Mathematics, CRC press, Boca Raton, Florida, 1998, doi:10.1201/9781482246582. [15] 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. [16] X. Hou, Y. Lu and X. Xu, A characterization of (γt, 2γ)-block graphs, Util. Math. 82 (2010), 155–159. [17] V. E. Levit and D. Tankus, Well-dominated graphs without cycles of lengths 4 and 5, Discrete Math. 340 (2017), 1793–1801, doi:10.1016/j.disc.2017.02.021. 222 Ars Math. Contemp. 20 (2021) 209–222 [18] J. Pfaff, R. Laskar and S. T. Hedetniemi, Linear algorithms for independent domination and total domination in series-parallel graphs, Congr. Numer. 45 (1984), 71–82, proceedings of the Fifteenth Southeastern Conference on Combinatorics, Graph Theory and Computing (Baton Rouge, Louisiana, 1984). [19] M. D. Plummer, Well-covered graphs: a survey, Quaestiones Math. 16 (1993), 253–287, doi: 10.1080/16073606.1993.9631737. [20] J. Topp and L. Volkmann, Well covered and well dominated block graphs and unicyclic graphs, Math. Pannon. 1 (1990), 55–66, http://mathematica-pannonica.ttk.pte.hu/ articles/mp01-2/mp01-2-055-066.pdf. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 223–231 https://doi.org/10.26493/1855-3974.2173.71a (Also available at http://amc-journal.eu) Nordhaus-Gaddum type inequalities for the distinguishing index Monika Pilśniak AGH University, Department of Discrete Mathematics, al. Mickiewicza 30, 30-059 Krakow, Poland Received 6 November 2019, accepted 1 March 2021, published online 3 November 2021 Abstract The distinguishing index of a graph G, denoted by D′(G), is the least number of colours in an edge colouring of G not preserved by any nontrivial automorphism. This invariant is defined for any graph without K2 as a connected component and without two isolated vertices, and such a graph is called admissible. We prove the Nordhaus-Gaddum type relation: 2 ≤ D′(G) +D′(G) ≤ ∆(G) + 2 for every admissible connected graph G of order |G| ≥ 7 such that G is also admissible. Keywords: Symmetry breaking in graphs, distinguishing index, Nordhaus-Gaddum type bounds. Math. Subj. Class. (2020): 05C15, 05C25 1 Introduction and main result We consider finite graphs and their edge colourings, not necessarily proper. Such a colour- ing breaks an automorphism of a graph if there exists an edge that is mapped into an edge with a different colour by that automorphism. A colouring is called asymmetric (or dis- tinguishing), if it breaks all non-trivial automorphisms. The minimum number of colours in an asymmetric colouring of a graph G is called the distinguishing index of G and is denoted by D′(G). Obviously, the distinguishing index is defined only for graphs without K2 as a component and with at most one isolated vertex. We call such graphs admissible. The following general upper bound for the distinguishing index of connected graphs with respect to the maximum degree was proved by Kalinowski and Pilśniak in [8]. E-mail address: pilsniak@agh.edu.pl (Monika Pilśniak) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 224 Ars Math. Contemp. 20 (2021) 223–231 Theorem 1.1 ([8]). If G is a connected graph with at least three vertices, then D′(G) ≤ ∆(G) unless G is C3, C4 or C5. This result was improved in [10], where all connected graphs with the distinguishing index equal to the maximum degree were characterized. In particular, a tree is called sym- metric (respectively, bisymmetric) if it contains a central vertex v0 (resp. a central edge e0), all leaves are at the same distance from v0 (resp. e0) and all vertices that are not leaves have the same degree. Theorem 1.2 ([10]). Let G be a connected graph of order at least three. Then D′(G) ≤ ∆(G)− 1 unless G is a cycle, a symmetric or a bisymmetric tree, K4 or K3,3. In the same paper, the conjecture for 2-connected graphs was formulated, and quite recently was confirmed in [7] in a bit stronger form, as follows. Theorem 1.3 ([7]). If G is a connected graph with minimum degree δ(G) ≥ 2, then D′(G) ≤ ⌈√ ∆(G) ⌉ + 1. The main goal of the paper is a proof of a Nordhaus-Gaddum type inequalities for the distinguishing index of a graph. Our investigation was motivated by the renowned result of Nordhaus-Gaddum [9] who determined in 1956 lower and upper bounds for the sum of the chromatic numbers of a graph and its complement (actually, the upper bound was first proved by Zykov [12] in 1949). Since then, Nordhaus-Gaddum type bounds were obtained for many graph invariants. An exhaustive survey is given in [1]. In particular, it was considered by Collins and Trenk in [3] for the distinguishing chromatic number χD(G), which is the minimum number of colours in an asymmetric proper vertex colouring of a graph G. The Nordhaus-Gaddum type inequalities were also investigated in [10] for the chromatic distinguishing index χ′D(G) of a graph G defined for asymmetric proper edge colourings. It was proved therein that if G is an admissible graph of order n ≥ 3 distinct from K1,4, then n− 1 ≤ χ′D(G) + χ′D(G) ≤ 2(n− 1). Both upper and lower bounds are similar to Vizing bounds proved for the chromatic index of a graph [11] but in the proof for the chromatic distinguishing index we have to be more careful. It was also conjectured in [10] that 2 ≤ D′(G) +D′(G) ≤ max{∆(G),∆(G)}+ 2 if both graphs G and G are admissible and of order n ≥ 7. It was confirmed for some classes of graphs, in particular for trees, claw-free graphs, 3-connected graphs and traceable graphs. Here, we prove the stronger version of Nordhaus-Gaddum type inequality for the distinguishing index. M. Pilśniak: Nordhaus-Gaddum type inequalities for the distinguishing index 225 Theorem 1.4 (Main Theorem). If both G and G are admissible graphs of order n ≥ 7, and G is connected, then 2 ≤ D′(G) +D′(G) ≤ ∆(G) + 2. The lower bound by 2 is obvious. Indeed, if a graph G is asymmetric, that means it has a trivial automorphism group, then the distinguishing index of both G and G equals 1. Moreover, Theorem 1.4 is tight. To see this, consider a symmetric (or bisymmetric) tree T of order n ≥ 7. Then by [5] T contains an asymmetric spanning tree if T is different from a star, so D′(T ) = 2 by Proposition 2.2. So, it follows from Theorem 1.2 that D′(T ) +D′(T ) = ∆(T ) + 2 for symmetric and bisymmetric trees. It has to be noted that there exist graphs of order less than 7 violating the right inequal- ity in Theorem 1.4. For example, D′(K3,3) = 3, D′(K3,3) = 4, whence D′(K3,3) + D′(K3,3) = ∆(K3,3) + 4. Also, D′(C5) + D′(C5) = ∆(C5) + 4, and D′(K1,i) + D′(K1,i) = ∆(K1,i) + 3 for i = 3, 4, 5. In Section 2 we recall some results useful in the proof of Theorem 1.4, which is given in Section 3. 2 Known bounds for D′ Let us recall some useful results, before we start to prove the Main Theorem. A graph that contains a Hamiltonian path, i.e. a path that visits each vertex of the graph, is called traceable. Following [2], we define the k-th Bondy-Chvátal closure clk(G) of a graph G as the graph obtained from G by recursively joining pairs of non-adjacent vertices with degree-sum at least k. By the well-known theorem of Bondy and Chvátal [2], a graph G of order n is traceable whenever cln−1(G) is traceable. We begin with the distinguishing index of complete graphs and of traceable graphs. Proposition 2.1 ([8]). D′(Kn) = { 2, if n ≥ 6, 3, if n = 3, 4, 5. The following simple observation is very useful in some proofs. A subgraph H of a graph G is called almost spanning if H is a spanning subgraph of a graph G− v for some v ∈ V (G). Proposition 2.2 ([10]). If H is a spanning or almost spanning subgraph with at least three vertices of a graph G, then D′(G) ≤ D′(H) + 1. In particular, the spanning path in a traceable graph needs two colours to break its non- trivial automorphism, so a traceable graph has the distinguishing index at most 3. Actually we have a stronger result. Theorem 2.3 ([10]). Let G be a traceable graph of order at least 3. If G has at least 7 vertices, then D′(G) ≤ 2. Moreover, if G is of order at most 6, then D′(G) ≤ 3. The distinguishing index of complete bipartite graphs was determined independently by Fisher and Isaak [4] and by Imrich, Jerebic and Klavžar [6]. Actually, they formulated their result for the distinguishing number D(Kp □Kq) of the Cartesian product of complete graphs, but D′(Kp,q) = D(Kp □Kq). 226 Ars Math. Contemp. 20 (2021) 223–231 Theorem 2.4 ([4, 6]). Let p, q, d be integers such that d ≥ 2 and (d− 1)p < q ≤ dp. Then D′(Kp,q) = { d, if q ≤ dp − ⌈logd p⌉ − 1, d+ 1, if q ≥ dp − ⌈logd p⌉+ 1. If q = dp − ⌈logd p⌉, then the distinguishing index D′(Kp,q) is either d or d + 1 and can be computed recursively in O(log∗(q)) time. In the rest of the paper, we make use of the following immediate corollary. Corollary 2.5. If p ≤ q, then D′(Kp,q) ≤ ⌈ p √ q⌉+ 1. The following simple observation is used later in this section. Proposition 2.6. If H is an admissible disconnected graph of order at least 7, then D′(H) ≤ |H| − 2. Proof. Theorem 1.2 implies that the only connected graph H with D′(H) ≥ |H| − 1 is K3,K4, C4 or a star K1,n−1 for n ≥ 3. If all components H1, . . . ,Hs of H are pairwise non-isomorphic, then D′(H) = max{D′(Hi) : i = 1, . . . , s}, so D′(H) ≤ |H| − 2. If H contains t ≥ 2 copies of a graph H1 as its components, so |H| ≥ t|H1|, then we colour one of them distinguishingly and use one extra colour for each other copy. Hence, D′(tH1) ≤ D′(H1) + t− 1 ≤ |H1|+ t− 1 ≤ |H| t + t− 1 ≤ |H| − 2. We easily extend a result of [10] for trees to forests. First, recall a result of Hedetniemi et al. [5] on packing two trees into Kn. Theorem 2.7 ([5]). A complete graph Kn contains edge disjoint copies of any two trees of order n distinct from a star K1,n−1. Proposition 2.8. Let G be an admissible graph of order n such that G is an admissible forest. Then D′(G) ≤ 2 if n ≥ 7, and D′(G) ≤ 3 otherwise. Proof. The case when G is a tree was proved in [10]. Otherwise, it easily follows from Theorem 2.7 that G contains a Hamiltonian path Pn. Indeed, we can consider a tree F ′ spanned by G, and every tree distinct from a star is included in a subgraph F ′ of G. Thus D′(G) ≤ 2 if n ≥ 7, and D′(G) ≤ 3 if 3 ≤ n ≤ 6 by Theorem 2.3. Additionally, let us note the following observation for small graphs. Denote by W1 and W2 the two graphs from Figure 1 called windmills. W1 b b bb b b b W2 b b bb b b b Figure 1: Two windmills. M. Pilśniak: Nordhaus-Gaddum type inequalities for the distinguishing index 227 Proposition 2.9. If G is a connected graph of order at most 7 different from windmills W1 and W2, and from a star K1,n for n = 4, 5, 6, then D′(G) ≤ 3. Proof. Observe that if ∆(G) ≤ 4, then the claim holds by Theorem 1.2. So let ∆(G) ≥ 5. First assume that G does not have pendant edges. If the longest cycle in G is of order at least 6, then G is traceable and D′(G) ≤ 3 by Theorem 2.3. If the longest cycle C in G is of order 5, then we colour the edges of this cycle with 0, 1, 0, 2, 0 (in this order) and all chords with 1. Thus, each vertex of C has an incident edge coloured with 0. If there exists a vertex of G with noncoloured incident edges, then all of them we colour with 2 and all the remaining edges with 1. It is an asymmetric colouring of G, because outside the initial cycle C we create a monochromatic vertex with colour 2 and a monochromatic vertex with colour 1 or a bichromatic one with colours 1 and 2. If the longest cycle in G is of order 4, then a 2-connected G is isomorphic to K2,r or K2,r + e with r ≤ 5 where e is an edge between two vertices of maximum degree. Three colour suffice for an asymmetric colouring of r paths of length two between the two vertices of maximum degree. If such a 2-connected graph is joined with a triangle (in a cut vertex), then we can colour every edge of the triangle with a different colour. If two cycle of length 4 (with chords) meet in only one vertex, then an asymmetric colouring of one of them uses colour 1 twice, while the other one uses 2 twice (apart from one edge coloured with 0). If the longest cycle is a triangle C3, then G has a cut vertex v where three cycles meet. Then a pair of edges including v in every triangle is coloured with 1 and 2, while every remaining edge obtains a distinct colour 1, 2 or 3. Now assume that there exists a leaf v in G. First suppose there exists an induced subgraph B of G with minimal degree at least 2. Then we can colour it with three colours as above. So we have to distinguish now only pendant trees (in particular paths and edges) with a common vertex just fixed in B. It is always possible to do this with three colours if |G| ≤ 7 unless B has a pendant star K1,4. Then we obtain an exceptional graph W1. Finally, let G be a tree of order at most 7 different from a star K1,n with n = 3, . . . , 6. Then D′(G) ≤ ∆(G) − 1 by Theorem 1.2. Clearly, D′(G) ≤ 3 unless G = W2, which needs four colours like W1. 3 Proof of the Main Theorem Proof of Theorem 1.4. Let G be a connected graph of order n ≥ 7 such that both G and its complement G are admissible. Denote ∆ = ∆(G). Clearly, ∆ ≥ ∆(G) + 1 whenever G is disconnected. Next, if ∆ ≤ n−12 , then δ(G) = n− 1−∆ ≥ n− 1 2 , and G is connected. Otherwise, n−12 ≥ ∆ ≥ ∆(G) + 1 ≥ n−1 2 + 1. Hence, G satisfies the well-known Dirac’s condition for traceability, so G is traceable. Hence, D′(G) + D′(G) ≤ ∆+ 2 by Theorem 1.1 and Theorem 2.3. Assume then that ∆ ≥ ⌈n2 ⌉. Analogously, we can assume that ∆(G) ≥ ⌈ n 2 ⌉. Indeed, our theorem holds if ∆(G) ≤ n−12 . Then D ′(G) ≤ 2 and D′(G) ≤ ∆(G) ≤ n−12 ≤ ∆ if G is connected, or D′(G) ≤ ⌈n2 ⌉ ≤ ∆ if G is disconnected, by the following reasons. Theorem 1.1 implies that the only connected graphs H with D′(H) > ∆(H) are C3, C4 and C5. If all components H1, . . . ,Hs of G are pairwise non-isomorphic, then 228 Ars Math. Contemp. 20 (2021) 223–231 D′(G) = max{D′(Hi) : i = 1, . . . , s} ≤ max{3,∆(Hi)}, so D′(G) ≤ n2 . If G contains ti ≥ 2 copies of a graph Hi as its components, then let t = max{ti : i = 1, . . . , s} and let H = Hl where D′(Hl) = max{D′(Hi) : i = 1, . . . , s}, and we colour one copy of Hl distinguishingly and use one extra colour for each other copy. So, for H different from C3, C4 and C5 we have D′(G) ≤ D′(tHl) ≤ D′(Hl) + t− 1 ≤ ∆(Hl) + t− 1 ≤ |G| t + t− 2 ≤ n 2 , and for i ∈ {3, 4, 5} we obtain D′(tCi) ≤ t+ 2 ≤ ⌈ ti2 ⌉. We distinguish two cases. Case A: Both G and G are connected graphs without pendant edges. Then D′(G) ≤ ⌈ √ ∆⌉+1 and D′(G) ≤ ⌈ √ ∆(G)⌉+1. Hence, the inequality D′(G)+ D′(G) ≤ ∆+ 2 is weaker than ∆− ⌈√ ∆ ⌉ ≥ ⌈√ ∆(G) ⌉ . (3.1) First assume that ∆ ≥ ∆(G). It is easy to see that the inequality (3.1) is satisfied unless ∆ = ∆(G) = 5. In the latter case 8 ≤ n ≤ 10 and δ(G) = δ(G) = n − 6. We want to show that either G or its complement G is traceable. We say that a sequence (ai) is minorized by a sequence (bi) if bi ≤ ai for any i. If n = 8, then the degree sequence, ordered non-increasingly, of G (or G) is minorized by (5, 5, 4, 4, 2, 2, 2, 2) or by (5, 4, 4, 4, 3, 2, 2, 2). Indeed, we know by assumptions that b1 = 5, b8 = 2, two terms of bi have to be odd since the sum of degrees is even in every graph, and the sum of the fourth term of the sequence of G and the fifth term of the sequence of G equals n − 1 = 7, so one of them cannot be smaller than ⌈n−12 ⌉ = 4. Now, by definition of the (n − 1)-th Bondy-Chvátal closure cln−1(G), a vertex of degree five in G has degree n − 1 = 7 in cl7(G), so we have to add two new edges incident to it. Observe that adding new edges yields another vertex that has degree n − 1 in cl7(G), and this is the case at each step of creating cl7(G). Finally, cl7(G) = K8. Hence, G is traceable, by the Bondy-Chvátal theorem [2]. Similarly, if n = 9, we may assume that the degree sequence of G is minorized by (5, 4, 4, 4, 4, 4, 3, 3, 3) or by (5, 5, 4, 4, 4, 3, 3, 3, 3), and it is not difficult to see that cl8(G) = K9. For n = 10, the degree sequence of G is minorized by (5, 5, 4, . . . , 4) and here it is clear that cl9(G) = K10. For brevity, the details are left to the reader. Now assume that ∆(G) > ∆. Then it is easily seen that the inequality (3.1) holds for any ∆, ∆(G) and n unless either n = 8,∆ = 4,∆(G) = 5, or n = 9,∆ = 5,∆(G) = 6, or n = 10,∆ = 5,∆(G) ∈ {6, 7}, since ⌈n2 ⌉−⌈ √ ⌈n2 ⌉⌉ ≥ ⌈ √ n− 3⌉. The same argument as above confirms that G is traceable. Case B: A graph G is disconnected or δ(G) = 1. If G is a forest, then the conclusion follows from Proposition 2.8. Hence, assume that G contains a 2-connected block. We now consider decompositions of G into two subgraphs F1, F2 such that E(F1) ∪ E(F2) = E(G) and the vertex sets V (F1), V (F2) share at most one vertex which is a cut-vertex of G. Denote p = |F1|−1, q = |F2|−1 if F1 and F2 share a common vertex, and p = |F1|, q = |F2| if F1 and F2 are disjoint. Assume that p ≤ q and the difference q − p is smallest possible. Observe that ∆(G) ≥ q and G is spanned or almost spanned by a complete bipartite graph Kp,q . M. Pilśniak: Nordhaus-Gaddum type inequalities for the distinguishing index 229 First, suppose that q ≤ 2p − p. Then D′(G) ≤ 2, since G is (almost) spanned by an asymmetric spanning subgraph of Kp,q for p + q ≥ 7 by Proposition 3.10 in [10] (see also [6]). If G is connected, then ∆ = ∆(G) = n − 2 since δ(G) = 1 by the assumption of Case B. Moreover, D′(G) ≤ n− 2 by Theorem 1.1 as ∆(G) ≤ n− 2. Hence D′(G) +D′(G) ≤ 2 + n− 2 = ∆+ 2. If G is disconnected, then either δ(G) = 1 or δ(G) ≥ 2. If δ(G) = 1, then D′(G) ≤ n− 2 by Proposition 2.6 and D′(G) +D′(G) ≤ 2 + n− 2 = ∆+ 2. Now, let δ(G) ≥ 2, and assume (in the worst case) that G contains s components isomorphic to a connected graph H . Recall that ∆(G) ≥ n2 , as we assumed at the beginning of the proof. So, 3 ≤ |H| < n2s . If we use a distinct additional colour for every component H , then D ′(G) ≤ ⌈ √ ∆(G)⌉+ s ≤ ⌈ √ n− 2⌉+ s, by Theorem 1.3. So, we would like to show that⌈√ n− 2 ⌉ + s+ 2 ≤ n− |H|+ 2, (3.2) since ∆(G) ≥ n− |H|. To confirm the inequality (3.2), we estimate⌈√ n− 2 ⌉ + s ≤ ⌈√ n− 2 ⌉ + n 6 ≤ n− n 4 ≤ n− |H|. It is easy to verify that the second inequality is always true. The last inequality is obvious if s is at least 2. If s = 1 we do not need to distinguishing connected component one from another, hence we use the same colours in every component and D′(G) ≤ ⌈ √ ∆(G)⌉+1 ≤ ∆. So, this subcase is completed. Now, assume that q ≥ 2p − p+ 1. Then D′(G) ≤ p√q + 2 for p ≥ 2 by Corollary 2.5 and Proposition 2.2. In this case the graph G can be obtained from a 2-connected graph B by attaching a number of independent pendant subgraphs (of order at most p) to it or there is a component of order p and a 2-connected component B of order q. Hence, ∆(B) ≤ q and D′(G) ≤ ⌈√q⌉+ 1, by Theorem 1.3 for the block B, and by the observation that then every subgraph attached to B has one vertex fixed and the order at most p ≤ √q + 2. We obtain D′(G) + D′(G) ≤ ⌈ p√q⌉ + 2 + ⌈√q⌉ + 1. Recall that ∆ ≥ q, so it is enough to check whether ⌈ p√q⌉+ ⌈√q⌉+ 3 ≤ q + 2. Consequently, we obtain the inequality q − ⌈ p√q⌉ − ⌈√q⌉ − 1 ≥ 0. (3.3) For p = 3, we have q ≥ 23 − 3 + 1 = 6. For q = 6, p = 3 we have D′(G) ≤ 4 by Proposition 2.2, and D′(G) ≤ 3 by Proposition 2.9 for F2. Observe now that the inequality (3.3) is satisfied, because it holds for q = 7, and q − ⌈ 3√q⌉ − ⌈√q⌉ − 1 is an increasing function of q. The inequality (3.3) also holds for larger values of p since its left-hand side is non-decreasing with respect to p. If p = 2, then inequality (3.3) is satisfied for q ≥ 7. For q = 5 or q = 6, G is (almost) spanned by K2,5 or K2,6, so D′(G) ≤ 4 by Proposition 2.2, and D′(G) ≤ 3 by Proposition 2.9 for F2. For q = 4 we have n = 7 and our theorem is true by Proposition 2.9. Now let p = 1. If G is disconnected, then G contains a 2-connected block B of order n − 1 and an isolated vertex v. Hence, ∆ = n − 1, of course. Then D′(G) ≤ D′(B) ≤ 230 Ars Math. Contemp. 20 (2021) 223–231 ⌈ √ ∆(G)⌉ + 1. Moreover, D′(G) ≤ ⌈ √ ∆⌉ + 1 whenever δ(G) ≥ 2, by Theorem 1.3. Hence, D′(G) +D′(G) ≤ 2 ⌈√ ∆ ⌉ + 2, and 2⌈ √ ∆⌉+ 2 ≤ ∆+ 2 for ∆ ≥ 6, since ∆ = n− 1 ≥ 6. If the graph G is connected, then G can be obtained from a 2-connected block B by attaching a number (maybe zero) of independent pendant edges to it. It is enough to break all nontrivial automorphisms of B. Then D′(G) ≤ D′(B) ≤ ⌈ √ ∆(G)⌉ + 1. Moreover, ∆ = n−2 and D′(G) ≤ ⌈ √ ∆⌉+1 whenever δ(G) ≥ 2, by Theorem 1.3. Then 2⌈ √ ∆⌉+ 2 ≤ ∆+2 for ∆ ≥ 6, unless ∆ = 5. But in the latter case n = 7 and D′(G)+D′(G) ≤ 6 by Proposition 2.9. Finally, assume that p = 1 and δ(G) = 1. Then we consider decompositions of G into two subgraphs F ′1, F ′ 2 such that E(F ′ 1)∪E(F ′2) = E(G) and the vertex sets V (F ′1), V (F ′2) share one vertex which is a cut-vertex of G. Denote p′ = |F ′1| − 1, q′ = |F ′2| − 1. Assume that p′ ≤ q′ and the difference q′ − p′ is smallest possible. Recall that ∆(G) = n− 2 and G is spanned or almost spanned by a complete bipartite graph Kp′,q′ . If q′ ≤ 2p′ − p′, then D′(G) ≤ 2 like above (for p′ + q′ ≥ 7), and D′(G) ≤ ∆ by Theorem 1.1. So, we are done. If q′ ≥ 2p′ − p′ + 1, then D′(G) ≤ p′ √ q′ + 2 for p′ ≥ 2, and we obtain D′(G) +D′(G) ≤ ⌈√ q′ ⌉ + ⌈ p′ √ q′ ⌉ + 3 ≤ p′ + q′ + 1, since ∆ ≥ n − 2 = p′ + q′ − 1. For p′ = 2, we have q′ ≥ 22 − 2 + 1 = 3, and the inequality 2⌈ √ q′⌉ ≤ q′ is satisfied since n ≥ 7. Hence q′ ≥ 4. The inequality p′ + q′ − ⌈ √ q′⌉ − ⌈ p′ √ q′⌉ − 2 ≥ 0 also holds for larger values of p′ since its left-hand side is non-decreasing with respect to p′. Let p′ = 1. Then there exists a 2-connected block B′ in G with a number of indepen- dent pendant edges attached to it (at least one). Hence D′(G) ≤ D′(B′) ≤ ⌈ √ ∆⌉ + 1. Recall that also D′(G) ≤ ⌈ √ ∆(G)⌉+ 1 ≤ ⌈ √ ∆⌉+ 1, since ∆ = ∆(G) = n− 2. So we verify the following inequality 2 ⌈√ ∆ ⌉ + 2 ≤ ∆+ 2 for ∆ ∈ {n − 2, n − 1}, which is true for n ≥ 7 unless ∆ = 5. But then n = 7 and D′(G) +D′(G) ≤ 6 once more by Proposition 2.9. ORCID iD Monika Pilśniak https://orcid.org/0000-0002-3734-7230 References [1] M. Aouchiche and P. Hansen, A survey of Nordhaus-Gaddum type relations, Discrete Appl. Math. 161 (2013), 466–546, doi:10.1016/j.dam.2011.12.018. [2] J. A. Bondy and V. Chvátal, A method in graph theory, Discrete Math. 15 (1976), 111–135, doi:10.1016/0012-365x(76)90078-9. [3] K. L. Collins and A. N. Trenk, Nordhaus-Gaddum theorem for the distinguishing chromatic number, Electron. J. Combin. 20 (2013), #P46 (18 pages), doi:10.37236/2117. M. Pilśniak: Nordhaus-Gaddum type inequalities for the distinguishing index 231 [4] M. J. Fisher and G. Isaak, Distinguishing colorings of Cartesian products of complete graphs, Discrete Math. 308 (2008), 2240–2246, doi:10.1016/j.disc.2007.04.070. [5] S. M. Hedetniemi, S. T. Hedetniemi and P. J. Slater, A note on packing two trees into KN , Ars Combin. 11 (1981), 149–153. [6] W. Imrich, J. Jerebic and S. Klavžar, The distinguishing number of Cartesian products of com- plete graphs, European J. Combin. 29 (2008), 922–929, doi:10.1016/j.ejc.2007.11.018. [7] W. Imrich, R. Kalinowski, M. Pilśniak and M. Woźniak, The distinguishing index of con- nected graphs without pendant edges, Ars Math. Contemp. 18 (2020), 117–126, doi:10.26493/ 1855-3974.1852.4f7. [8] R. Kalinowski and M. Pilśniak, Distinguishing graphs by edge-colourings, European J. Com- bin. 45 (2015), 124–131, doi:10.1016/j.ejc.2014.11.003. [9] E. A. Nordhaus and J. W. Gaddum, On complementary graphs, Amer. Math. Monthly 63 (1956), 175–177, doi:10.2307/2306658. [10] M. Pilśniak, Improving upper bounds for the distinguishing index, Ars Math. Contemp. 13 (2017), 259–274, doi:10.26493/1855-3974.981.ff0. [11] V. G. Vizing, The chromatic class of multigraphs, Kibernetika (Kiev) 1 (1965), 29–39. [12] A. A. Zykov, On some properties of linear complexes, Mat. Sb. (N.S.) 24 (1949), 163–188, http://mi.mathnet.ru/eng/msb5974. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 233–241 https://doi.org/10.26493/1855-3974.2284.aeb (Also available at http://amc-journal.eu) Closed formulas for the total Roman domination number of lexicographic product graphs Abel Cabrera Martı́nez , Juan Alberto Rodrı́guez-Velázquez Universitat Rovira i Virgili, Departament d’Enginyeria Informàtica i Matemàtiques, Av. Paı̈sos Catalans 26, 43007 Tarragona, Spain Received 19 March 2020, accepted 6 January 2021, published online 6 November 2021 Abstract Let G be a graph with no isolated vertex and f : V (G) → {0, 1, 2} a function. Let Vi = {x ∈ V (G) : f(x) = i} for every i ∈ {0, 1, 2}. We say that f is a total Roman dominating function on G if every vertex in V0 is adjacent to at least one vertex in V2 and the subgraph induced by V1 ∪ V2 has no isolated vertex. The weight of f is ω(f) =∑ v∈V (G) f(v). The minimum weight among all total Roman dominating functions onG is the total Roman domination number of G, denoted by γtR(G). It is known that the general problem of computing γtR(G) is NP-hard. In this paper, we show that if G is a graph with no isolated vertex and H is a nontrivial graph, then the total Roman domination number of the lexicographic product graph G ◦H is given by γtR(G ◦H) = { 2γt(G) if γ(H) ≥ 2, ξ(G) if γ(H) = 1, where γ(H) is the domination number of H , γt(G) is the total domination number of G and ξ(G) is a domination parameter defined on G. Keywords: Total Roman domination, total domination, lexicographic product graph. Math. Subj. Class. (2020): 05C69, 05C76 1 Introduction Let G be a graph with no isolated vertex and f : V (G) → {0, 1, 2} a function. Let Vi = {x ∈ V (G) : f(x) = i} for every i ∈ {0, 1, 2}. We will identify f with the partition of E-mail addresses: abel.cabrera@urv.cat (Abel Cabrera Martı́nez), juanalberto.rodriguez@urv.cat (Juan Alberto Rodrı́guez-Velázquez) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 234 Ars Math. Contemp. 20 (2021) 233–241 V (G) induced by f and write f(V0, V1, V2). The weight of f is defined to be ω(f) = f(V (G)) = ∑ v∈V (G) f(v) = |V1|+ 2|V2|. A function f(V0, V1, V2) is said to be total Roman dominating function onG if every vertex in V0 is adjacent to at least one vertex in V2 and the subgraph induced by V1 ∪ V2 has no isolated vertex [17]. The minimum weight among all total Roman dominating functions on G is the total Roman domination number of G, denoted by γtR(G). In this article, we continue the study initiated in [5] on the total Roman domination number of lexicographic product graphs. In particular, we give closed formulas for the total Roman domination number of lexicographic product graphs. Let G and H be two graphs. The lexicographic product of G and H is the graph G ◦H whose vertex set is V (G ◦H) = V (G)× V (H) and (u, v)(x, y) ∈ E(G ◦H) if and only if ux ∈ E(G) or u = x and vy ∈ E(H). Notice that for any u ∈ V (G) the subgraph of G ◦ H induced by {u} × V (H) is isomorphic to H . For simplicity, we will denote this subgraph by Hu. For a basic introduction to the lexicographic product of two graphs we suggest the books [7, 12]. One of the main problems in the study of G ◦ H consists of finding exact values or tight bounds for specific parameters of these graphs and express them in terms of known invariants of G and H . In particular, we cite the following works on domination theory of lexicographic product graphs: (total) domination [14, 18, 19, 21], Roman domi- nation [14], weak Roman domination [20], rainbow domination [15], super domination [6], doubly connected domination [2], secure domination [13], double domination [3] and total Roman domination [5]. We assume that the reader is familiar with the basic concepts and terminology of domi- nation in graph. If this is not the case, we suggest the textbooks [8, 9, 11]. In particular, we use the standard notation γ(G) and γt(G) for the domination number and the total domi- nation number of a graph G, respectively. Throughout the paper, N(v) will denote the set of neighbours or open neighbourhood of v in G. The closed neighbourhood of v, denoted by N [v], equals N(v) ∪ {v}. A vertex v ∈ V (G) such that N [v] = V (G) is said to be a universal vertex. For the remainder of the paper, definitions will be introduced whenever a concept is needed. 2 The case where γ(H) ≥ 2 The next theorem merges two results obtained in [14] and [21]. Theorem 2.1 ([14] and [21]). For any graph G with no isolated vertex and any nontrivial graph H , γ(G ◦H) = { γ(G), if γ(H) = 1, γt(G), if γ(H) ≥ 2. Below we present two theorems that complete the tools we need to deduce our first result. Theorem 2.2 ([1]). For any graph G with no isolated vertex, 2γ(G) ≤ γtR(G) ≤ min{2γt(G), 3γ(G)}. A. Cabrera Martı́nez and J. A. Rodrı́guez-Velázquez: Closed formulas for the total Roman . . . 235 Theorem 2.3 ([4]). For any graph G with no isolated vertex and any nontrivial graph H , γt(G ◦H) = γt(G). From the results above we deduce the following main theorem. Theorem 2.4. For any graph G with no isolated vertex and any graph H with γ(H) ≥ 2, γtR(G ◦H) = 2γt(G). Proof. The result immediately follows by applying Theorems 2.1, 2.3 and 2.2, in this order, i.e., 2γt(G) = 2γ(G ◦H) ≤ γtR(G ◦H) ≤ 2γt(G ◦H) = 2γt(G). Notice that, since the general optimization problem of finding the total domination number of a graph is NP-hard [16], by Theorem 2.4 we can conclude that the problem of finding the total Roman domination number is NP-hard. Even so, we would like to point out that there are several families of graphs for which the total domination number can be found in polynomial time [10]. 3 The case where γ(H) = 1 The following two lemmas are the main tools in this section. Lemma 3.1. Let G be a graph with no isolated vertex. For any nontrivial graph H with γ(H) = 1, there exists a γtR(G ◦H)-function f satisfying the following two conditions. (i) f(V (Hu)) ≤ 2 for every u ∈ V (G). (ii) If f(V (Hu)) = 2, then f(u, v) = 2 for some universal vertex v of H . Proof. Given a TRDF f on G ◦H , we define the set Rf = {x ∈ V (G) : f(V (Hx)) ≥ 3}. Let f be a γtR(G◦H)-function such that |Rf | is minimum among all γtR(G◦H)-functions. Let v ∈ V (H) be a universal vertex and suppose that there exists u ∈ Rf . We differentiate the following two cases. Case 1. There exists u′ ∈ N(u) such that f(V (Hu′)) ≥ 1. Let f ′ be the function defined by f ′(V (Hu)) = f ′(u, v) = 2 and f ′(x, y) = f(x, y) for every x ∈ V (G) \ {u}. It is readily seen that f ′ is a γtR(G ◦H)-function with |Rf ′ | < |Rf |, which is a contradiction. Case 2. f(N(u) × V (H)) = 0. In this case, we choose a vertex u′ ∈ N(u) and define a function f ′ as f ′(V (Hu′)) = f ′(u′, v) = 1, f ′(V (Hu)) = f ′(u, v) = 2 and f ′(x, y) = f(x, y) for every x ∈ V (G) \ {u, u′}. As in Case 1, f ′ is a γtR(G ◦ H)-function with |Rf ′ | < |Rf |, which is a contradiction. According to the two cases above, (i) follows. Now, for any γtR(G ◦ H)-function f(V0, V1, V2) satisfying (i), we define R′f = {x ∈ V (G) : f(V (Hx)) = 2 and V (Hx) ∩ V2 = ∅}. Let g(V ′0 , V ′1 , V ′2) be a γtR(G◦H)-function such that |R′g| is minimum among all γtR(G ◦H)-functions satisfying (i). Suppose that there exists u ∈ R′g . If there exists u′ ∈ N(u) such that, g(V (Hu′)) = 2, then the function g′ defined by g′(V (Hu)) = g′(u, v) = 1 and g′(x, y) = g(x, y) for every x ∈ V (G) \ {u}, is a TRDF on G ◦ H of weight ω(g′) < ω(g) = γtR(G ◦H), which is a contradiction. Hence, g(N(u)× V (H)) ≤ 1 and we can differentiate the following two cases. 236 Ars Math. Contemp. 20 (2021) 233–241 Case 1′. There exists u′ ∈ N(u) such that g(V (Hu′)) = 1. In this case, we define a function g′ by g′(V (Hu)) = g′(u, v) = 2 and g′(x, y) = g(x, y) for every x ∈ V (G) \ {u}. Notice that g′ is a γtR(G ◦ H)-function satisfying (i) and |R′g′ | < |R′g|, which is a contradiction. Case 2′. g(N(u) × V (H)) = 0. We fix u′ ∈ N(u). Notice that there exists u′′ ∈ N(u′)\{u}, with V (Hu′′)∩V ′2 ̸= ∅. Hence, we can define a function g′ as g′(V (Hu′)) = g′(u′, v) = g′(V (Hu)) = g ′(u, v) = 1 and g′(x, y) = g(x, y) for every x ∈ V (G) \ {u, u′}. As in Case 1′, g′ is a γtR(G ◦H)-function satisfying (i) and |R′g′ | < |R′g|, which is a contradiction. According to the two cases above, R′g = ∅, and so there exists a γtR(G ◦H)-function h defined as h(V (Hu)) = h(u, v) = 2 whenever g(V (Hu)) = 2 and h(V (Hu)) = g(V (Hu)) otherwise. Therefore, h satisfies (i) and (ii). Lemma 3.2. Let G be a graph with no isolated vertex and H a nontrivial graph with γ(H) = 1. Let f(V0, V1, V2) be a γtR(G ◦H)-function, A = {x ∈ V (G) : V (Hx)∩V1 ̸= ∅} and B = {x ∈ V (G) : V (Hx) ∩ V2 ̸= ∅}. If f satisfies Lemma 3.1, then B is a dominating set and A ∪B is a total dominating set of G. Proof. Let f(V0, V1, V2) be a γtR(G ◦ H)-function which satisfies Lemma 3.1. Let C = V (G) \ (A∪B). Obviously, if x ∈ C, then V (Hx) ⊆ V0, which implies that x is adjacent to some vertex in B and, since H is a nontrivial graph and f satisfies Lemma 3.1, if x ∈ A, then there exists y ∈ V (H) such that (x, y) ∈ V0, and so x is adjacent to some vertex in B. Hence, B is a dominating set of G. Now, since the subgraph of G ◦H induced by V1 ∪ V2 does not have isolated vertices, the subgraph of G induced by A∪B does not have isolated vertices, which implies that A ∪B is total dominating set of G. For any graphG, let D(G) be the set of dominating sets ofG, and Dt(G) the set of total dominating sets of G. We now proceed to introduce our main tool, which is the following domination parameter. ξ(G) = min{|A|+ 2|B| : B ∈ D(G) and A ∪B ∈ Dt(G)}. We say that an ordered pair (A,B) of subsets of V (G) is a ξ(G)-pair if B ∈ D(G), A ∪B ∈ Dt(G) and ξ(G) = |A|+ 2|B|. Theorem 3.3. For any graph G with no isolated vertex and any nontrivial graph H with γ(H) = 1, γtR(G ◦H) = ξ(G). Proof. Let v be a universal vertex ofH . From any ξ(G)-pair (A,B) we define the function f(V0, V1, V2) as V2 = B×{v}, V1 = A×{v} and V0 = V (G◦H)\(V1∪V2). Since V2 is a dominating set ofG◦H and V1∪V2 is a total dominating set ofG◦H , we can conclude that f is a TRDF onG◦H . Therefore, γtR(G◦H) ≤ ω(f) = |V1|+2|V2| = |A|+2|B| = ξ(G). Now, let f ′(V ′0 , V ′ 1 , V ′ 2) be a γtR(G ◦ H)-function which satisfies Lemma 3.1. Let A = {x ∈ V (G) : f ′(V (Hx)) = 1} and B = {x ∈ V (G) : f ′(V (Hx)) = 2}. By Lemma 3.2, B is a dominating set of G and A∪B is a total dominating set, which implies that ξ(G) ≤ |A|+ 2|B| = |V ′1 |+ 2|V ′2 | = γtR(G ◦H). Therefore, the result follows. A. Cabrera Martı́nez and J. A. Rodrı́guez-Velázquez: Closed formulas for the total Roman . . . 237 a db c e f 2 1 2 1 2 a db c e f 2 2 2 2 Figure 1: The labels correspond to two different γtR(G)-functions f1(V0, V1, V2), on the left, and f2(W0,W1,W2), on the right. In this case, γtR(G) = 2γt(G) = 8, V2 = {a, d, f} is a γ(G)-set and W2 = {a, b, e, f} is the only γt(G)-set. Let G be the graph shown in Figure 1 and H a nontrivial graph with γ(H) = 1. Notice that γtR(G ◦ H) = ξ(G) = γtR(G) = 8, where f1(V0, V1, V2) and f2(W0,W1,W2) are γtR(G)-functions for V1 = {b, e}, V2 = {a, d, f}, W1 = ∅, W2 = {a, b, e, f}. Furthermore, both (V1, V2) and (W1,W2) are ξ(G)-pairs, where V2 is a γ(G)-set and |V1|+ |V2| > γt(G), while W2 is a γt(G)-set which does not contain any γ(G)-set. The following bounds were given in [5]. In fact the lower bound was stated for any connected non-trivial graph G, although it also holds for any graph G with no isolated vertex. Theorem 3.4 ([5]). For any graph H and any graph G with no isolated vertex, γtR(G) ≤ γtR(G ◦H) ≤ 2γt(G). Furthermore, if γ(H) = 1, then γtR(G ◦H) ≤ 3γ(G). In order to improve some of these bounds, we need to introduce some additional termi- nology. Given a set S ⊆ V (G), we define ψ(S) = min{|S′| : S′ ⊆ V (G) \ S and S ⊆ N(S′ ∪ S)}. We also define the following parameter associated to G. µ(G) = min{ψ(S) : S is a γ(G)-set}. It is readily seen that 0 ≤ µ(G) ≤ γ(G). Furthermore, µ(G) = 0 if and only if γt(G) = γ(G), while µ(G) = γ(G) if and only if for every γ(G)-set S and every pair of different vertices x, y ∈ S we have that N [x] ∩ N [y] = ∅, i.e., if and only if every γ(G)-set is a 2-packing of G. With the notation above in mind, we state the following theorem. Theorem 3.5. Let G and H be two graphs with no isolated vertex. If γ(H) = 1, then max{γtR(G), γt(G) + γ(G)} ≤ γtR(G ◦H) ≤ min{2γ(G) + µ(G), 2γt(G)}. Proof. Our main tool is Theorem 3.3. For any ξ(G)-pair (A,B) we have that γtR(G◦H) = ξ(G) = 2|B|+ |A| ≥ |(A ∪B)|+ |B| ≥ γt(G) + γ(G). Now, let S be a γ(G)-set with µ(G) = ψ(S) and S′ ⊆ V (G) \ S a set of minimum cardinality among the subsets of V (G)\S satisfying that S ⊆ N(S′∪S). Since S∪S′ is a total dominating set, γtR(G◦H) = ξ(G) ≤ |S∪S′|+ |S| = 2|S|+ |S′| = 2γ(G)+µ(G). Finally, by Theorem 3.4, γtR(G) ≤ γtR(G ◦ H) ≤ 2γt(G), which completes the proof. 238 Ars Math. Contemp. 20 (2021) 233–241 Since µ(G) ≤ γ(G), we can conclude that the bound γtR(G ◦ H) ≤ 2γ(G) + µ(G) is never worse than the known bound γtR(G ◦ H) ≤ 3γ(G). In order to see that the upper bounds given by Theorem 3.5 are tight, we take the graph G shown in Figure 1 and any nontrivial graphH with γ(H) = 1. In this case, γtR(G◦H) = 2γt(G) = 2γ(G) + µ(G) = 8. We would point out the following result which is a direct consequence of Theorems 2.2 and 3.5. Theorem 3.6. If G is a graph with γt(G) = γ(G) and H is a nontrivial graph with γ(H) = 1, then γtR(G ◦H) = γtR(G) = 2γ(G). We now proceed to characterize the graphs achieving the lower bounds given by Theo- rem 3.5. Theorem 3.7. Let G and H be two graphs with no isolated vertex. If γ(H) = 1, then the following statements are equivalent. (i) γtR(G ◦H) = γtR(G). (ii) There exists a γtR(G)-function f(V0, V1, V2) such that V2 is dominating set of G. Proof. If there exists a γtR(G)-function f(V0, V1, V2) such that V2 is dominating set of G, then γtR(G ◦ H) = ξ(G) ≤ |V1 ∪ V2| + |V2| = |V1| + 2|V2| = γtR(G). Since γtR(G) ≤ γtR(G ◦H), we conclude that γtR(G ◦H) = γtR(G). Conversely, assume that γtR(G ◦ H) = γtR(G). Let g(V ′0 , V ′1 , V ′2) be a γtR(G ◦ H)- function satisfying Lemma 3.1. Let A = {x ∈ V (G) : g(V (Hx)) = 1} and B = {x ∈ V (G) : g(V (Hx)) = 2}. By Lemma 3.2, B is a dominating set of G and A ∪ B is a total dominating set. Hence, we can define a TRDF h(V ′′0 , V ′′ 1 , V ′′ 2 ) from V ′′ 1 = A and V ′′ 2 = B. Since ω(h) = |A|+ 2|B| = |V ′1 |+ 2|V ′2 | = γtR(G ◦H) = γtR(G), we conclude that h is a γtR(G)-function where V ′′2 is a dominating set, as desired. The next result gives a characterization for the case γtR(G ◦ H) = γt(G) + γ(G) whenever γ(H) = 1. Theorem 3.8. Let G and H be two graphs with no isolated vertex. If γ(H) = 1, then the following statement are equivalent. (i) γtR(G ◦H) = γt(G) + γ(G). (ii) There exists a γt(G)-set that contains some γ(G)-set. Proof. If there exists a γt(G)-set X which contains a γ(G)-set B, then γtR(G ◦ H) = ξ(G) ≤ |X \ B| + 2|B| = |X| + |B| = γt(G) + γ(G), and by (i) we conclude that γtR(G ◦H) = γt(G) + γ(G). Conversely, assume that γtR(G◦H) = γt(G)+γ(G) and let (A,B) be a ξ(G)-pair. If the total dominating setA∪B is a γt(G)-set, then we are done, asB is a dominating set and from γt(G)+ γ(G) = γtR(G ◦H) = ξ(G) = |A|+2|B| = |A∪B|+ |B| = γt(G)+ |B| we deduce that B is a γ(G)-set. Suppose to the contrary, that |A ∪ B| > γt(G). In such a case, γt(G) + γ(G) = ξ(G) = |A|+ 2|B| ≥ |A ∪B|+ |B| > γt(G) + γ(G), which is a contradiction. Therefore, the result follows. Figure 2 shows a graph G such that γtR(G ◦H) = γt(G) + γ(G) = 7 > 6 = γtR(G) for every nontrivial graph H with γ(H) = 1. A. Cabrera Martı́nez and J. A. Rodrı́guez-Velázquez: Closed formulas for the total Roman . . . 239 a db c e Figure 2: The γt(G)-set D = {a, b, d, e} contains the γ(G)-set S = {a, b, d}. 4 Small values of γtR(G ◦ H) In this short section we characterize the graphs G and H for which γtR(G ◦H) ∈ {3, 4}. Theorem 4.1. For any graph G and H with no isolated vertex, the following statements are equivalent. (i) γtR(G ◦H) = 3. (ii) γ(G) = γ(H) = 1. Proof. If γtR(G ◦ H) = 3, then by Theorem 2.4 we deduce that γ(H) = 1. Moreover, by Theorem 3.5 we have that 3 = γtR(G ◦H) ≥ γt(G) + γ(G) ≥ 3. Hence, γ(G) = 1, as required. Conversely, if γ(G) = γ(H) = 1, then by Theorem 3.8 we deduce that γtR(G ◦H) = 3. Theorem 4.2. For any graph G and H with no isolated vertex, γtR(G ◦ H) = 4 if and only if one of the following conditions are satisfied. (i) γt(G) = 2 and γ(H) ≥ 2. (ii) γt(G) = γ(G) = 2 and γ(H) = 1. Proof. We first notice that if conditions (i) or (ii) holds, then by Theorem 2.4 or by Theo- rem 3.5, respectively, it follows that γtR(G ◦H) = 4. Conversely, assume that γtR(G ◦ H) = 4. If γ(H) ≥ 2, then Theorem 2.4 leads to γt(G) = 2. From now on, we assume that γ(H) = 1. By Theorem 3.8, we have that 4 = γtR(G ◦ H) ≥ γt(G) + γ(G). Hence, 1 ≤ γ(G) ≤ 2. If γ(G) = 1, then by Theorem 4.1 we obtain that γtR(G ◦H) = 3, which is a contradiction. Hence, γ(G) = 2 and so γt(G) = 2. Therefore, the result follows. 5 Open problems By Theorem 3.3 we learned that, if we want to know the behaviour of γtR(G ◦ H) when γ(H) = 1, then it is crucial to obtain the exact value or derive tight bounds on ξ(G). In this sense, the study of ξ(G) is an interesting challenge. In particular, Theorem 3.5 states that max{γtR(G), γt(G) + γ(G)} ≤ ξ(G) ≤ min{2γ(G) + µ(G), 2γt(G)}. The graphs achieving the equalities ξ(G) = γtR(G) and ξ(G) = γt(G)+ γ(G) were char- acterized in Theorems 3.7 and 3.8, respectively. Therefore, the problems of characterizing the graphs achieving the equalities ξ(G) = 2γt(G) and ξ(G) = 2γ(G) + µ(G) = 3γ(G) remain open. 240 Ars Math. Contemp. 20 (2021) 233–241 ORCID iDs Abel Cabrera Martı́nez https://orcid.org/0000-0003-2806-4842 Juan Alberto Rodrı́guez-Velázquez https://orcid.org/0000-0002-9082-7647 References [1] H. Abdollahzadeh Ahangar, M. A. Henning, V. Samodivkin and I. G. Yero, Total Ro- man domination in graphs, Appl. Anal. Discrete Math. 10 (2016), 501–517, doi:10.2298/ aadm160802017a. [2] B. H. Arriola and S. R. Canoy, Jr., Doubly connected domination in the corona and lexico- graphic product of graphs, Appl. Math. Sci. 8 (2014), 1521–1533, doi:10.12988/ams.2014. 4136. [3] A. Cabrera Martı́nez, S. Cabrera Garcı́a and J. A. Rodrı́guez-Velázquez, Double domination in lexicographic product graphs, Discrete Appl. Math. 284 (2020), 290–300, doi:10.1016/j.dam. 2020.03.045. [4] A. Cabrera Martı́nez and J. A. Rodrı́guez-Velázquez, Total protection of lexicographic product graphs, Discuss. Math. Graph Theory (2020), in press, doi:10.7151/dmgt.2318. [5] N. Campanelli and D. Kuziak, Total Roman domination in the lexicographic product of graphs, Discrete Appl. Math. 263 (2019), 88–95, doi:10.1016/j.dam.2018.06.008. [6] M. Dettlaff, M. Lemańska, J. A. Rodrı́guez-Velázquez and R. Zuazua, On the super domination number of lexicographic product graphs, Discrete Appl. Math. 263 (2019), 118–129, doi:10. 1016/j.dam.2018.03.082. [7] R. Hammack, W. Imrich and S. Klavžar, Handbook of Product Graphs, Discrete Mathematics and its Applications, CRC Press, Boca Raton, FL, 2nd edition, 2011, doi:10.1201/b10959. [8] T. W. Haynes, S. T. Hedetniemi and P. J. Slater (eds.), Domination in Graphs, Volume 2: Ad- vanced Topics, volume 209 of Monographs and Textbooks in Pure and Applied Mathematics, Marcel Dekker, New York, 1998, doi:10.1201/9781315141428. [9] 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, New York, 1998, doi:10.1201/9781482246582. [10] M. A. Henning, A survey of selected recent results on total domination in graphs, Discrete Math. 309 (2009), 32–63, doi:10.1016/j.disc.2007.12.044. [11] 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. [12] W. Imrich and S. Klavžar, Product Graphs: Structure and Recognition, Wiley-Interscience Series in Discrete Mathematics and Optimization, Wiley-Interscience, New York, 2000. [13] D. J. Klein and J. A. Rodrı́guez-Velázquez, Protection of lexicographic product graphs, Dis- cuss. Math. Graph Theory (2019), in press, doi:10.7151/dmgt.2243. [14] T. Kraner Šumenjak, P. Pavlič and A. Tepeh, On the Roman domination in the lexicographic product of graphs, Discrete Appl. Math. 160 (2012), 2030–2036, doi:10.1016/j.dam.2012.04. 008. [15] T. Kraner Šumenjak, D. F. Rall and A. Tepeh, Rainbow domination in the lexicographic product of graphs, Discrete Appl. Math. 161 (2013), 2133–2141, doi:10.1016/j.dam.2013.03.011. [16] R. Laskar, J. Pfaff, S. M. Hedetniemi and S. T. Hedetniemi, On the algorithmic complexity of total domination, SIAM J. Algebraic Discrete Methods 5 (1984), 420–425, doi:10.1137/ 0605040. A. Cabrera Martı́nez and J. A. Rodrı́guez-Velázquez: Closed formulas for the total Roman . . . 241 [17] C.-H. Liu and G. J. Chang, Roman domination on strongly chordal graphs, J. Comb. Optim. 26 (2013), 608–619, doi:10.1007/s10878-012-9482-y. [18] J. Liu, X. Zhang and J. Meng, Domination in lexicographic product digraphs, Ars Combin. 120 (2015), 23–32. [19] R. J. Nowakowski and D. F. Rall, Associative graph products and their independence, domina- tion and coloring numbers, Discuss. Math. Graph Theory 16 (1996), 53–79, doi:10.7151/dmgt. 1023. [20] M. Valveny, H. Pérez-Rosés and J. A. Rodrı́guez-Velázquez, On the weak Roman domination number of lexicographic product graphs, Discrete Appl. Math. 263 (2019), 257–270, doi:10. 1016/j.dam.2018.03.039. [21] X. Zhang, J. Liu and J. Meng, Domination in lexicographic product graphs, Ars Combin. 101 (2011), 251–256. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 243–260 https://doi.org/10.26493/1855-3974.2359.a7b (Also available at http://amc-journal.eu) Wiener-type indices of Parikh word representable graphs* Nobin Thomas Research scholar, APJ Abdul Kalam Technological University, Thiruvananthapuram, Kerala 695 016 India, and Amal Jyothi College of Engineering, Kanjirappally, Kerala 686 518 India Lisa Mathew Amal Jyothi College of Engineering, Kanjirappally, Kerala 686 518 India Sastha Sriram Department of Mathematics, School of Arts, Science and Humanities, SASTRA Deemed University, Tanjore, Tamil Nadu 613 401 India K. G. Subramanian † School of Mathematics, Computer Science and Engineering, Liverpool Hope University, Liverpool L16 9JD, United Kingdom Received 10 June 2020, accepted 14 February 2021, published online 9 November 2021 Abstract A new class of graphs G(w), called Parikh word representable graphs (PWRG), corre- sponding to words w that are finite sequence of symbols, was considered in the recent past. Several properties of these graphs have been established. In this paper, we consider these graphs corresponding to binary core words of the form aub over a binary alphabet {a, b}. We derive formulas for computing the Wiener index of the PWRG of a binary core word. Sharp bounds are established on the value of this index in terms of different parameters related to binary words over {a, b} and the corresponding PWRGs. Certain other Wiener- type indices that are variants of Wiener index are also considered. Formulas for computing these indices in the case of PWRG of a binary core word are obtained. Keywords: Graphs, words, Parikh matrix, Parikh word representable graphs. Math. Subj. Class. (2020): 68R10, 68R15 *The authors would like to thank the reviewers for their very useful comments which enabled them to revise the paper improving the presentation of the paper. †Honorary Visiting Professor. E-mail addresses: nobinvazhayil@gmail.com (Nobin Thomas), lisamathew@amaljyothi.ac.in (Lisa Mathew), sriram.discrete@gmail.com (Sastha Sriram), kgsmani1948@gmail.com (K. G. Subramanian) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 244 Ars Math. Contemp. 20 (2021) 243–260 1 Introduction While words that are finite sequences of symbols are the fundamental and central objects in computing models developed in theoretical studies of computer science, graphs are mathe- matical models of pairwise relations between objects found useful for analyzing and solv- ing different kinds of problems. An interesting area of investigation is relating graphs and words and there are many studies in this direction (see, for example, [7, 10, 15, 18, 19, 24]). On the other hand, in the study of numerical quantities related to subwords (also called scattered subwords) of a word, the notion of Parikh matrix of a word over an ordered alphabet was introduced in [26]. This has opened up a new direction of research in the area of combinatorics on words [23] and many problems on words and subwords have been investigated (see, for example, [1, 2, 4, 25, 33, 34, 36, 37, 38] and references therein), resulting in a number of interesting results. Parikh word representable graph (PWRG) is one such notion introduced in [3] linking the two areas of study on properties of words and of graphs. Based on the notions of subwords of a word and the Parikh matrix of a word [26] with entries of the matrix giving the counts of certain subwords in the word, PWRG related to a word was introduced in [3]. Relationship of these graphs with corresponding words and partitions was recently studied in [27]. In the field of chemical graph theory [14], undirected graphs, referred to as molecular graphs are considered providing graph representations of organic compounds or equiva- lently their molecular structures with atoms other than hydrogen often represented by ver- tices and covalent chemical bonds by edges. In fact in chemical graph theory there have been attempts to capture the molecular structure in terms of the topological index of the corresponding graph. There has been a great interest in various topological indices associ- ated with graphs due to their application in the area of chemical graph theory [8]. There are a number of studies (see, for example, [14]) of various topological indices of graphs estab- lishing formulae for computing the indices and also providing upper and lower bounds on the values of such indices. The Wiener index (also called Wiener number) [40] is the first topological index introduced by Harold Wiener. Knor et al. [22] provide an excellent sum- mary of results relating to Wiener index besides providing conjectures and problems on this index. Wiener index and its variants for different classes of graphs are widely investigated indices (see, for example, [9, 12, 14, 21, 28, 29, 39] and references therein). In this paper we study the Wiener index of a PWRG of a binary core word and derive formulas for computing this index besides establishing sharp bounds on their values, given different parameters related to the graphs. We also obtain formulas for evaluating certain other indices that are variants of Wiener index, such as multiplicative Wiener index [13], terminal Wiener index [11], peripheral Wiener index [16], hyper-Wiener index [20, 30] in the case of a PWRG of a binary core word. 2 Preliminaries The basic notions and notations relating to words and subwords can be found in [23, 31]. We recall some essential concepts and results. An ordered alphabet Σ which is a set of symbols {a1, a2, . . . , as} with an ordering < on its symbols is written as Σ = {a1 < a2 < · · · < as} . A word v is a subword of a word w over Σ if and only if we can find words x1, x2, . . . , xn, y0, y1, . . . , yn over Σ, some of them possibly empty, such that w = y0x1y1x2y2 · · · yn−1xnyn and v = x1x2 · · ·xn. The number of occurrences of a word u as a subword of w is denoted by |w|u. For example, in the word w = aababaaab = N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 245 a2baba3b over the ordered binary alphabet Σ = {a < b}, the number of distinct occur- rences of the subword ab is 11 so that |w|ab = 11. The set of all words over an alphabet Σ, including the empty word λ with no symbols, is denoted by Σ∗. Definition 2.1 ([6]). A binary word w over an alphabet {a, b} is said to be fair if |w|ab = |w|ba. Example 2.2. The binary word abbbaab is a fair word since |w|ab = |w|ba = 6. Definition 2.3 ([38]). Consider the binary word w ∈ Σ∗ where Σ = {a < b}. The core of w, denoted by core(w), is the unique word w0 of w with the smallest possible length such that w ∈ b∗w0a∗. A word w ∈ Σ∗ is said to be a core word if and only if core(w) = w. Clearly, a non empty word w over Σ = {a < b} is a core word if and only if w starts with a and ends with b. We now recall the relationship between binary core words and partitions following the discussion in [38, pages 62–63]. Lemma 2.4 ([38]). Every nonempty binary core word can be identified with a partition of a positive integer. Proof. Suppose w ∈ Σ∗ is a nonempty core word and has the form an1ban2b · · · an|w|b b where n1 ≥ 1 and nk is nonnegative for each k, 2 ≤ k ≤ |w|b. Thus w can be identified with the partition |w|ab = (n1 + n2 + · · ·+ n|w|b) + · · ·+ (n2 + n1) + n1. Clearly, distinct core words are identified with distinct partitions. Conversely, suppose that m1 +m2 + · · · +ml is a partition of some positive integer, where m1 ≥ m2 ≥ · · · ≥ ml ≥ 1. It is clear that the word w = amlbaml−1−mlbaml−2−ml−1b · · · am1−m2b can be identified with the given partition. We shall use the notation p(w) = (n1+n2+ · · ·+nl)+(n1+n2+ · · ·+nl−1)+ · · ·+ (n1 +n2 + · · ·+nl−i+1) + · · ·+ (n1 +n2) + (n1) to indicate the partition corresponding to the word w = an1ban2b · · · anlb. We now recall the notion of Parikh word representable graph (PWRG) [3]. For basic concepts pertaining to graphs we refer to [5]. Definition 2.5 ([3]). For a word w = a1a2 · · · an of length n where for 1 ≤ i ≤ n, ai ∈ Σ = {a < b}, we associate a simple graph G = G(w) with n vertices {1, 2, . . . , n}. Each vertex i has the label ai and represents the position of the letter ai, 1 ≤ i ≤ n, in w. For each occurrence of the subword ab in w, there is an edge in G(w) joining the vertices corresponding to the positions of a and b in w. We say that the graph G is Parikh binary word representable by the binary word w. In other words, we say that a graph G is Parikh binary word representable if there exists a binary word w such that G is Parikh binary word representable by the binary word w. 246 Ars Math. Contemp. 20 (2021) 243–260 Since every connected Parikh binary word representable graph corresponds to a core word, we deal with only core words and the corresponding graphs in the rest of this paper. As an illustration, if the core word is w = aabab, then in the Parikh word representable graph as shown in Figure 1, the vertices 1, 2 and 4 have label a while the vertices 3 and 5 have the label b. The number of edges in the graph is |w|ab = 5. For example there is a subword ab in w formed by the symbol a in position 1 and the symbol b in position 3 and so in the graph there is an edge joining the vertex 1 with the vertex 3. 3 1 42 5 G(aabab) Figure 1: The Parikh word representable graph of the word aabab. 3 Wiener index of Parikh word representable graphs Let G = (V,E) be a connected graph with vertex set V (G) and edge set E(G). The distance between the vertices u and v of G is denoted by d(u, v) and is defined as the length of a shortest path between u and v in G. Definition 3.1. The Wiener index W (G) of a connected graph G = (V,E), is the sum of distances d(u, v) between all the vertices u and v of G. In other words W (G) = ∑ {u,v}⊆V (G) d(u, v). We now obtain a formula for computing the Wiener index of Parikh word representable graph of a binary word. Theorem 3.2. The Wiener index of a Parikh word representable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is W (G(w)) = ( l∑ i=1 ni )2 + l∑ i=1 (l + 2i− 3)ni + l(l − 1). Proof. In the Parikh word representable graph G(w) corresponding to the word w = an1ban2b · · · anlb, we consider pairs of vertices (u, v), with u, v ∈ {1, 2, . . . , n} where the label of u appears before the label of v in w. There are now four cases to be considered: (i) u and v are both labeled a; (ii) u is labeled a and v is labeled b; (iii) u is labeled b and v is labeled a; N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 247 (iv) u and v are both labeled b. The contribution to the Wiener index of the Parikh word representable graph from each of these four cases may be calculated as follows: (i) The distance between any two vertices labeled a is 2 and there are n1+n2+ · · ·+nl vertices labeled a. Hence the total contribution from these pairs of vertices is (n1 + n2 + · · ·+ nl)C2 × 2 = (n1 + n2 + · · ·+ nl)2 − (n1 + n2 + · · ·+ nl). (ii) The distance between u labeled a and v labeled b is 1 and there are n1+(n1+n2)+ · · ·+ (n1 + n2 + · · ·+ nl) such pairs. Hence the total contribution is n1 + (n1 + n2) + · · ·+ (n1 + n2 + · · ·+ nl) = ln1 + (l − 1)n2 + · · ·+ nl. (iii) The distance between u labeled b and v labeled a is 3 and there are (n2 +n3 + · · ·+ nl) + (n3 + · · · + nl) + · · · + nl = n2 + 2n3 + · · · + (l − 1)nl such pairs. Hence the total contribution is 3(n2 + 2n3 + · · ·+ (l − 1)nl). (iv) The distance between any two vertices labeled b is 2 and there are l vertices labeled b. Hence the total contribution from these pairs of vertices is lC2 × 2 = l(l − 1). Hence the Wiener index of G(w) is given by W (G(w)) = (n1 + n2 + · · ·+ nl)2 + (l − 1)n1 + (l + 1)n2 + · · · + 3(l − 1)nl + l(l − 1) = ( l∑ i=1 ni )2 + l∑ i=1 (l + 2i− 3)ni + l(l − 1). Example 3.3. For the PWRG in Figure 1 corresponding to the word w = a2bab, we have l = 2, n1 = 2, n2 = 1 and so W (G(w)) = 16 which can also be verified from the formula in the Definition 3.1 by actually computing the distances d(u, v) for all unordered pairs (u, v) of vertices. We now derive an alternate form of the expression for the Wiener index of the Parikh word representable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l. An interesting aspect of this alternate form is that the expression in the formula is elegant involving only the parameters related to the word. Theorem 3.4. The Wiener index of a Parikh word representable graph G(w), for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is W (G(w)) = |w|2 − |w|+ |w|a|w|b − 2|w|ab. 248 Ars Math. Contemp. 20 (2021) 243–260 Proof. Since w = an1ban2b · · · anlb, we have ∑l i=1 ni = |w|a, l = |w|b. Also |w|a|w|b − |w|ab = l ( l∑ i=1 ni ) − [(n1 + n2 + · · ·+ nl) + · · ·+ (n1 + n2) + n1] = l∑ i=1 ini − |w|a so that l∑ i=1 ini = |w|a|w|b + |w|a − |w|ab. Hence from Theorem 3.2, the Wiener index W (G(w)) = |w|2a + (|w|b − 3)|w|a + |w|b(|w|b − 1) + 2 l∑ i=1 ini = |w|2a + |w|2b + 3|w|a|w|b − |w|a − |w|b − 2|w|ab = |w|2 − |w|+ |w|a|w|b − 2|w|ab using |w| = |w|a + |w|b. Corollary 3.5. The Wiener index of a Parikh word representable graph G(w) for a fair word w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is W (G(w)) = |w|2 − |w|. Proof. For a binary word w, we have |w|ab + |w|ba = |w|a|w|b. Since w is a fair word, |w|ab = |w|ba so that |w|a|w|b − 2|w|ab = |w|ba − |w|ab = 0. Hence from Theorem 3.4, W (G(w)) = |w|2 − |w|. Theorem 3.6. The Wiener index W (G(w)) of a Parikh word representable graph G(w) = (V1 ∪ V2, E) with |V1| = |w|a = p, |V2| = |w|b = q for the word w = an1ban2b · · · anqb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is bounded above by p2 + q2 + 3pq − 3p − 3q + 2 and below by p2 + q2 + pq − p− q. The bounds are attained on G(abq−1ap−1b) and G(apbq) respectively. Proof. Since G(w) is connected, |w|ab = |E| ≥ p + q − 1 [5]. Also |w|ab ≤ pq [26]. Hence from Theorem 3.4, the Wiener index of G(w) is W (G(w)) = p2 + q2 + 3pq − p− q − 2|w|ab ≤ p2 + q2 + 3pq − p− q − 2(p+ q − 1) = p2 + q2 + 3pq − 3p− 3q + 2 which is the Wiener index of the Parikh word representable graph G(abq−1ap−1b) and W (G(w)) ≥ p2 + q2 + 3pq − p− q − 2pq = p2 + q2 + pq − p− q which is the Wiener index of the Parikh word representable graph G(apbq). N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 249 Remark 3.7. One of the conjectures listed in [22, page 333], states that for a graph G with diameter d and order 2d + 1, the Wiener index W (G) ≤ W (C2d+1) where C2d+1 denotes a cycle of length 2d+ 1. Since the diameter of any PWRG G(w) corresponding to the binary core word w is 3, this conjecture holds good for G(w), if the order of G(w) is 7, which also equals |w|. In fact, if the binary core word w with |w| = 7, is over the ordered alphabet {a < b}, the maximum Wiener index of G(w) equals W (C7) which is 42 and this is attained, by Theorem 3.6, on G(ab3a2b) or G(ab2a3b). We shall now find an expression for an upper bound on the Wiener index of Parikh word representable graph with a fixed number of edges. We use the following lemma. Lemma 3.8. Given a fixed value of e and of l, the maximum value of W (G(w)) over all Parikh word representable graphs of the form G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges, is attained on G(abl−1ae−lb). Proof. Since the number of edges in G(w) is e = |w|ab, |w|b = l and e ≥ |w|a + |w|b − 1 as G(w) is a connected graph with |w|a + |w|b vertices, we have |w|a ≤ e− l + 1. Hence from Theorem 3.4, the Wiener index of G(w) is W (G(w)) = |w|a(|w|a − 1) + l2 + 3|w|al − l − 2e ≤ (e− l + 1)(e− l) + l2 + 3|w|al − l − 2e ≤ (e− l + 1)(e− l) + l2 + 3(e− l + 1)l − l − 2e = e2 − l2 + el + l − e which is the Wiener index of the Parikh word representable graph G(abl−1ae−lb). Theorem 3.9. An upper bound of the Wiener index W (G(w)) of a Parikh word repre- sentable graph G(w), w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges is given by W (G(w)) ≤ { 5m2 − 6m+ 2(m ≥ 1), if e = 2m− 1; 5m2 −m(m ≥ 1), if e = 2m. The bound is sharp and is attained on G(abm−1am−1b) when e = 2m − 1 and on G(abm−1amb) when e = 2m. Proof. From Lemma 3.8, the Wiener index of the Parikh word representable graph G(w), w = an1ban2b · · · anlb, n1 ≥ 1, has a maximum for G(abl−1ae−lb) and is given by W (G(w)) = e2 − l2 + el + l − e. We now use the fact that a quadratic expression ax2 + bx + c, a < 0, has a maximum when x = − b2a . When e = 2m − 1,m ≥ 1, we have W (G(w)) = −l2 + 2ml + (4m2 − 6m+ 2). If e = 2m− 1 has a fixed value, this quadratic expression in l has a maximum when l = m and the maximum is 5m2 − 6m+ 2. When e = 2m,m ≥ 1, we have W (G(w)) = −l2 + l(1 + 2m) + (4m2 − 2m). Again if e = 2m has a fixed value, this quadratic expression has a maximum when l = [m+ 12 ] = m where [x] is the integral part of x and the maximum is 5m 2 + 2m. We shall now evaluate the Wiener index of a Parikh word representable graph corre- sponding to a specific partition of a given integer. 250 Ars Math. Contemp. 20 (2021) 243–260 Theorem 3.10. Suppose m1+m2+ · · ·+ml is a partition of some positive integer, where m1 ≥ m2 ≥ · · · ≥ ml ≥ 1. Then the Wiener index of the Parikh word representable graph G corresponding to this partition is given by W (G) = m21 − 2e+ (3l − 1)m1 + l(l − 1) where e is the number of edges of G. Proof. From Lemma 2.4, the word w = amlbaml−1−mlbaml−2−ml−1b · · · am1−m2b corre- sponds to the given partition. Now |w| = m1 + l, |w|a = m1, |w|b = l, so that using the formula in Theorem 3.4, we have W (G) = W (G(w)) = |w|2 − |w|+ |w|a|w|b − 2|w|ab = (m1 + l) 2 − (m1 + l) + lm1 − 2e = m21 − 2e+ (3l − 1)m1 + l(l − 1) since |w|ab = e. 4 Multiplicative Wiener index Definition 4.1 ([32]). The Wiener polynomial of a graph G is W (G;x) = ∑ {u,v}⊆V (G) xd(u,v). Theorem 4.2. The Wiener polynomial of a Parikh word representable graph G(w), for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is W (G(w);x) = (|w|a|w|b−|w|ab)x3+ 1 2 ( |w|a(|w|a−1)+ |w|b(|w|b−1) ) x2+(|w|ab)x. Proof. We consider pairs of vertices (u, v) in the Parikh word representable graph G(w) corresponding to the word w = an1ban2b · · · anlb, with u, v ∈ {1, 2, . . . , n} where the label of u appears before the label of v in w. As discussed in the proof of Theorem 3.2, the vertex pairs are of four types, namely, types 1, 2, 3 and 4. Also, as discussed in the proof of Theorem 3.4, there are 12 |w|a(|w|a − 1) pairs of vertices of type 1 and 1 2 |w|b(|w|b − 1) of type 4 and the distance between each such pair is 2. Likewise, there are |w|ab pairs of vertices of type 2 with distance 1 and |w|a|w|b − |w|ab pairs of vertices of type 3 with distance 3. Hence the Wiener polynomial is (|w|a|w|b − |w|ab)x3 + 1 2 ( |w|a(|w|a − 1) + |w|b(|w|b − 1) ) x2 + (s|w|ab)x. Definition 4.3 ([40]). The Wiener polarity index of G, denoted by Wp(G), is defined as Wp(G) = |{(u, v) ⊆ V (G) : d(u, v) = 3}|. Theorem 4.4. The Wiener polarity index of a Parikh word representable graph G(w), for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is given by Wp(G(w)) = |w|a|w|b − |w|ab. N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 251 Proof. In the Parikh word representable graph G(w), for the word w = an1ban2b · · · anlb as in the hypothesis, the pairs of vertices (u, v) of type 3 as mentioned in the proof of Theorem 4.2, are at a distance 3 and these pairs contribute to the Wiener polarity index of G(w). In fact there are n2 + n3 + · · ·+ nl vertices with label a that are at distance 3 from the vertex with label, the first b in w. Likewise for other vertices corresponding to the other b′s in w. Hence Wp(G(w)) = (n2 + n3 + · · ·+ nl) + (n3 + n4 + · · ·+ nl) + · · ·+ nl = l∑ i=1 ini − l∑ i=1 ni = w|a|w|b − |w|ab. Definition 4.5 ([13]). The multiplicative version of the Wiener index of a graph G is π(G) = ∏ {u,v}⊆V (G) d(u, v). Theorem 4.6. The multiplicative Wiener index of a Parikh word representable graph G(w), for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is π(G(w)) = 2 |w|a(|w|a−1)+|w|b(|w|b−1) 2 3|w|a|w|b−|w|ab . Proof. Considering pairs of vertices as in Theorem 4.2, we obtain the required result, since there are |w|a(|w|a−1)+|w|b(|w|b−1)2 pairs of vertices at distance 2 in G(w) while there are |w|a|w|b − |w|ab pairs of vertices at distance 3. It is to be noted that pairs of vertices at distance 1 contribute value 1 to the product defining π(G(w)). Corollary 4.7. The multiplicative Wiener index of a Parikh word representable graph G(w) for a fair word w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is π(G(w)) = 2 |w|2−|w| 2 ( 3 2 )|w|ab . Proof. For a binary word w, we have |w|ab + |w|ba = |w|a|w|b. Since w is a fair word, |w|ab = |w|ba so that |w|a|w|b − 2|w|ab = |w|ba − |w|ab = 0. Also |w| = |w|a + |w|b. Hence from Theorem 4.6, π(G(w)) = 2 |w|2−|w| 2 ( 3 2 )|w|ab . Lemma 4.8. Given a fixed value of e and of l, the maximum value of π(G(w)) over all Parikh word representable graphs of the form G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges, is attained on G(abl−1ae−lb). Proof. Since |w|b = l and the number of edges in G(w) is e = |w|ab, we have |w|a ≤ e− l+1 as G(w) is a connected graph with |w|a+ |w|b vertices so that e ≥ |w|a+ |w|b−1. Note that in a connected graph G with n vertices and e edges, e ≥ n− 1 [5]. Hence from Theorem 4.6, the multiplicative Wiener index of G(w) is π(G(w)) = 2 |w|a(|w|a−1)+|w|b(|w|b−1) 2 3|w|a|w|b−|w|ab ≤ 2 (e−l+1)(e−l)+l(l−1) 2 3(e−l+1)l−e which is the multiplicative Wiener index of the Parikh word representable graph G(abl−1 ae−lb). 252 Ars Math. Contemp. 20 (2021) 243–260 Theorem 4.9. An upper bound of the multiplicative Wiener index π(G(w)) of a Parikh word representable graph G(w), w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges is given by π(G(w)) ≤ { 2m 2−m3m 2−2m+1(m ≥ 1), if e = 2m− 1; 2m 2 3m 2−m(m ≥ 1), if e = 2m. The bound is sharp and is attained on G(abm−1am−1b) when e = 2m − 1 and on G(abm−1amb) when e = 2m. Proof. From Lemma 4.8, the multiplicative Wiener index of the Parikh word representable graph G(w), w = an1ban2b · · · anlb, n1 ≥ 1, has a maximum for G(abl−1ae−lb) and is given by π(G(w)) = 2 (e−l+1)(e−l)+l(l−1) 2 3(e−l+1)l−e. On taking logarithms we get, ln(π(G(w))) = [(e− l + 1)(e− l) + l2 − l] ln 2 2 + [(e− l + 1)l − e] ln 3 = l2(ln 2− ln 3)− l((e+ 1)(ln 2− ln 3) + e ( (e+ 1) 2 ln 2− ln 3 ) . We now use the fact that a quadratic expression ax2+ bx+ c, a < 0, has a maximum when x = − b2a . Let F (l) = l 2(ln 2− ln 3)− l((e+1)(ln 2− ln 3)+ e( (e+1)2 ln 2− ln 3). When e = 2m− 1,m ≥ 1, F (l) has a maximum when l = m and so π(G(w)) has the maximum 2m 2−m3m 2−2m+1. When e = 2m,m ≥ 1, F (l) has a maximum when l = [m + 12 ] = m where [x] is the integral part of x and π(G(w)) has the maximum 2m 2 3m 2−m. Theorem 4.10. The multiplicative Wiener index π(G(w)) of a Parikh word representable graph G(w) = (V1 ∪ V2, E) with |V1| = |w|a = p, |V2| = |w|b = q for the word w = an1ban2b · · · anqb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is bounded above by 2 p(p−1)+q(q−1) 2 3pq−p−q+1 and below by 2 p(p−1)+q(q−1) 2 . The bounds are attained on G(abq−1ap−1b) and G(apbq) respectively. Proof. Since G(w) is connected, |w|ab = |E| ≥ p + q − 1 [5]. Also |w|ab ≤ pq [26]. Hence from Theorem 4.6, the multiplicative Wiener index of G(w) is π(G(w)) ≥ 2 p(p−1)+q(q−1) 2 which is the multiplicative Wiener index of the Parikh word representable graph G(apbq) and π(G(w)) ≤ 2 p(p−1)+q(q−1) 2 3pq−p−q+1 which is the multiplicative Wiener index of the Parikh word representable graph G(abq−1 ap−1b). N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 253 5 Hyper-Wiener index of Parikh word representable graphs We now derive formuas for computing hyper-Wiener index of Parikh word representable graphs of binary words. Definition 5.1 ([20, 30]). The hyper-Wiener index of a connected graph G is given by WW (G) = ∑ {u,v}⊆V (G) d(u, v) + d2(u, v) where d(u, v) is the distance between the vertices u and v of G. Theorem 5.2. The hyper-Wiener index WW (G(w)) of a Parikh word representable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is WW (G(w)) = 3 ( l∑ i=1 ni )2 + l∑ i=1 (2l + 10i− 13)ni + 3l(l − 1). Proof. As in the proof of Theorem 3.2, there are four cases for the vertex pairs (u, v). The contribution to the hyper-Wiener index of the Parikh word representable graph from each of these cases may be calculated as follows: (i) The contribution from pairs of vertices labeled a is (n1+n2+ · · ·+nl)C2× (2+4) = 3(n1+n2+ . . .+nl)2−3(n1+n2+ · · ·+nl). (ii) The contribution from pairs of vertices (u, v) where u is labeled a and v is labeled b is 2(n1 + (n1 + n2) + . . .+ (n1 + n2 + . . .+ nl)) = 2(ln1 + (l − 1)n2 + · · ·+ nl). (iii) The contribution from pairs of vertices (u, v) where u is labeled b and v is labeled a is 12(n2 + 2n3 + · · ·+ (l − 1)nl)). (iv) The contribution from pairs of vertices labeled b is lC2 × 6 = 3l(l − 1). Hence the hyper-Wiener index of G(w) is given by WW (G(w) = 3 ( l∑ i=1 ni )2 + l∑ i=1 (2l + 10i− 13)ni + 3l(l − 1). We now derive an alternate form of the expression for the hyper-Wiener index of the Parikh word representable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l. 254 Ars Math. Contemp. 20 (2021) 243–260 Theorem 5.3. The hyper-Wiener index of a Parikh word representable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is WW (G(w)) = 3|w|2a + 3|w|2b + 12|w|a|w|b − 3|w|a − 3|w|b − 10|w|ab = 3|w|2 − 3|w|+ 6|w|a|w|b − 10|w|ab. Proof. Since w = an1ban2b · · · anlb, we have ∑l i=1 ni = |w|a, l = |w|b. Also, as in the proof of Theorem 3.4, l∑ i=1 ini = |w|a|w|b + |w|a − |w|ab. Hence from Theorem 5.2, the hyper-Wiener index WW (G(w)) = 3|w|2a + (2|w|b − 13)|w|a + 3|w|b(|w|b − 1) + 10 l∑ i=1 ini = 3|w|2a + 3|w|2b + 12|w|a|w|b − 3|w|a − 3|w|b − 10|w|ab = 3|w|2 − 3|w|+ 6|w|a|w|b − 10|w|ab using |w| = |w|a + |w|b. Theorem 5.4. The hyper-Wiener index of a Parikh word representable graph G(w) = (V1 ∪ V2, E) with |V1| = |w|a = p, |V2| = |w|b = q for the word w = an1ban2b · · · anqb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is bounded above and below by 3(p2 + q2) + 12pq − 13p− 13q + 10 and 3(p2 + q2) + 2pq − 3p− 3q. The bounds are attained on G(abq−1ap−1b) and G(apbq) respectively. Proof. As done in the proof of Theorem 3.6, using the inequalities |w|ab ≥ p+ q − 1 [5], |w|ab ≤ pq [26], we have from Theorem 5.3, the hyper-Wiener index of G(w) is WW (G(w)) = 3p2 + 3q2 + 12pq − 3p− 3q − 10|w|ab ≤ 3(p2 + q2) + 12pq − 13p− 13q + 10 which is the hyper-Wiener index of the Parikh word representable graph G(abq−1ap−1b) and W (G(w)) ≥ 3(p2 + q2) + 2pq − 3p− 3q which is the Wiener index of the Parikh word representable graph G(apbq). The hyper-Wiener index of a Parikh word representable graph corresponding to a spe- cific partition of a given integer can be evaluated proceeding as in the proof of Theorem 3.10 and is given in the following theorem. N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 255 Theorem 5.5. Suppose m1 +m2 + · · ·+ml is a partition of some positive integer, where m1 ≥ m2 ≥ · · · ≥ ml ≥ 1. Then the hyper-Wiener index of the Parikh word representable graph G corresponding to this partition is given by WW (G) = 3(m21 + (4l − 1)m1 + l(l − 1))− 10e where e is the number of edges of G. We shall now find an expression for an upper bound on the hyper-Wiener index of Parikh word representable graph with a fixed number of edges. We use the following lemma. Lemma 5.6. Given a fixed value of e and of l, the maximum value of WW (G(w)) over all Parikh word representable graphs of the form G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges, is attained on G(abl−1ae−lb). Proof. Proceeding as done in the proof of Lemma 3.8, the number of edges in G(w) is e = |w|ab, |w|b = l and e ≥ |w|a+ |w|b−1 as G(w) is a connected graph with |w|a+ |w|b vertices, we have |w|a ≤ e − l + 1. Hence from Theorem 5.3, the hyper-Wiener index of G(w) is WW (G(w)) = 3|w|a(|w|a − 1) + 3l2 + 12l|w|a − 3l − 10e ≤ 3(e− l + 1)(e− l) + 3l2 + 12l(e− l + 1)− 3l − 10e = 3e2 − 6l2 + 6el + 6l − 7e which is the hyper-Wiener index of the Parikh word representable graph G(abl−1ae−lb). Theorem 5.7. An upper bound of the Wiener index W (G(w)) of a Parikh word repre- sentable graph G(w) for w = an1ban2b · · · anlb, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, with e edges is given by WW (G) ≤ { 18m2 − 26m+ 10(m ≥ 1), if e = 2m− 1; 18m2 − 8m(m ≥ 1), if e = 2m. The bound is sharp and is attained on G(abm−1am−1b) when e = 2m − 1 and on G(abm−1amb) when e = 2m. Proof. From Lemma 5.6, the hyper-Wiener index of the Parikh word representable graph G(w), w = an1ban2b · · · anlb, n1 ≥ 1, has a maximum for G(w) for w = abl−1ae−lb and is given by WW (G(w)) = 3e2 − 6l2 + 6el + 6l − 7e. We use the fact that a quadratic expression ax2+bx+c, a < 0, has a maximum when x = − b2a . When e = 2m−1,m ≥ 1, we have WW (G(w)) = −6l2 + 12ml + (12m2 − 26m+ 10). If e = 2m− 1 has a fixed value, this quadratic expression in l has a maximum when l = m and the maximum is 18m2 − 26m + 10. When e = 2m,m ≥ 1, we have WW (G(w)) = −6l2 + l(6 + 12m) + (12m2 − 14m). Again if e = 2m has a fixed value, this quadratic expression has a maximum when l = [m+ 12 ] = m where [x] is the integral part of x and the maximum is 18m2 − 8m. 256 Ars Math. Contemp. 20 (2021) 243–260 6 Terminal and peripheral Wiener indices By considering only pendant vertices in a graph, a special kind of Wiener index, called terminal Wiener index, has been introduced and studied [11]. Here we consider this notion in the context of Parikh word representable graphs. Definition 6.1 ([11]). The terminal Wiener index TW (G) of a connected graph G is the sum of distances between all pairs of pendant vertices of G. Theorem 6.2. The terminal Wiener index of a Parikh word representable graph G = G(an1ban2b · · · anlb), n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ l, is given by TW (G) = { nl(nl − 1), if n1 ̸= 1; nl(nl − 1) + k(k − 1) + 3knl, if n1 = 1, for 2 ≤ j ≤ k ≤ l, nj = 0. Proof. If n1 ̸= 1, the only possible pendant vertices are the a’s in the last block and the distance between any two of them is 2 so that TW (G) = 2× nlC2 = nl(nl − 1). Note that if nl = 0, then TW (G) = 0. If n1 = 1, n2 = n3 = · · · = nk = 0 and nk+1 ̸= 0, 2 ≤ k < l, l > 1, then the first k b’s are also a pendant vertices and the distance between any one of these b’s and any one of the a’s in the last block is 3 while that between any two b’s is 2 as well as the distance between any two a′s in the last block is 2. Hence TW (G) = nl(nl − 1) + k(k − 1) + 3knl. When k = l, then n1 = 1 and ni = 0, 2 ≤ i ≤ l so that all the b′s are the only pendant vertices and hence TW (G) = l(l − 1). If n1 = 1, l = 1, clearly TW (G) = 1 as the corresponding graph G(ab) has only one edge with the end labels a and b. Another special kind of Wiener index, called peripheral Wiener index, has also been studied [16]. Definition 6.3 ([16]). The peripheral Wiener index of a connected graph G is given by Wp(G) = ∑ {u,v}⊆P (G) d(u, v) where P (G) is the set of all peripheral vertices which are the vertices of G with their eccentricities equal to diameter of G. Theorem 6.4. The peripheral Wiener index of the Parikh word representable graph G(w) for w = an1ban2ban3b · · · anrbl−r+1, n1 ≥ 1, nk ≥ 0 for 2 ≤ k ≤ r − 1, and nr > 0 with r at least 2 and at the most l, is PW (G(w)) = ( r∑ i=2 ni )2 + r∑ i=2 (r + 2i− 4)ni + (r − 1)(r − 2). Furthermore PW (G(w)) = W (G(w)) if w = anbl, where W (G(w)) is the Wiener index of G(w). N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 257 Proof. From the definition of G(w), the Parikh word representable graph corresponding to the word w = an1ban2ban3b · · · anrbl−r+1, it is clear that G(w) is a bipartite graph of diameter 3 and radius 2. Also,the vertices corresponding to all a′s in ani , 2 ≤ i ≤ r and to all b′s except the last l − r + 1 b′s are of eccentricity 3 and hence they are the peripheral vertices of G(w). We consider pairs of peripheral vertices (u, v), where the label of u appears before the label of v in w. As in Theorem 3.2, the vertex pairs of (u, v) can be divided into four cases as given below: 1. u and v are both labeled a; 2. u is labeled a and v is labeled b; 3. u is labeled b and v is labeled a; 4. u and v are both labeled b. The contribution to the peripheral Wiener index of the Parikh word representable graph from each of these cases may be calculated as follows: 1. The contribution from the pairs of peripheral vertices labeled a is (n2+n3+n4+· · ·+nr)C2×2 = (n2+n3+n4+· · ·+nr)2−(n2+n3+n4+· · ·+nr). 2. The contribution from the pairs of peripheral vertices (u, v) where u is labeled a and v is labeled b is n2 + (n2 + n3) + (n2 + n3 + n4) + · · ·+ (n2 + n3 + · · ·+ nr−1) = (r − 2)n2 + (r − 3)n3 + · · ·+ nr−1. 3. The contribution from the pairs of peripheral vertices (u, v) where u is labeled b and v is labeled a is 3[(n2 + n3 + · · ·+ nr) + (n3 + n4 + · · ·+ nr) + · · ·+ nr] = 3[(r − 1)nr + (r − 2)nr−1 + · · ·+ n2]. 4. The contribution from the pairs of peripheral vertices labeled b is (r − 1)C2 × 2 = (r − 1)(r − 2). Hence the peripheral Wiener index of G(w) is given by PW (G(w)) = ( r∑ i=2 ni )2 + r∑ i=2 (r + 2i− 4)ni + (r − 1)(r − 2). Furthermore, if w = anbl then all the vertices of G(w) labeled a as well as b are peripheral vertices of G(w). Thus PW (G(w)) = W (G(w)) where W (G(w)) is the Wiener index of G(w). 258 Ars Math. Contemp. 20 (2021) 243–260 7 Conclusion We have derived formulas for evaluating the Wiener index and certain other variants of Wiener topological indices for Parikh word representable graphs [3] of binary core words. There are problems that remain to be investigated. For example, the lower bound of Wiener index of a Parikh word representable graph G(w) of a binary core word w when the number of edges of G(w) is a given fixed value, needs to be examined. Bipartite graphs have been utilized in studies of structural features in the areas of molecular biology and chemistry (see, for example, [17, 35]). Parikh word representable graphs (PWRG) corresponding to binary core words are bipartite graphs. It will be of interest to examine the role of PWRG in such studies of structural features and relationships. Also, it will also be of interest to study other kinds of topological indices for this class of graphs. ORCID iDs Nobin Thomas https://orcid.org/0000-0002-4057-3220 Lisa Mathew https://orcid.org/0000-0002-3722-3326 Sastha Sriram https://orcid.org/0000-0003-0604-2159 K. G. Subramanian https://orcid.org/0000-0001-8726-5850 References [1] A. Atanasiu, Binary amiable words, Internat. J. Found. Comput. Sci. 18 (2007), 387–400, doi: 10.1142/s0129054107004735. [2] A. Atanasiu, R. Atanasiu and I. Petre, Parikh matrices and amiable words, Theoret. Comput. Sci. 390 (2008), 102–109, doi:10.1016/j.tcs.2007.10.022. [3] S. Bera and K. Mahalingam, Structural properties of word representable graphs, Math. Comput. Sci. 10 (2016), 209–222, doi:10.1007/s11786-016-0257-1. [4] S. Bera, K. Mahalingam and K. G. Subramanian, Properties of Parikh matrices of binary words obtained by an extension of a restricted shuffle operator, Internat. J. Found. Comput. Sci. 29 (2018), 403–413, doi:10.1142/s0129054118500119. [5] G. Bondy and U. Murty, Graph Theory with Applications, North-Hollans, 1982. [6] A. Černý, On fair words, J. Autom. Lang. Comb. 14 (2009), 163–174, doi:10.25596/ jalc-2009-163. [7] A. Collins, S. Kitaev and V. V. Lozin, New results on word-representable graphs, Discrete Appl. Math. 216 (2017), 136–141, doi:10.1016/j.dam.2014.10.024. [8] M. V. Diudea, Basic Chemical Graph Theory, in: Multi-shell Polyhedral Clusters, Springer, Cham, volume 10 of Carbon Materials: Chemistry and Physics, 2018 pp. 1–21, doi:10.1007/ 978-3-319-64123-2 1. [9] M. Eliasi, G. Raeisi and B. Taeri, Wiener index of some graph operations, Discrete Appl. Math. 160 (2012), 1333–1344, doi:10.1016/j.dam.2012.01.014. [10] A. L. L. Gao, S. Kitaev and P. B. Zhang, On 132-representable graphs, Australas. J. Combin. 69 (2017), 105–118, https://ajc.maths.uq.edu.au/pdf/69/ajc_v69_p105. pdf. [11] I. Gutman, B. Furtula and M. Petrović, Terminal Wiener index, J. Math. Chem. 46 (2009), 522–531, doi:10.1007/s10910-008-9476-2. N. Thomas et al.: Wiener-type indices of Parikh word representable graphs 259 [12] I. Gutman, S. Li and W. Wei, Cacti with n-vertices and t cycles having extremal Wiener index, Discrete Appl. Math. 232 (2017), 189–200, doi:10.1016/j.dam.2017.07.023. [13] I. Gutman, W. Linert, I. Lukovits and Ž. Tomović, The multiplicative version of the Wiener index, J. Chem. Inf. Comput. Sci. 40 (2000), 113–116, doi:10.1021/ci990060s. [14] I. Gutman and O. E. Polansky, Mathematical Concepts in Organic Chemistry, Springer-Verlag, Berlin, 1986, doi:10.1007/978-3-642-70982-1. [15] M. M. Halldórsson, S. Kitaev and A. Pyatkin, Graphs capturing alternations in words, in: Y. Gao, H. Lu, S. Seki and S. Yu (eds.), Developments in Language Theory, Springer, Berlin, Heidelberg, volume 6224 of Lecture Notes in Computer Science, 2010 pp. 436–437, doi: 10.1007/978-3-642-14455-4 41, proceedings of the 14th International Conference (DLT 2010) held at the University of Western Ontario, London, ON, August 17 – 20, 2010. [16] H. Hua, On the peripheral Wiener index of graphs, Discrete Appl. Math. 258 (2019), 135–142, doi:10.1016/j.dam.2018.11.031. [17] P. Iyer and J. Bajorath, Mechanism-based bipartite matching molecular series graphs to iden- tify structural modifications of receptor ligands that lead to mechanism hopping, Med. Chem. Commun. 3 (2012), 441–448, doi:10.1039/c2md00281g. [18] S. Kitaev and V. Lozin, Words and Graphs, Monographs in Theoretical Computer Science, An EATCS Series, Springer, Cham, 2015, doi:10.1007/978-3-319-25859-1. [19] S. Kitaev, P. Salimov, C. Severs and H. Ulfarsson, Word-representability of line graphs, Open J. Discrete Math. 1 (2011), 96–101, doi:10.4236/ojdm.2011.12012. [20] D. J. Klein, I. Lukovits and I. Gutman, On the definition of the hyper-Wiener index for cycle- containing structures, J. Chem. Inf. Comput. Sci. 35 (1995), 50–52, doi:10.1021/ci00023a007. [21] M. Knor and R. Škrekovski, Wiener index of generalized 4-stars and of their quadratic line graphs, Australas. J. Combin. 58 (2014), 119–126, https://ajc.maths.uq.edu.au/ pdf/58/ajc_v58_p119.pdf. [22] M. Knor, R. Škrekovski and A. Tepeh, Mathematical aspects of Wiener index, Ars Math. Con- temp. 11 (2016), 327–352, doi:10.26493/1855-3974.795.ebf. [23] M. Lothaire, Combinatorics on Words, Cambridge Mathematical Library, Cambridge Univer- sity Press, Cambridge, 1997, doi:10.1017/cbo9780511566097. [24] Y. Mandelshtam, On graphs representable by pattern-avoiding words, Discuss. Math. Graph Theory 39 (2019), 375–389, doi:10.7151/dmgt.2128. [25] A. Mateescu and A. Salomaa, Matrix indicators for subword occurrences and ambiguity, Inter- nat. J. Found. Comput. Sci. 15 (2004), 277–292, doi:10.1142/s0129054104002418. [26] A. Mateescu, A. Salomaa, K. Salomaa and S. Yu, A sharpening of the Parikh mapping, Theor. Inform. Appl. 35 (2001), 551–564, doi:10.1051/ita:2001131. [27] L. Mathew, N. Thomas, S. Bera and K. G. Subramanian, Some results on Parikh word repre- sentable graphs and partitions, Adv. Appl. Math. 107 (2019), 102–115, doi:10.1016/j.aam.2019. 02.009. [28] K. Pattabiraman and P. Paulraja, Wiener index of the tensor product of a path and a cycle, Discuss. Math. Graph Theory 31 (2011), 737–751, doi:10.7151/dmgt.1576. [29] K. Pattabiraman and P. Paulraja, Wiener and vertex PI indices of the strong product of graphs, Discuss. Math. Graph Theory 32 (2012), 749–769, doi:10.7151/dmgt.1647. [30] M. Randić, Novel molecular descriptor for structure—property studies, Chem. Phys. Lett. 211 (1993), 478–483, doi:10.1016/0009-2614(93)87094-j. 260 Ars Math. Contemp. 20 (2021) 243–260 [31] G. Rozenberg and A. Salomaa (eds.), Handbook of Formal Languages, Volume 1: Word, Lan- guage, Grammar, Springer-Verlag, Berlin, 1997, doi:10.1007/978-3-642-59136-5. [32] B. E. Sagan, Y.-N. Yeh and P. Zhang, The Wiener polynomial of a graph, Int. J. Quantum Chem. 60 (1996), 959–969, doi:10.1002/(sici)1097-461x(1996)60:5⟨959::aid-qua2⟩3.0.co;2-w. [33] A. Salomaa, Parikh matrices: subword indicators and degrees of ambiguity, in: H.-J. Böckenhauer, D. Komm and W. Unger (eds.), Adventures Between Lower Bounds and Higher Altitudes, Springer, Cham, volume 11011 of Lecture Notes in Computer Science, pp. 100–112, 2018, doi:10.1007/978-3-319-98355-4 7. [34] K. G. Subramanian, A. M. Huey and A. K. Nagar, On Parikh matrices, Internat. J. Found. Comput. Sci. 20 (2009), 211–219, doi:10.1142/s0129054109006528. [35] W. R. Taylor, Protein structure comparison using bipartite graph matching and its application to protein structure classification, Mol. Cell. Proteom. 1 (2002), 334–339, doi:10.1074/mcp. t200001-mcp200. [36] W. C. Teh, On core words and the Parikh matrix mapping, Internat. J. Found. Comput. Sci. 26 (2015), 123–142, doi:10.1142/s0129054115500069. [37] W. C. Teh, Parikh-friendly permutations and uniformly Parikh-friendly words, Australas. J. Combin. 76 (2020), 208–219, https://ajc.maths.uq.edu.au/pdf/76/ajc_v76_ p208.pdf. [38] W. C. Teh and K. H. Kwa, Core words and Parikh matrices, Theoret. Comput. Sci. 582 (2015), 60–69, doi:10.1016/j.tcs.2015.03.037. [39] H. Wang, The extremal values of the Wiener index of a tree with given degree sequence, Dis- crete Appl. Math. 156 (2008), 2647–2654, doi:10.1016/j.dam.2007.11.005. [40] H. Wiener, Structural determination of paraffin boiling points, J. Am. Chem. Soc. 69 (1947), 17–20, doi:10.1021/ja01193a005. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 261–274 https://doi.org/10.26493/1855-3974.2248.d3f (Also available at http://amc-journal.eu) A double Sylvester determinant* Darij Grinberg † Drexel University, Korman Center, 15 S 33rd Street, Philadelphia PA, 19104, USA Received 7 February 2020, accepted 22 August 2020, published online 18 November 2021 Abstract Given two (n+ 1) × (n+ 1)-matrices A and B over a commutative ring, and some k ∈ {0, 1, . . . , n}, we consider the ( n k ) × ( n k ) -matrix W whose entries are (k + 1)×(k + 1)- minors of A multiplied by corresponding (k + 1)× (k + 1)-minors of B. Here we require the minors to use the last row and the last column (which is why we obtain an ( n k ) × ( n k ) - matrix, not a ( n+1 k+1 ) × ( n+1 k+1 ) -matrix). We prove that the determinant detW is a multiple of detA if the (n+ 1, n+ 1)-th entry of B is 0. Furthermore, if the (n+ 1, n+ 1)-th entries of both A and B are 0, then detW is a multiple of (detA) (detB). This extends a previous result of Olver and the author. Keywords: Determinant, compound matrix, Sylvester’s determinant, polynomials. Math. Subj. Class. (2020): 15A15, 11C20 1 Introduction Let n and k be nonnegative integers, and let A = (ai,j)1≤i≤n+1, 1≤j≤n+1 be an (n+ 1)× (n+ 1)-matrix over some commutative ring. Let Pk be the set of all k-element subsets of {1, 2, . . . , n}. For any such subset K ∈ Pk, let K+ denote the subset K ∪ {n+ 1} of {1, 2, . . . , n+ 1}. If U and V are two subsets of {1, 2, . . . , n+ 1}, then subVU A shall denote the |U |×|V |-submatrix of A containing only the entries au,v with u ∈ U and v ∈ V. Let WA be the Pk × Pk-matrix1 whose (I, J)-th entry (for all I ∈ Pk and J ∈ Pk) is det ( subJ+I+ A ) . *The author would like to thank Christian Krattenthaler, Peter Olver and Victor Reiner for enlightening dis- cussions, and Peter Olver for the joint work that led to this paper. The SageMath computer algebra system [14] has been used for experimentation leading up to some of the results below. †Homepage: http://www.cip.ifi.lmu.de/˜grinberg/ E-mail address: darijgrinberg@gmail.com (Darij Grinberg) 1This means a matrix whose rows and columns are indexed by the k-element subsets of {1, 2, . . . , n}. If you pick a total order on the set Pk , then you can view such a matrix as an (n k ) × (n k ) -matrix. cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 262 Ars Math. Contemp. 20 (2021) 261–274 (Thus, the entries of WA are all (k + 1) × (k + 1)-minors of A that use the last row and the last column.) A particular case of a celebrated result going back to Sylvester [15] (see [12, §2.7] or [13, Teorema 2.9.1] or [10] for modern proofs) then says that det (WA) = a p n+1,n+1 · (detA) q , where p = ( n− 1 k ) and q = ( n− 1 k − 1 ) . Now, consider a second (n+ 1) × (n+ 1)-matrix B = (bi,j)1≤i≤n+1, 1≤j≤n+1 over the same ring. Let WA,B (later to be just called W ) be the Pk ×Pk-matrix whose (I, J)-th entry (for all I ∈ Pk and J ∈ Pk) is det ( subJ+I+ A ) det ( subJ+I+ B ) . What can be said about det (WA,B)? In general, very little2. However, under some as- sumptions, it splits off factors. Namely, we shall show (Theorem 2.1) that det (WA,B) is a multiple of detA if bn+1,n+1 = 0. We shall then conclude (Theorem 2.2) that if both an+1,n+1 and bn+1,n+1 are 0, then det (WA,B) is a multiple of (detA) (detB). In either case, the quotient (usually a much more complicated polynomial3) remains mysterious; our proofs are indirect and reveal little about it. Our second result generalizes a curious prop- erty of ( n 2 ) × ( n 2 ) -determinants [6, Theorem 10] that arose from the study of the n-body problem (see Example 2.4 for details). 2 The theorems Let us first introduce the standing notations. Let N = {0, 1, 2, . . .}. Let K be a commutative ring. If a and b are two elements of K, then we write a | b when b is a multiple of a (that is, b ∈ Ka). If m ∈ N, then [m] shall mean the set {1, 2, . . . ,m}. Fix an n ∈ N. If K is any subset of [n], then K+ shall mean the subset K ∪ {n+ 1} of [n+ 1]. Fix k ∈ {0, 1, . . . , n}. Let Pk be the set of all k-element subsets of [n]. This is a finite set; thus, any Pk × Pk-matrix (i.e., any matrix whose rows and columns are indexed by k-element subsets of [n]) has a well-defined determinant4. Such matrices appear frequently in classical determinant theory (see, e.g., the “k-th compound determinants” in [11] and in [12, §2.6], as well as the related “Generalized Sylvester’s identity” in [12, §2.7] and [13, Teorema 2.9.1] and [10]). If A ∈ Ku×v is a u× v-matrix, and if I ⊆ [u] and J ⊆ [v], then subJI A shall mean the submatrix of A obtained by removing all rows whose indices are not in I and removing all columns whose indices are not in J . (Rigorously speaking, if A = (ai,j)1≤i≤u, 1≤j≤v and I = {i1 < i2 < · · · < ip} and J = {j1 < j2 < · · · < jq}, then subJI A is defined to be the matrix (aix,jy )1≤x≤p, 1≤y≤q .) When |I| = |J |, then the submatrix sub J I A is square; its determinant det(subJI A) is called a minor of A. 2For example, if n = 3 and k = 2, then det ( WA,B ) is an irreducible polynomial in the (altogether 2 (n+ 1)2 = 32) variables ai,j and bi,j with 110268 monomials. 3Again, irreducible in the case when n = 3 and k = 2. 4Here, we are using the concepts of P × P -matrices (where P is a finite set) and their determinants. Both of these concepts are folklore; a brief introduction can be found in [5, §1]. D. Grinberg: A double Sylvester determinant 263 Our main two results are the following: Theorem 2.1. Let A = (ai,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) and B = (bi,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) be such that bn+1,n+1 = 0. Let W be the Pk × Pk-matrix whose (I, J)-th entry (for all I ∈ Pk and J ∈ Pk) is det ( subJ+I+ A ) det ( subJ+I+ B ) . Then detA | detW . Theorem 2.2. Let A = (ai,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) and B = (bi,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) be such that an+1,n+1 = 0 and bn+1,n+1 = 0. Define the Pk × Pk-matrix W as in Theorem 2.1. Then (detA) (detB) | detW . Example 2.3. For this example, set k = 1. Then Pk = P1 = {{1} , {2} , . . . , {n}}. Thus, the map [n] → Pk, i 7→ {i} is a bijection. Use this bijection to identify the elements 1, 2, . . . , n of [n] with the elements {1} , {2} , . . . , {n} of Pk. Thus, the Pk×Pk-matrix W in Theorem 2.1 becomes the n×n- matrix( det ( sub {j}+ {i}+ A )︸ ︷︷ ︸ =ai,jan+1,n+1 −ai,n+1an+1,j det ( sub {j}+ {i}+ B )︸ ︷︷ ︸ =bi,jbn+1,n+1 −bi,n+1bn+1,j ) 1≤i≤n, 1≤j≤n = ( (ai,jan+1,n+1 − ai,n+1an+1,j) (bi,j bn+1,n+1︸ ︷︷ ︸ =0 −bi,n+1bn+1,j) ) 1≤i≤n, 1≤j≤n = ((ai,jan+1,n+1 − ai,n+1an+1,j)(−bi,n+1bn+1,j))1≤i≤n, 1≤j≤n. This is the matrix obtained from (ai,jan+1,n+1 − ai,n+1an+1,j)1≤i≤n, 1≤j≤n by multiply- ing the i-th row with −bi,n+1 for all i ∈ [n] and multiplying the j-th column with bn+1,j for all j ∈ [n]. Thus, the claim of Theorem 2.1 follows from the classical fact that det ( (ai,jan+1,n+1 − ai,n+1an+1,j)1≤i≤n, 1≤j≤n ) = an−1n+1,n+1 · detA. This fact is known as Chio pivotal condensation (see, e.g., [7, Theorem 0.1]), and is a particular case of Sylvester’s identity ([12, §2.7]). 264 Ars Math. Contemp. 20 (2021) 261–274 Example 2.4. For this example, set k = 2, and consider the situation of Theorem 2.1 again. Then Pk = P2 = {{i, j} | 1 ≤ i < j ≤ n}. If {i, j} ∈ P2 and {k, l} ∈ P2 satisfy i < j and k < l, then the ({i, j} , {k, l})-th entry of W is det ( sub {k,l}+ {i,j}+ A ) det ( sub {k,l}+ {i,j}+ B ) = det  ai,k ai,l ai,n+1aj,k aj,l aj,n+1 an+1,k an+1,l an+1,n+1  det  bi,k bi,l bi,n+1bj,k bj,l bj,n+1 bn+1,k bn+1,l 0  . Note that bn+1,n+1 = 0. If we furthermore assume that an+1,n+1 = 0, and an+1,i = ai,n+1 = 1 for all i ∈ [n] , and bn+1,i = bi,n+1 = 1 for all i ∈ [n] , then this entry rewrites as det ai,k ai,l 1aj,k aj,l 1 1 1 0  det bi,k bi,l 1bj,k bj,l 1 1 1 0  = (aj,k + ai,l − ai,k − aj,l) (bj,k + bi,l − bi,k − bj,l) . Hence, [6, Theorem 10] can be obtained from Theorem 2.2 by setting k = 2 and A = CS and B = CT (and observing that the matrix W then equals to WS,T ). 3 The proofs Our proofs of Theorem 2.1 and Theorem 2.2 will rely on some basic commutative algebra: the notion of a unique factorization domain (“UFD”); the concepts of coprime, prime and irreducible elements; the localization of a commutative ring at a multiplicative subset. This all appears in most textbooks on abstract algebra; for example, [8, Sections VIII.4 and VIII.10] is a good reference5. The content of a polynomial p over a UFD is defined to be the greatest common divisor of the coefficients of p. For example, the polynomial 4x2 + 6y2 ∈ Z [x, y] has content gcd (4, 6) = 2. (Of course, in a general UFD, the greatest common divisor is defined only up to multiplication by a unit.) The following known facts are crucial to us: Proposition 3.1. A polynomial ring over Z in finitely many indeterminates is always a UFD. □ Proposition 3.1 appears, e.g., in [8, Remark after Corollary 8.21]. For a constructive proof of Proposition 3.1, we refer to [9, Chapter IV, Theorems 4.8 and 4.9] or to [2, Es- say 1.4, Corollary of Theorem 1 and Corollary 1 of Theorem 2]. Proposition 3.2. Let p be an irreducible element of a UFD K. Then the quotient ring K/ (p) is an integral domain. 5We call “multiplicative subset” what Knapp (in [8, Section VIII.10]) calls a “multiplicative system”. D. Grinberg: A double Sylvester determinant 265 Proof of Proposition 3.2. First of all, we recall that any irreducible element of a UFD is prime (indeed, this follows from [8, Proposition 8.13]). Thus, the element p of K is prime. Hence, [8, Proposition 8.14] shows that the ideal (p) of K is prime. Therefore, the quotient ring K/ (p) is an integral domain. This proves Proposition 3.2. We shall furthermore use the following properties of contents (whose proofs are easy): Proposition 3.3. Let U be a UFD. Let p ∈ U [x1, x2, . . . , xm] be a polynomial over U. Assume that the content of p is 1. Also assume that p is irreducible when considered as a polynomial in F [x1, x2, . . . , xm], where F is the field of fractions of U. Then p is also irreducible when considered as a polynomial in U [x1, x2, . . . , xm]. Proposition 3.4. Let U be a UFD. Let p, q ∈ U [x1, x2, . . . , xm] be two polynomials over U. Assume that both p and q have content 1, and assume furthermore that p and q don’t have any indeterminates in common (i.e., there is no i ∈ [m] such that degxi p > 0 and degxi q > 0). Then p and q are coprime. The next simple fact states that for any positive integer n, the determinant of the “generic n × n-matrix” (i.e., of the n × n-matrix whose n2 entries are distinct indeter- minates in a polynomial ring over Z) is irreducible as a polynomial over Z: Corollary 3.5. Let n be a positive integer. Let G be the multivariate polynomial ring Z [ ai,j | (i, j) ∈ [n]2 ] . Let A ∈ Gn×n be the n × n-matrix (ai,j)1≤i≤n, 1≤j≤n. Then the element detA of G is irreducible. Proof of Corollary 3.5. A well-known fact (e.g., [1, Lemma 5.12]) shows that detA is irreducible as an element of Q [ ai,j | (i, j) ∈ [n]2 ] . This yields (using Proposition 3.3) that detA is irreducible as an element of Z [ ai,j | (i, j) ∈ [n]2 ] as well, since the polynomial detA has content 1. This proves Corollary 3.5. An element a of a commutative ring A is said to be regular if every b ∈ A satisfying ab = 0 must satisfy b = 0. (Regular elements are also known as non-zero-divisors.) In a polynomial ring, each indeterminate is regular; hence, each monomial (without coefficient) is regular (since any product of two regular elements is regular). We recall a few standard concepts from commutative algebra. Let K be a commutative ring. A multiplicative subset of K means a subset S of K that contains the unity 1K of K and has the property that every a, b ∈ S satisfy ab ∈ S. If S is a multiplicative subset of K, then the localization of K at S is defined as follows: Let ∼ be the binary relation on the set K× S defined by ((r, s) ∼ (r′, s′)) ⇐⇒ (t (rs′ − sr′) = 0 for some t ∈ S) . Then it is easy to see that ∼ is an equivalence relation. The set L of its equivalence classes [(r, s)] can be equipped with a ring structure via the rules [(r, s)] + [(r′, s′)] = [(rs′ + sr′, ss′)] and [(r, s)] · [(r′, s′)] = [(rr′, ss′)] (with zero element [(0, 1)] and unity [(1, 1)]). The resulting ring L is commutative, and is known as the localization of K at S. (This generalizes the construction of Q from Z known from high school.) The element [(r, s)] of L is denoted by rs . There is a canonical ring homomorphism from K to L that sends each r ∈ K to [(r, 1)] = r1 ∈ L. 266 Ars Math. Contemp. 20 (2021) 261–274 When all elements of the multiplicative subset S are regular, the statement “t(rs′ − sr′) = 0 for some t ∈ S” in the definition of the relation ∼ can be rewritten in the equivalent (but much simpler) form “rs′ = sr′” (which is even more reminiscent of the construction of Q). The following fact is easy to see: Proposition 3.6. Let K be a commutative ring. Let S be a multiplicative subset of K such that all elements of S are regular. Let L be the localization of the ring K at S. Then: (a) The canonical ring homomorphism from K to L is injective. We shall thus consider it as an embedding. (b) If K is an integral domain, then L is an integral domain. (c) Let a and b be two elements of K. Then we have the following logical equivalence: (a | b in L) ⇐⇒ (a | sb in K for some s ∈ S) . Matrices over arbitrary commutative rings can behave a lot less predictably than matri- ces over fields. However, matrices over integral domains still show a lot of the latter good behavior, such as the following: Proposition 3.7. Let P be a finite set. Let M be an integral domain. Let W ∈ MP×P be a P × P -matrix over M. Let u ∈ MP be a vector such that u ̸= 0 and Wu = 0. Here, u is considered as a “column vector”, so that Wu is defined by Wu = ∑ q∈P wp,quq  p∈P , where W = (wp,q)(p,q)∈P×P and u = (up)p∈P . Then detW = 0. Proof of Proposition 3.7. Let m = |P |. Then we can view the P × P -matrix W as an m × m-matrix (by “numerical reindexing”, as explained in [5, §1]), and we can view the vector u as a column vector of size m. Let us do this from here on. Let F be the quotient field of the integral domain M. Thus, there is a canonical embed- ding of M into F. Hence, we can view the matrix W ∈ Mm×m as a matrix over F, and we can view the vector u ∈ Mm as a vector over F. Let us do so from here on. We are now in the realm of classical linear algebra over fields: The vector u ∈ Fm is nonzero (since u ̸= 0) and belongs to the kernel of the m × m-matrix W ∈ Fm×m (since Wu = 0). Hence, the kernel of the matrix W is nontrivial. In other words, this matrix W is singular. Thus, detW = 0 by a classical fact of linear algebra. This proves Proposition 3.7. Let us next recall an identity for determinants (a version of the Cauchy–Binet formula): Lemma 3.8. Let n ∈ N, m ∈ N and p ∈ N. Let A ∈ Kn×p be an n × p-matrix. Let B ∈ Kp×m be a p×m-matrix. Let k ∈ N. Let P be a subset of [n] such that |P | = k. Let Q be a subset of [m] such that |Q| = k. Then det ( subQP (AB) ) = ∑ R⊆[p]; |R|=k det ( subRP A ) · det ( subQR B ) . D. Grinberg: A double Sylvester determinant 267 Lemma 3.8 is [4, Corollary 7.251] (except that we are using the notation subJI C for what is called subw(J)w(I) C in [4]). It also appears in [3, Chapter I, (19)] (where it is stated using p-tuples instead of subsets). The next lemma is just a particular case of Theorem 2.1, but it is a helpful stepping stone on the way to proving the latter theorem: Lemma 3.9. Let A = (ai,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) and B = (bi,j)1≤i≤n+1, 1≤j≤n+1 ∈ K (n+1)×(n+1) be such that bn+1,n+1 = 0. Assume further that an+1,j = 0 for all j ∈ [n] . (3.1) Define the Pk × Pk-matrix W as in Theorem 2.1. Then detA | detW . The following proof is inspired by [6, proof of Theorem 10]. Proof of Lemma 3.9. We WLOG assume that K is the polynomial ring over Z in n2 + (n+ 1) + ((n+ 1) 2 − 1) indeterminates ai,j for all i ∈ [n] and j ∈ [n] ; ai,n+1 for all i ∈ [n+ 1] ; bi,j for all i ∈ [n+ 1] and j ∈ [n+ 1] except for bn+1,n+1. And, of course, we assume that the entries of A and B that are not zero by assumption are these indeterminates.6 The ring K is a UFD (by Proposition 3.1). We WLOG assume that n > 0 (otherwise, the result follows from detW = det ( 0 ) = 0). The set Pk is nonempty (since k ∈ {0, 1, . . . , n}); thus, |Pk| ≥ 1. Let A be the n × n-matrix (ai,j)1≤i≤n, 1≤j≤n ∈ K n×n. Then, because of (3.1), we have detA = an+1,n+1 · detA (3.2) (by [4, Theorem 6.43], applied to n+ 1 instead of n). The matrix A is a completely generic n× n-matrix (i.e., its entries are distinct indeter- minates); thus, its determinant detA is an irreducible polynomial in the polynomial ring Z [ ai,j | (i, j) ∈ [n]2 ] (by Corollary 3.5). Hence, detA also is an irreducible polynomial in the ring K (since K differs from Z [ ai,j | (i, j) ∈ [n]2 ] only in having more variables, which clearly cannot contribute any factors to detA). Thus, Proposition 3.2 (applied to p = detA) shows that the quotient ring K/(detA) is an integral domain. Let M be the quotient ring K/(detA). Then M is an integral domain (since K/(detA) is an integral domain). All monomials in the variables bi,j (with (i, j) ̸= (n+ 1, n+ 1)) are nonzero in M. Likewise, an+1,n+1 ̸= 0 in M. 6These assumptions are legitimate, because if we can prove Lemma 3.9 under these assumptions, then the universal property of polynomial rings shows that Lemma 3.9 holds in the general case. 268 Ars Math. Contemp. 20 (2021) 261–274 Let w be the element ∏ j∈[n] bn+1,j ∈ M. (Strictly speaking, we mean the canonical projection of ∏ j∈[n] bn+1,j ∈ K onto the quotient ring M.) Then, w is a nonzero element of the integral domain M (since bn+1,j ̸= 0 in M for all j ∈ [n]). For each i ∈ [n], we define zi ∈ M by zi = ∏ j∈[n]; j ̸=i bn+1,j (projected onto M). This is a nonzero element of M. In M, we have bn+1,izi = bn+1 ∏ j∈[n]; j ̸=i bn+1,j = ∏ j∈[n] bn+1,j = w (3.3) for all i ∈ [n]. We need another piece of notation: If M is a p × q-matrix, and if u ∈ [p] and v ∈ [q], then M∼u,∼v denotes the (p− 1)× (q − 1)-matrix obtained from M by removing the u-th row and the v-th column. The matrix A∼1,∼(n+1) has determinant 0 (because (3.1) shows that its last row consists of zeroes). In other words, det ( A∼1,∼(n+1) ) = 0. Also, due to (3.1), we see that each i ∈ [n] satisfies det (A∼1,∼i) = an+1,n+1 · det ( A∼1,∼i ) (3.4) (by [4, Theorem 6.43], applied to A∼1,∼i instead of A), because the last row of the matrix A∼1,∼i is (0, 0, . . . , 0, an+1,n+1). For each i ∈ [n+ 1], we define an element ui ∈ M by ui = { zi (−1)i det (A∼1,∼i) , if i ∈ [n] ; 1, if i = n+ 1. Claim 1. All these n+ 1 elements u1, u2, . . . , un+1 of M are nonzero. Proof of Claim 1. Let i ∈ [n]. Then, det ( A∼1,∼i ) ̸= 0 in M because det ( A∼1,∼i ) is a polynomial of smaller degree than detA, and thus is not a multiple of detA. Now, ui = zi (−1)i =an+1,n+1·det(A∼1,∼i) (by (3.4))︷ ︸︸ ︷ det (A∼1,∼i) = zi︸︷︷︸ ̸=0 in M ̸=0 in M︷ ︸︸ ︷ (−1)i an+1,n+1︸ ︷︷ ︸ ̸=0 in M · ̸=0 in M︷ ︸︸ ︷ det ( A∼1,∼i ) ̸= 0 in M (since M is an integral domain). Thus, u1, u2, . . . , un are nonzero. Moreover, un+1 is nonzero (since un+1 = 1). Thus, we are done. Let u = (uJ)J∈Pk ∈ M Pk be the vector defined by uJ = ∏ j∈J uj . D. Grinberg: A double Sylvester determinant 269 Then the entries of the vector u are nonzero (because they are products of the nonzero elements u1, u2, . . . , un+1 of the integral domain M). Since the vector u has at least one entry (because |Pk| ≥ 1), we thus conclude that u ̸= 0. Let ∆ be the diagonal matrix ∆ = diag (u1, u2, . . . , un+1) ∈ M(n+1)×(n+1). Let x ∈ Mn+1 be the column vector defined by x = ( (−1)1 det (A∼1,∼1) , (−1)2 det (A∼1,∼2) , . . . , (−1)n+1 det(A∼1,∼(n+1)) )T . Let (e1, e2, . . . , en+1) be the standard basis of the free M-module Mn+1. Thus, for any (n+ 1)× (n+ 1)-matrix C ∈ M(n+1)×(n+1) and any j ∈ {1, 2, . . . , n+ 1}, we have (the j-th column of the matrix C) = Cej . (3.5) Now, using Laplace expansion, it is easy to see that Ax = − detA · e1. (3.6) To prove Equation (3.6), consider the adjugate adjA of the matrix A. A standard fact ([4, Theorem 6.100]) says that A ·adjA = detA ·In+1. But the definition of adjA reveals that the first column of the matrix adjA is −x. Hence, the first column of the matrix A · adjA is A · (−x) = −Ax. On the other hand, the first column of the matrix A ·adjA is detA ·e1 (since A · adjA = detA · In+1). Comparing the preceding two sentences, we conclude that −Ax = detA · e1, so that Ax = −detA · e1. This proves Equation (3.6). Also, Equation (3.5) (applied to C = BT and j = n+ 1) yields BT en+1 = ( the (n+ 1)-st column of the matrix BT ) = (bn+1,1, bn+1,2, . . . , bn+1,n+1) T . Hence, ∆BT en+1 = ∆(bn+1,1, bn+1,2, . . . , bn+1,n+1) T = (u1bn+1,1, u2bn+1,2, . . . , un+1bn+1,n+1) T (3.7) (since ∆ = diag (u1, u2, . . . , un+1)). Claim 2. We have uibn+1,i = w · (−1)i det (A∼1,∼i) for each i ∈ [n+ 1] . (3.8) Proof of Claim 2. Let i ∈ [n+ 1]. If i = n+ 1, then both sides of (3.8) are zero (because bn+1,n+1 = 0 and det ( A∼1,∼(n+1) ) = 0). If i ̸= n+1, then i ∈ [n] and thus the definition of ui yields ui = zi(−1)i det(A∼1,∼i). Hence, uibn+1,i = zi (−1)i det (A∼1,∼i) bn+1,i = bn+1,izi︸ ︷︷ ︸ =w (by (3.3)) (−1)i det (A∼1,∼i) = w · (−1)i det (A∼1,∼i) . Hence, Equation (3.8) is proven in both cases. 270 Ars Math. Contemp. 20 (2021) 261–274 Now, (3.7) becomes ∆BT en+1 = (u1bn+1,1, u2bn+1,2, . . . , un+1bn+1,n+1) T = ( w · (−1)1 det (A∼1,∼1) , w · (−1)2 det (A∼1,∼2) , . . . , w · (−1)n+1 det ( A∼1,∼(n+1) ) )T (by (3.8)) = w · ( (−1)1 det (A∼1,∼1) , (−1)2 det (A∼1,∼2) , . . . , (−1)n+1 det ( A∼1,∼(n+1) ))T ︸ ︷︷ ︸ =x (by the definition of x) = wx. Hence, A∆BT en+1 = Awx = w · =− detA·e1 (by (3.6))︷︸︸︷ Ax = −w · detA︸ ︷︷ ︸ =an+1,n+1·detA (by (3.2)) · e1 = −w · an+1,n+1 · detA︸ ︷︷ ︸ =0 (since we are in M) · e1 = 0. In other words, the (n+ 1)-st column of the matrix A∆BT is 0 (since the (n+ 1)-st col- umn of the matrix A∆BT is A∆BT en+1 (by (3.5), applied to C = A∆BT and j = n+1)). Now, fix I ∈ Pk. Then, the last column of the matrix subI+I+(A∆BT ) is 0 (because this column is a piece of the (n+ 1)-st column of the matrix A∆BT , but as we have just shown the latter column is 0). Thus, det ( subI+I+(A∆B T ) ) = 0. But Lemma 3.8 (applied to M, n + 1, n + 1, n + 1, ∆BT , k + 1, I+ and I+ instead of K, n, m, p, B, k, P and Q) yields det ( subI+I+(A∆B T ) ) = ∑ R⊆[n+1]; |R|=k+1 det ( subRI+ A ) det ( subI+R (∆B T ) ) . Comparing this with det ( subI+I+(A∆B T ) ) = 0, we obtain 0 = ∑ R⊆[n+1]; |R|=k+1 det ( subRI+ A ) det ( subI+R (∆B T ) ) . In the sum on the right hand side, all addends for which n + 1 /∈ R are zero (because if R ⊆ [n+ 1] satisfies |R| = k + 1 and n+ 1 /∈ R, then the last row of the matrix subRI+ A consists of zeroes (by (3.1), since n + 1 /∈ R but n + 1 ∈ I+), and therefore we have det ( subRI+ A ) = 0), and thus can be discarded. Hence, we are left with 0 = ∑ R⊆[n+1]; |R|=k+1; n+1∈R det ( subRI+ A ) det ( subI+R (∆B T ) ) . D. Grinberg: A double Sylvester determinant 271 But the subsets R of [n+ 1] satisfying |R| = k + 1 and n+ 1 ∈ R can be parametrized as J+ with J ranging over Pk. Hence, this rewrites further as 0 = ∑ J∈Pk det ( subJ+I+ A ) det ( subI+J+(∆B T ) ) . It is easily seen that det ( subI+J+(∆B T ) ) = det ( subJ+I+ B ) uJ for each J ∈ Pk (indeed, recall the definition of ∆ and the fact that un+1 = 1 and that det ( CT ) = detC for each square matrix C). Thus, the above equality simplifies to 0 = ∑ J∈Pk det ( subJ+I+ A ) det ( subJ+I+ B ) uJ . Now, forget that we fixed I . We thus have proven that 0 = ∑ J∈Pk det ( subJ+I+ A ) det ( subJ+I+ B ) uJ (3.9) for each I ∈ Pk. This rewrites as Wu = 0 (indeed, the left hand side of (3.9) is the I-th entry of the zero vector 0, whereas the right hand side of (3.9) is the I-th entry of Wu). Now, consider the matrix W as a matrix in MPk×Pk . Then, Proposition 3.7 (applied to P = Pk) yields detW = 0 in M (since u ̸= 0 and Wu = 0). In view of the definition of M, this rewrites as detA | detW in K. Let us consider the matrix W again as a matrix over K. Each entry of W has the form det ( subJ+I+ A ) det ( subJ+I+ B ) for some I, J ∈ Pk. Claim 3. det ( subJ+I+ A ) is a multiple of an+1,n+1 for all I, J ∈ Pk. Proof of Claim 3. Let I, J ∈ Pk. Then, the equality (3.1) shows that the last row of the matrix subJ+I+ A is (0, 0, . . . , 0, an+1,n+1). Hence, an application of [4, Theorem 6.43] shows that det ( subJ+I+ A ) = an+1,n+1 det ( subJI A ) . Thus, det ( subJ+I+ A ) is a multiple of an+1,n+1. By Claim 3, all entries of W are multiples of an+1,n+1. Hence, the determinant of W is a multiple of (an+1,n+1) |Pk|, thus a multiple of an+1,n+1 (since |Pk| ≥ 1). In other words, an+1,n+1 | detW in K. Recall that K is a UFD. Also, the two polynomials an+1,n+1 and detA in K both have content 1, and don’t have any indeterminates in common; thus, these two polynomials are coprime (by Proposition 3.4). Hence, any polynomial in K that is divisible by both an+1,n+1 and detA must be divisible by the product an+1,n+1 · detA as well. Thus, from an+1,n+1 | detW and detA | detW , we obtain an+1,n+1 · detA | detW . In view of (3.2), this rewrites as detA | detW . This proves Lemma 3.9. We shall now derive Theorem 2.2 from Lemma 3.9, following the same idea as in [12, §2.7] and [13, Teorema 2.9.1] and [10]: Proof of Theorem 2.1. We WLOG assume that n > 0 (otherwise, the result follows from detW = det ( 0 ) = 0). 272 Ars Math. Contemp. 20 (2021) 261–274 We WLOG assume that K is the polynomial ring over Z in (n+ 1)2 + ((n+ 1)2 − 1) indeterminates ai,j for all i ∈ [n+ 1] and j ∈ [n+ 1] ; bi,j for all i ∈ [n+ 1] and j ∈ [n+ 1] except for bn+1,n+1. And, of course, we assume that the entries of A and B that are not zero by assumption are these indeterminates. Proposition 3.1 shows that the ring K is a UFD (since it is a polynomial ring over Z). Let S be the multiplicative subset { apn+1,n+1 | p ∈ N } of K. Then, all elements of S are regular (since they are monomials in a polynomial ring). Let L be the localization of the commutative ring K at the multiplicative subset S. Then, Proposition 3.6(a) shows that the canonical ring homomorphism from K to L is injective; we shall thus consider it as an embedding. Also, Proposition 3.6(b) shows that L is an integral domain. Claim 1. We claim that detA | detW in L. (3.10) Proof of Claim 1. Consider A, B and W as matrices over L. The entry an+1,n+1 of A is invertible in L (by the construction of L). Hence, we can subtract appropriate scalar multiples7 of the (n+ 1)-st column of A from each other column of A to ensure that all entries of the last row of A become 0, except for an+1,n+1. (Specifically, for each j ∈ [n], we have to subtract aj,n+1/an+1,n+1 times the (n+ 1)-st column of A from the j- th column of A.) All these column transformations preserve the determinant detA, and also preserve the minors det ( subJ+I+ A ) for all I, J ∈ Pk (because when the (n+ 1)-st column of A is subtracted from another column of A, the matrix subJ+I+ A either stays the same or undergoes an analogous column transformation8, which preserves its determinant); thus, they preserve the matrix W . Hence, we can replace A by the result of these transformations. This new matrix A satisfies (3.1). Hence, Lemma 3.9 (applied to L instead of K) yields that detA | detW in L. This proves (3.10). But we must prove that detA | detW in K. Fortunately, this is easy: Since K embeds into L, we can translate our result “detA | detW in L” as “detA | apn+1,n+1 detW in K for an appropriate p ∈ N” (by Proposition 3.6(c), applied to a = detA and b = detW ). Consider this p. Claim 2. The polynomial an+1,n+1 ∈ K is coprime to detA. Proof of Claim 2. The polynomial detA contains the monomial a1,n+1a2,n · · · an+1,1 =∏ i∈[n+1] ai,n+2−i, and thus is not a multiple of an+1,n+1. Hence, it is coprime to an+1,n+1 (since the only non-unit divisor of an+1,n+1 is an+1,n+1 itself, up to scaling by units). So we know that an+1,n+1 is coprime to detA. Hence, its power a p n+1,n+1 is co- prime to detA as well. Hence, we can cancel the apn+1,n+1 from the divisibility detA | apn+1,n+1 detW , and conclude that detA | detW in K. This proves Theorem 2.1. 7The scalars, of course, come from L here. 8Here we are using the fact that n + 1 ∈ J+ (so that the matrix subJ+I+ A contains part of the (n+ 1)-st column of A). D. Grinberg: A double Sylvester determinant 273 Proof of Theorem 2.2. We WLOG assume that K is the polynomial ring over Z in the ((n+ 1) 2 − 1) + ((n+ 1)2 − 1) indeterminates ai,j for all i ∈ [n+ 1] and j ∈ [n+ 1] except for an+1,n+1; bi,j for all i ∈ [n+ 1] and j ∈ [n+ 1] except for bn+1,n+1. And, of course, we assume that the entries of A and B that are not zero by assumption are these indeterminates. The ring K is a UFD (by Proposition 3.1). WLOG assume that n > 0 (otherwise, the result follows from detW = det ( 0 ) = 0). Thus, the monomial a1,n+1a2,n · · · an+1,1 = ∏ i∈[n+1] ai,n+2−i occurs in the polynomial detA with coefficient ±1. Hence, the polynomial detA has content 1. Similarly, the polynomial detB has content 1. Theorem 2.1 yields detA | detW . The same argument yields detB | detW (since the matrices A and B play symmetric roles in the construction of W ). But Proposition 3.4 shows that the polynomials detA and detB in K are coprime (because they have content 1, and don’t have any indeterminates in common). Thus, any polynomial in K that is divisible by both detA and detB must be divisible by the product (detA) (detB) as well. Thus, from detA | detW and detB | detW , we obtain (detA) (detB) | detW . This proves Theorem 2.2. 4 Further questions While Theorems 2.1 and 2.2 are now proven, the field appears far from fully harvested. Three questions readily emerge: Question 4.1. What can be said about detWdetA (in Theorem 2.1) and detW (detA)(detB) (in Theo- rem 2.2)? Are there formulas? Question 4.2. Are there more direct proofs of Theorems 2.1 and 2.2, avoiding the use of polynomial rings and their properties and instead “staying inside K”? Such proofs might help answer the previous question. Question 4.3. The entries of our matrix W were products of minors of two (n+ 1) × (n+ 1)-matrices that each use the last row and the last column. What can be said about products of minors of two (n+m)× (n+m)-matrices that each use the last m rows and the last m columns, where m is an arbitrary positive integer? The “Generalized Sylvester’s identity” in [12, §2.7] answers this for the case of one matrix. It is not quite obvious what the right analogues of the conditions an+1,n+1 = 0 and bn+1,n+1 = 0 are; furthermore, nontrivial examples become even more computationally challenging. ORCID iD Darij Grinberg https://orcid.org/0000-0002-9661-8432 References [1] J. Désarménien, J. P. S. Kung and G.-C. Rota, Invariant theory, Young bitableaux, and combi- natorics, Advances in Math. 27 (1978), 63–92, doi:10.1016/0001-8708(78)90077-4. [2] H. M. Edwards, Essays in Constructive Mathematics, Springer-Verlag, New York, 2005. 274 Ars Math. Contemp. 20 (2021) 261–274 [3] F. R. Gantmacher, The Theory of Matrices, Volume 1, AMS Chelsea Publishing, Providence, RI, 1998. [4] D. Grinberg, Notes on the combinatorial fundamentals of algebra, arXiv:2008.09862v1 [math.CO]. [5] D. Grinberg, MathOverflow post #317105 (answer to “A Putnam problem with a twist”), 2019, https://mathoverflow.net/q/317105. [6] D. Grinberg and P. J. Olver, The n body matrix and its determinant, SIAM J. Appl. Algebra Geom. 3 (2019), 67–86, doi:10.1137/18m1175410. [7] K. Karnik and A. Zhang, Combinatorial proof of Chio Pivotal Condensation, http://www. cip.ifi.lmu.de/˜grinberg/primes2015/kazh-exp.pdf. [8] A. W. Knapp, Basic Algebra, published by A. W. Knapp, digital 2nd edition, 2016, http: //www.math.stonybrook.edu/˜aknapp/download.html. [9] R. Mines, F. Richman and W. Ruitenburg, A Course in Constructive Algebra, Universitext, Springer-Verlag, New York, 1988, doi:10.1007/978-1-4419-8640-5. [10] E. Mohr, Einfacher Beweis des verallgemeinerten Determinantensatzes von Sylvester nebst einer Verschärfung, Math. Nachr. 10 (1953), 257–260, doi:10.1002/mana.19530100502. [11] T. Muir, A Treatise on the Theory of Determinants, Dover Publications, New York, 1960, re- vised and enlarged by William H. Metzler. [12] V. V. Prasolov, Problems and Theorems in Linear Algebra, volume 134 of Translations of Mathematical Monographs, American Mathematical Society, Providence, RI, 1994, doi:10. 1090/mmono/134, translated from the Russian manuscript by D. A. Leı̆tes. [13] V. V. Prasolov, Zadachi i teoremy linejnoj algebry, Izdatelstvo MZNMO, Moskva, 2nd edition, 2015. [14] The Sage Developers, SageMath, the Sage Mathematics Software System (Version 7.6), 2017, https://www.sagemath.org/. [15] J. J. Sylvester, On the relation between the minor determinants of linearly equivalent quadratic functions, Philos. Magazine 1 (1851), 295–305, doi:10.1080/14786445108646735. ISSN 1855-3966 (printed edn.), ISSN 1855-3974 (electronic edn.) ARS MATHEMATICA CONTEMPORANEA 20 (2021) 275–288 https://doi.org/10.26493/1855-3974.2229.5f1 (Also available at http://amc-journal.eu) Boundary-type sets of strong product of directed graphs* Bijo S. Anand Department of Mathematics, Sree Narayana College, Punalur, Kollam, India-691305 Manoj Changat Department of Futures Studies, University of Kerala, Trivandrum, India-695581 Prasanth G. Narasimha-Shenoi † , Mary Shalet Thottungal Joseph ‡ Department of Mathematics, Government College Chittur, Palakkad, India-678104 Received 26 January 2020, accepted 8 March 2021, published online 18 November 2021 Abstract Let D = (V,E) be a strongly connected digraph and let u and v be two vertices in D. The maximum distance md(u, v) is defined as md(u, v) = max{d⃗(u, v), d⃗(v, u)}, where d⃗(u, v) denotes the length of a shortest directed u-v path in D. This is a metric. The boundary, contour, eccentricity and periphery sets of a strongly connected digraph D with respect to this metric have been defined. The boundary-type sets of the strong product of two digraphs is investigated in this article. Keywords: Maximum distance, boundary-type sets, strongly connected digraph, strong product. Math. Subj. Class. (2020): 05C12, 05C20, 05C76 *The authors thank the anonymous referees for their valuable suggestions, which helped improve this article. †Supported by Science and Engineering Research Board, a statutory body of Government of India under their Extra Mural Research Funding No. EMR/2015/002183 and MATRICS Scheme No. MTR/2018/000012. Supported by Kerala State Council for Science Technology and Environment of Government of Kerala under their SARD project grant Council(P) No. 436/2014/KSCSTE. ‡Corresponding author. Supported by Science and Engineering Research Board, a statutory body of Gov- ernment of India under their Extra Mural Research Funding No. EMR/2015/002183. Supported by Kerala State Council for Science Technology and Environment of Government of Kerala under their SARD project grant Coun- cil(P) No. 436/2014/KSCSTE. E-mail addresses: bijos_anand@yahoo.com (Bijo S. Anand), mchangat@keralauniversity.ac.in (Manoj Changat), prasanthgns@gmail.com (Prasanth G. Narasimha-Shenoi), mary_shallet@yahoo.co.in (Mary Shalet Thottungal Joseph) cb This work is licensed under https://creativecommons.org/licenses/by/4.0/ 276 Ars Math. Contemp. 20 (2021) 275–288 1 Introduction Directed graphs or in short digraphs have immense applications in almost all areas of sci- ence and even in sociology. A directed network is a network in which each edge has a direction, pointing from one vertex to another. They can be represented as directed graphs. Road traffic networks are the most frequently met examples of one-way networks. A two-way street is one in which vehicles are allowed to travel in both directions. The advan- tages of a one-way street network over a two-way street pattern are discussed in [14]. But when one-way traffic is introduced in a two-way network, the distance between places in one of the directions may increase. So the problem of designing a network is to minimize the distance between places and the cost of construction. The one-way problem was first studied by Robbins [13]. It finds applications in various fields like computer science, biology, etc. In [2], directed graphs are used to analyze the local properties of internet connectivity. Neurons are connected in intricate communication networks established during development to convey sensory information from peripheral receptors of sensory neurons to the central nervous system and to convey commands from the central nervous system to effector organs [12]. The boundary-type sets of a graph, the boundary, contour, eccentricity, and periph- ery sets of a graph were studied in [5] and [7]. It is very difficult to identify the various boundary-type sets in large networks. So we try to decompose the network into smaller networks and identify the boundary-type sets. The four standard graph products, namely Cartesian, direct, strong, and lexicographic products can be extended to digraphs as well. Marc Hellmuth and Tilen Marc developed a polynomial-time algorithm for determining the prime factor decomposition of strong product of digraphs [11]. The directed distance defined in digraphs is generally not a metric. As we are concerned with the problem of designing the network to minimize the distance between places at a minimum cost, we consider the distance maximum distance or in short, m-distance which is a metric that was introduced by Chartrand and Tian in [8]. It gives the maximum of the directed distances in either direction and is denoted by md(u, v). So minimizing md(u, v) results in minimizing the distance between the nodes in both directions. An undirected graph G can be identified as a symmetric digraph, that is, one for which (u, v) ∈ E(G) if and only if (v, u) ∈ E(G), and the metric md is the usual distance metric in undirected graphs. The boundary-type sets of the Cartesian product of two digraphs were studied in [6]. In this paper, a similar study is conducted for the strong product of digraphs. 2 Preliminaries A directed graph or a digraph D consists of a non-empty finite set V (D) of elements called vertices and a finite set E(D) of ordered pairs of distinct vertices called arcs or edges [1]. We call V (D) the vertex set and E(D) the edge set of D. We write D = (V,E) to denote the digraph D with vertex set V and edge set E. For an edge (u, v), the first vertex u of the ordered pair is the tail of the edge and the second vertex v is the head; together they are the endpoints. This definition of a digraph does not allow loops (edges whose head and tail coincide) or parallel edges (pairs of edges with the same tail and the same head). The underlying graph UD of a digraph D is the simple graph with the vertex set V (D) and the unordered pair (x, y) ∈ E(UD) if and only if either (x, y) ∈ E(D) or (y, x) ∈ E(D). The following concepts are taken from [1]. B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 277 For a vertex v in a digraph D = (V,E), the neighborhoods are defined as follows: N+D (v) = {w ∈ V : (v, w) ∈ E}, N − D (v) = {u ∈ V : (u, v) ∈ E}. The sets N+D (v), N − D (v), and ND(v) = N + D (v)∪N − D (v) are called the out-neighborhood, in-neighborhood, and neighborhood of v. These neighborhoods are called open neighbor- hoods of v. Similarly, we can define closed neighborhoods of v (neighbors including v). The closed neighborhood of v is denoted by ND[v]. That is, ND[v] = ND(v) ∪ {v}. A directed path is a directed graph with V (P ) ̸= ∅ with distinct vertices u1, u2, . . . , uk and edges e1, e2, . . . , ek−1 such that ei is an edge directed from ui to ui+1 for 1 ≤ i ≤ k − 1. In this article, a path will always mean a ‘directed path’. A digraph is strongly connected or strong if, for each ordered pair (u, v) of vertices, there is a path from u to v. A digraph is weakly connected if its underlying graph is connected. A strong component of a digraph D is a maximal induced subdigraph of D which is strong. If D1, D2, . . . , Dt are the strong components of D, then V (D1) ∪ V (D2) ∪ · · · ∪ V (Dt) = V (D) and V (Di) ∩ V (Dj) = ∅ for every i ̸= j. The length of a path is the number of edges in the path. The directed distance d⃗(u, v) between two vertices u, v ∈ V (D) is the length of the shortest directed path from u to v, or infinity if no such path exists. Note that this distance is not a metric, as generally d⃗(u, v) ̸= d⃗(v, u). So in [8], Chartrand and Tian introduced two other distances between the vertices u and v in a strong digraph, namely the maximum distance md(u, v) = max{d⃗(u, v), d⃗(v, u)} and the sum distance sd(u, v) = d⃗(u, v)+d⃗(v, u), both of which are metrics. In this article, we deal with the maximum distance, md . Remark 2.1. md(u, v) is denoted by d(u, v) hereafter. The m-eccentricity of a vertex v, the m-radius and the m-diameter of a digraph D are also defined in [8]. Consistent with our notation d(u, v) for maximum distance be- tween the vertices u and v, we denote them respectively as ecc(v), rad(D), and diam(D). Thus, ecc(v) = maxu∈V (D) d(v, u), rad(D) = minv∈V (D) ecc(v), and diam(D) = maxv∈V (D) ecc(v), where ecc(v) denotes m-eccentricity of v. If a digraph D is strongly connected, then the maximum distance between every pair of vertices is finite, and hence the m-eccentricity of every vertex in D is finite. Otherwise, D has more than one strong component, and the maximum distance between two vertices lying in different strong components of D is infinity. So if D is not strongly connected, then the m-eccentricity of every vertex in D is infinity. 2.1 Definitions of boundary-type sets We define the boundary-type sets of a digraph D with respect to the metric maximum distance. Most of the following definitions are analogous to the definitions in [7]. Let D be a strong digraph and u, v ∈ V (D). The vertex v is said to be a boundary vertex of u if no neighbor of v is further away from u than v. Hereafter, we denote ND(v) and ND[v] by N(v) and N [v], respectively. A vertex v is called a boundary vertex of D if it is the boundary vertex of some vertex u ∈ V (D). Definition 2.2. The boundary ∂(D) of D is the set of all of its boundary vertices ∂(D) = {v ∈ V (D) : ∃u ∈ V (D) such that ∀w ∈ N(v), d(u,w) ≤ d(u, v)}. 278 Ars Math. Contemp. 20 (2021) 275–288 Given u, v ∈ V (D), the vertex v is called an eccentric vertex of u if no vertex in V (D) is further away from u than v; that is, if d(u, v) = ecc(u). A vertex v is called an eccentric vertex of digraph D if it is the eccentric vertex of some vertex u ∈ V (D). Definition 2.3. The eccentricity Ecc(D) of a digraph D is the set of all of its eccentric vertices Ecc(D) = {v ∈ V (D) : ∃u ∈ V (D) such that ecc(u) = d(u, v)}. In a similar way, we can define the eccentricity of any proper subset W of the vertex set V (D): Ecc(W ) = {v ∈ V (D) : ∃u ∈ W such that ecc(u) = d(u, v)}. A vertex v ∈ V (D) is called a peripheral vertex of D if no vertex in V (D) has an eccentricity greater than ecc(v); that is, if the eccentricity of v is equal to the diameter diam(D) of D. Definition 2.4. The periphery Per(D) of a digraph D is the set of all of its peripheral vertices Per(D) = {v ∈ V (D) : ecc(u) ≤ ecc(v),∀u ∈ V (D)} = {v ∈ V (D) : ecc(v) = diam(D)}. A vertex v ∈ V (D) is called a contour vertex of digraph D if no neighbor vertex of v has an eccentricity greater than ecc(v). The following definition is from [5]. Definition 2.5. The contour Ct(D) of a digraph D is the set of all of its contour vertices Ct(D) = {v ∈ V (D) : ecc(u) ≤ ecc(v),∀u ∈ N(v)}. As in the case of undirected graphs [3] we have, 1. Per(D) ⊆ Ct(D) ∩ Ecc(D), 2. Ecc(D) ∪ Ct(D) ⊆ ∂(D). This is because a peripheral vertex is a vertex having the maximum eccentricity in the digraph D and so every peripheral vertex in D is a contour vertex in D as well as the eccentric vertex of a diametrical vertex in D. If v is an eccentric vertex of a vertex u, then v is a boundary vertex of u. Also if v is a contour vertex, then ecc(u) ≤ ecc(v) for all u ∈ N(v). So there exists some vertex w ∈ V (D) such that d(w, u) ≤ d(w, v) for all u ∈ N(v), and hence v is a boundary vertex of w. The open neighborhood N(v) can be replaced by the closed neighborhood N [v] in the definitions of the boundary and the contour sets. This does not affect the definitions and is necessary for proving the relationship between the boundary and the contour sets of the strong product of two digraphs and its factors. B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 279 3 Strong product of directed graphs The strong product D1 ⊠ D2 of two digraphs D1 and D2 with vertex sets V (D1) = {u1, u2, . . . , um} and V (D2) = {v1, v2, . . . , vn} is the digraph having the vertex set V (D1) × V (D2) and with arc set E(D1 ⊠ D2) defined as follows. A vertex (ui, vr) is adjacent to (uj , vs) in D1 ⊠D2 if either 1. (ui, uj) ∈ E(D1), vr = vs, or 2. ui = uj , (vr, vs) ∈ E(D2), or 3. (ui, uj) ∈ E(D1), (vr, vs) ∈ E(D2). The strong product of digraphs is commutative [10]. The distance between two vertices (g, h) and (g′, h′) in the strong product G ⊠ H of two graphs G and H is given in [9] as follows: dG⊠H((g, h), (g ′, h′)) = max{dG(g, g′), dH(h, h′)}. So in the case of two digraphs D1 and D2, it follows that the directed distance d⃗D1⊠D2((ui, vr), (uj , vs)) = max{d⃗D1(ui, uj), d⃗D2(vr, vs)}. Lemma 3.1. Let D1 and D2 be two strongly connected digraphs. Then dD1⊠D2((ui, vr), (uj , vs)) = max{dD1(ui, uj), dD2(vr, vs)}, eccD1⊠D2(ui, vr) = max{eccD1(ui), eccD2(vr)}. Proof. dD1⊠D2((ui, vr), (uj , vs)) = max{d⃗D1⊠D2((ui, vr), (uj , vs)), d⃗D1⊠D2((uj , vs), (ui, vr))} = max{max{d⃗D1(ui, uj), d⃗D2(vr, vs)},max{d⃗D1(uj , ui), d⃗D2(vs, vr)}} = max{max{d⃗D1(ui, uj), d⃗D2(vr, vs), d⃗D1(uj , ui), d⃗D2(vs, vr)}} = max{max{d⃗D1(ui, uj), d⃗D1(uj , ui)},max{d⃗D2(vr, vs), d⃗D2(vs, vr)}} = max{dD1(ui, uj), dD2(vr, vs)}. Hence it follows that eccD1⊠D2(ui, vr) = max{dD1⊠D2((ui, vr), (uj , vs)) : (uj , vs) ∈ V (D1 ⊠D2)} = max{max{dD1(ui, uj), dD2(vr, vs)} : uj ∈ V (D1), vs ∈ V (D2)} = max{max{dD1(ui, uj) : uj ∈ V (D1)},max{dD2(vr, vs) : vs ∈ V (D2)}} = max{eccD1(ui), eccD2(vr)}. Corollary 3.2. Let D1 and D2 be two strongly connected digraphs. Then rad(D1 ⊠D2) = max{rad(D1), rad(D2)}, diam(D1 ⊠D2) = max{diam(D1),diam(D2)}. 280 Ars Math. Contemp. 20 (2021) 275–288 Proof. rad(D1 ⊠D2) = min (ui,vr)∈V (D1⊠D2) ecc(ui, vr) = min ui∈V (D1) vr∈V (D2) max{eccD1(ui), eccD2(vr)} = max{ min ui∈V (D1) ecc(ui), min vr∈V (D2) ecc(vr)} = max{rad(D1), rad(D2)}, diam(D1 ⊠D2) = max (ui,vr)∈V (D1⊠D2) ecc(ui, vr) = max ui∈V (D1) vr∈V (D2) max{eccD1(ui), eccD2(vr)} = max{ max ui∈V (D1) ecc(ui), max vr∈V (D2) ecc(vr)} = max{diam(D1),diam(D2)}. The strong product of two directed graphs is strongly connected if and only if both the digraphs are strongly connected [9]. Also if G and H are two undirected graphs, NG⊠H [(g, h)] = NG[g]×NH [h] [9]. Since the neigbors of a vertex in a directed graph are exactly its neighbors in the underlying graph, it follows that ND1⊠D2 [(ui, vr)] = NG⊠H [(ui, vr)] = NG[ui]×NH [vr] = ND1 [ui]×ND2 [vr], where G and H are the underlying graphs of D1 and D2, respectively. In [4], Cáceres et al. presented a description of the boundary-type sets of two undirected graphs and the description of the boundary is as follows. For two graphs G and H , ∂(G ⊠H) = (∂(G) × V (H)) ∪ (V (G) × ∂(H)). But this result does not hold in the case of directed graphs. Consider the strong product, D1 ⊠ D2 of the digraphs D1 and D2 in Figure 1. The eccentricity of each vertex is displayed near the vertex in red color. Per(D1) = Ecc(D1) = Ct(D1) = {u1, u4}, Per(D2) = Ecc(D2) = Ct(D2) = {v1, v2}, and Per(D1 ⊠D2) = Ecc(D1 ⊠D2) = Ct(D1 ⊠D2) = {(u1, v1), (u4, v1), (u1, v2), (u4, v2)}. ∂(D1) = {u1, u4}, ∂(D2) = {v1, v2}, and ∂(D1 ⊠D2) = {(u1, v1), (u4, v1), (u1, v2), (u4, v2)}. The reason for (u2, v1), (u2, v2), (u3, v1), (u3, v2) /∈ ∂(D1 ⊠ D2) is explained after the proof of Theorem 3.3. Now we present the results concerning the boundary-type sets of the strong product of two strongly connected digraphs. In all these results, D1 and D2 can be interchanged due to the commutativity of strong product of digraphs. B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 281 v2 1 v1 1 D2 (u1, v1) 3 (u2, v1) 2 (u3, v1) 2 (u4, v1) 3 (u1, v2) 3 (u2, v2) 2 (u3, v2) 2 (u4, v2) 3 u1 3 u2 2 u3 2 u4 3 D1 Figure 1: D1 ⊠D2. We have, ∂(D1 ⊠D2) ⊆ [∂(D1)× V (D2)] ∪ [V (D1)× ∂(D2)]. To this end, let (ui, vr) ∈ ∂(D1 ⊠ D2). Then there exists a vertex (uj , vs) ∈ V (D1 ⊠ D2) such that d((uj , vs), (ui, vr)) ≥ d((uj , vs), (uk, vq)) for every (uk, vq) ∈ N [(ui, vr)]. This implies, max{d(uj , ui), d(vs, vr)} ≥ max{d(uj , uk), d(vs, vq)} for every uk ∈ N [ui] and for every vq ∈ N [vr]. Hence d(uj , ui) ≥ d(uj , uk) for ev- ery uk ∈ N [ui], or d(vs, vr) ≥ d(vs, vq) for every vq ∈ N [vr]. Thus, ui ∈ ∂(D1) or vr ∈ ∂(D2) or both. That is, if (ui, vr) ∈ ∂(D1 ⊠D2), then at least one of the vertices ui and vr must be a boundary vertex in the corresponding factor graph. Theorem 3.3. Let D1 and D2 be two strongly connected digraphs. Then ∂(D1 ⊠D2) = A1 ∪A2 ∪A3, where A1 = ∂(D1)× ∂(D2), A2 = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ ∂(D1), vr /∈ ∂(D2), and ∃vt ∈ V (D2) such that d(vt, vq) ≤ ecc(ui),∀vq ∈ N [vr]}, A3 = {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ ∂(D1), vr ∈ ∂(D2), and ∃uℓ ∈ V (D1) such that d(uℓ, uk) ≤ ecc(vr),∀uk ∈ N [ui]}. Proof. Suppose that (ui, vr) ∈ ∂(D1 ⊠D2). 282 Ars Math. Contemp. 20 (2021) 275–288 Then there exists a vertex (uj , vs) ∈ V (D1 ⊠ D2) such that d((uj , vs), (ui, vr)) ≥ d((uj , vs), (uk, vq)) for all vertices (uk, vq) ∈ N [(ui, vr)]. Since d((uj , vs), (ui, vr)) = max{d(uj , ui), d(vs, vr)} and d((uj , vs), (uk, vq)) = max{d(uj , uk), d(vs, vq)}, we get max{d(uj , ui), d(vs, vr)} ≥ max{d(uj , uk), d(vs, vq)} for all uk ∈ N [ui], vq ∈ N [vr]. We distinguish four cases: 1. max{d(uj , ui), d(vs, vr)} = d(uj , ui) and d(vs, vr) ≥ d(vs, vq) for all vq ∈ N [vr]; 2. max{d(uj , ui), d(vs, vr)} = d(uj , ui) and d(vs, vr) ≥ d(vs, vq) does not hold for all vq ∈ N [vr]; 3. max{d(uj , ui), d(vs, vr)} = d(vs, vr) and d(uj , ui) ≥ d(uj , uk) for all uk ∈ N [ui]; 4. max{d(uj , ui), d(vs, vr)} = d(vs, vr) and d(uj , ui) ≥ d(uj , uk) does not hold for all uk ∈ N [ui]. In cases 1 and 3, d(uj , ui) ≥ d(uj , uk) for all uk ∈ N [ui] and d(vs, vr) ≥ d(vs, vq) for all vq ∈ N [vr]. So ui ∈ ∂(D1), vr ∈ ∂(D2), and hence (ui, vr) ∈ A1. In case 2, ui ∈ ∂(D1) and vr is not a boundary vertex of vs in D2. If there exists any vertex vt such that vr is a boundary vertex of vt, then we get (ui, vr) ∈ A1. Otherwise, since vr /∈ ∂(D2), for every vertex vt ∈ V (D2), there exists some vertex vq ∈ N [vr] such that d(vt, vr) < d(vt, vq). Hence if (ui, vr) is a boundary vertex of a vertex (uℓ, vt) in D1 ⊠D2, then d((uℓ, vt), (ui, vr)) = max{d(uℓ, ui), d(vt, vr)} = d(uℓ, ui) > d(vt, vr), for otherwise d(uℓ, ui) ≤ d(vt, vr) and so we get d((uℓ, vt), (ui, vr)) = d(vt, vr) < d(vt, vq) = d((uℓ, vt), (ui, vq)), where (ui, vq) ∈ N [(ui, vr)]. Let (uk, vq) ∈ N [(ui, vr)]. Then d((uℓ, vt), (uk, vq)) = max{d(uℓ, uk), d(vt, vq)}. If (ui, vr) is a boundary vertex of (uℓ, vt), then max{d(uℓ, ui), d(vt, vr)} ≥ max{d(uℓ, uk), d(vt, vq)}. So the necessary condition for the vertex (ui, vr) such that ui ∈ ∂(D1) and vr /∈ ∂(D2) to be a boundary vertex of the vertex (uℓ, vt) in D1 ⊠D2 is d(uℓ, ui) ≥ d(vt, vq) for all vq ∈ N [vr]. Since ecc(ui) ≥ d(uℓ, ui) for all uℓ ∈ V (D1), the necessary condition becomes ecc(ui) ≥ d(vt, vq) for all vq ∈ N [vr]. Thus, (ui, vr) ∈ A2. Thus in case 2, (ui, vr) ∈ A1 ∪A2. In case 4, vr ∈ ∂(D2) and ui is not a boundary vertex of uj in D1. As in case 2, it follows that (ui, vr) ∈ A1 ∪A3. Thus in all cases, we get ∂(D1 ⊠D2) ⊆ A1 ∪A2 ∪A3. Conversely, suppose that (ui, vr) ∈ A1 ∪ A2 ∪ A3. First let (ui, vr) ∈ A1. Then ui ∈ ∂(D1) and vr ∈ ∂(D2). So there exists vertices uj ∈ V (D1), vs ∈ V (D2) such that d(uj , ui) ≥ d(uj , uk) for every uk ∈ N [ui], and d(vs, vr) ≥ d(vs, vq) for every vq ∈ N [vr]. Hence in D1 ⊠ D2, d((uj , vs), (ui, vr)) = max{d(uj , ui), d(vs, vr)} ≥ max{d(uj , uk), d(vs, vq)} = d((uj , vs), (uk, vq)) for all vertices (uk, vq) ∈ N [(ui, vr)]. Thus, A1 ⊆ ∂(D1 ⊠D2). Now let (ui, vr) ∈ A2. Then ui ∈ ∂(D1), vr /∈ ∂(D2) and there exists some vertex vt ∈ V (D2) such that d(vt, vq) ≤ ecc(ui), for all vq ∈ N [vr]. Since ui ∈ ∂(D1), there exists at least one vertex uj ∈ V (D1) such that d(uj , ui) ≥ d(uj , uk) for every uk ∈ N [ui]. Of these vertices, let ub be a vertex such that d(ub, ui) = ecc(ui). Hence in D1 ⊠D2, d((ub, vt), (ui, vr)) = max{d(ub, ui), d(vt, vr)} ≥ max{d(ub, uk), d(vt, vq)} = d((ub, vt), (uk, vq)) B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 283 for all (uk, vq) ∈ N [(ui, vr)], since d(vt, vq) ≤ ecc(ui) = d(ub, ui) for all vq ∈ N [vr]. Thus, (ui, vr) is a boundary vertex of (ub, vt) in D1 ⊠D2 and hence A2 ⊆ ∂(D1 ⊠D2). By analogous arguments and since the strong product of digraphs is commutative, it follows that A3 ⊆ ∂(D1 ⊠D2). Hence A1 ∪A2 ∪A3 ⊆ ∂(D1 ⊠D2). Now consider Figure 1. ecc(v1) = ecc(v2) = 1. N [u2] = N [u3] = {u1, u2, u3, u4}, d(u1, u4) = 3, d(u1, u2) = d(u1, u3) = d(u2, u4) = d(u3, u4) = 2, and d(u2, u3) = 1. u2 /∈ ∂(D1) and hence (u2, v1), (u2, v2) /∈ ∂(D1 ⊠ D2), since there is no vertex uℓ ∈ V (D1) such that d(uℓ, uk)) ≤ 1 for all uk ∈ N [u2]. For similar reasons, (u3, v1), (u3, v2) /∈ ∂(D1 ⊠D2). Consider the strong product of two connected undirected graphs. In the case of an undirected graph, the maximum distance between two vertices is the usual distance between the vertices. Also, since we deal with the distance between any two distinct vertices, it doesn’t matter whether the undirected graphs are simple or not; that is, whether they contain loops or parallel edges. So we state the result for any two connected nontrivial (not equal to K1) undirected graphs. Remark 3.4. The description for the boundary set of the strong product of two graphs (undirected graphs) G and H presented in [4] holds only for the product of two nontrivial graphs G and H . To this end, let H = K1 = ({v}, ∅). We have, ∂(K1) = {v} (since all vertices of a complete graph are boundary vertices of the graph), and hence ∂(G) =̂ ∂(G⊠K1) = (∂(G)× {v}) ∪ (V (G)× {v}) =̂ V (G), which is not true in general. Corollary 3.5. Let D1 and D2 be two nontrivial connected undirected graphs. Then ∂(D1 ⊠D2) = [∂(D1)× V (D2)] ∪ [V (D1)× ∂(D2)]. Proof. By Theorem 3.3, if D1 and D2 are two strongly connected digraphs, ∂(D1⊠D2) = A1∪A2∪A3. Since D1 and D2 are given to be two nontrivial undirected graphs, ecc(ui) ≥ 1 for all ui ∈ V (D1), ecc(vr) ≥ 1 for all vr ∈ V (D2), d(ui, uk) = 1 for all uk ∈ N(ui), and d(vr, vq) = 1 for all vq ∈ N(vr). Thus, A1 = ∂(D1)× ∂(D2), A2 = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ ∂(D1), vr /∈ ∂(D2), and ∃vt ∈ V (D2) such that d(vt, vq) ≤ ecc(ui),∀vq ∈ N(vr)} = ∂(D1)× V (D2), A3 = {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ ∂(D1), vr ∈ ∂(D2), and ∃uℓ ∈ V (D1) such that d(uℓ, uk) ≤ ecc(vr),∀uk ∈ N(ui)} = V (D1)× ∂(D2). Therefore, ∂(D1⊠D2) = A1∪A2∪A3 = [∂(D1)×V (D2)]∪ [V (D1)×∂(D2)]. Theorem 3.6. Let D1 and D2 be two strongly connected digraphs. 1. If diam(D1) < diam(D2), then Per(D1 ⊠D2) = V (D1)× Per(D2). 2. If diam(D1) = diam(D2), then Per(D1 ⊠D2) = [Per(D1)× V (D2)] ∪ [V (D1)× Per(D2)]. 284 Ars Math. Contemp. 20 (2021) 275–288 Proof. 1. Let diam(D2) = n. Let vr ∈ Per(D2). Then for all ui ∈ V (D1), ecc(ui, vr) = max{ecc(ui), ecc(vr)} = n. Hence (ui, vr) ∈ Per(D1 ⊠ D2). Also if vr /∈ Per(D2), then since ecc(ui, vr) < n, (ui, vr) /∈ Per(D1⊠D2). Hence it follows that Per(D1⊠D2) = V (D1)×Per(D2). 2. Let diam(D1) = diam(D2) = n. If ui ∈ Per(D1), then for all vr ∈ V (D2), (ui, vr) ∈ Per(D1 ⊠D2), since ecc(ui, vr) = max{ecc(ui), ecc(vr)} = n. Hence (ui, vr) ∈ Per(D1 ⊠ D2). Similarly, if vr ∈ Per(D2), then for all ui ∈ V (D1), (ui, vr) ∈ Per(D1 ⊠ D2). Hence it follows that [Per(D1) × V (D2)] ∪ [V (D1) × Per(D2)] ⊆ Per(D1 ⊠D2). Conversely, if (ui, vr) ∈ Per(D1 ⊠ D2), then ecc(ui, vr) = max{diam(D1), diam(D2)} = n. Thus, at least one of ecc(ui) and ecc(vr) must be necessarily equal to n. Hence ui ∈ Per(D1) or vr ∈ Per(D2), and therefore, Per(D1 ⊠D2) ⊆ [Per(D1)× V (D2)] ∪ [V (D1)× Per(D2)]. Theorem 3.7. Let D1 and D2 be two strongly connected digraphs. 1. If rad(D1) = rad(D2), then Ecc(D1 ⊠D2) = [Ecc(D1)× V (D2)] ∪ [V (D1)× Ecc(D2)]. 2. If rad(D1) < rad(D2), then Ecc(D1 ⊠D2) = [ ⋃ ecc(ui)≥rad(D2) Ecc(ui)× V (D2) ] ∪ [V (D1)× Ecc(D2)] . Proof. 1. First we will prove that Ecc(D1 ⊠ D2) ⊆ [Ecc(D1) × V (D2)] ∪ [V (D1) × Ecc(D2)]. Let (ui, vr) ∈ Ecc(D1 ⊠ D2). Then there exists a vertex (uj , vs) such that ecc(uj , vs) = d((uj , vs), (ui, vr)) = max{d(uj , ui), d(vs, vr)}. Since ecc(uj , vs) = max{ecc(uj), ecc(vs)}, and ecc(uj) ≥ d(uj , ui) and ecc(vs) ≥ d(vs, vr), at least one of ecc(uj) = d(uj , ui) and ecc(vs) = d(vs, vr) must hold. So necessarily ui is an eccentric vertex of uj , or vr is an eccentric vertex of vs. Hence (ui, vr) ∈ [Ecc(D1)× V (D2)] ∪ [V (D1)× Ecc(D2)]. Let rad(D1) = rad(D2) = n. Let ui ∈ Ecc(D1). Then there exists a vertex uj ∈ V (D1) such that ecc(uj) = d(uj , ui). Consider the vertex (ui, vr) ∈ V (D1 ⊠D2), where vr is an arbitrary vertex in D2. Since rad(D2) = n, there exists a vertex vs ∈ V (D2) such that ecc(vs) = n. Hence d(vs, vr) ≤ n and so ecc(uj , vs) = max{ecc(uj), ecc(vs)} = max{ecc(uj), n} = ecc(uj). Thus, d((uj , vs), (ui, vr)) = max{d(uj , ui), d(vs, vr)} = ecc(uj) = ecc(uj , vs). So (ui, vr) is an eccentric vertex of (uj , vs). Thus if ui ∈ Ecc(D1), then (ui, vr) ∈ Ecc(D1 ⊠ D2) for all vr ∈ V (D2). Similarly, we can prove that if vq ∈ Ecc(D2), then (uk, vq) ∈ Ecc(D1 ⊠D2) for all uk ∈ V (D1). Hence [Ecc(D1)× V (D2)]∪ [V (D1)×Ecc(D2)] ⊆ Ecc(D1 ⊠D2), and so the result holds. B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 285 2. Let rad(D1) < rad(D2) = n. Let ui ∈ V (D1), vr ∈ V (D2). Here two cases arise: Case 1. vr ∈ Ecc(D2). Then there exists a vertex vs ∈ V (D2) such that ecc(vs) = d(vs, vr). Let up ∈ V (D1) be such that ecc(up) = rad(D1). Then since rad(D2) > ecc(up), ecc(up, vs) = max{ecc(up), ecc(vs)} = ecc(vs). Also, d((up, vs), (ui, vr)) = max{d(up, ui), d(vs, vr)} = ecc(vs). Thus, (ui, vr) is an eccentric vertex of (up, vs). So in this case, V (D1)× Ecc(D2) ⊆ Ecc(D1 ⊠D2). Case 2. vr /∈ Ecc(D2). Let vq ∈ V (D2) be such that ecc(vq) = rad(D2). Let ⋃ ecc(ui)≥rad(D2) Ecc(ui) = A. Let uk ∈ A. Then there exists a vertex up ∈ V (D1) such that ecc(up) ≥ rad(D2) and ecc(up) = d(up, uk). Then d((up, vq), (uk, vr)) = max{d(up, uk), d(vq, vr)} = d(up, uk) = ecc(up) = ecc(up, vq) and hence (uk, vr) is an eccentric vertex of (up, vq). Hence in this case, ⋃ ecc(ui)≥rad(D2) Ecc(ui)× V (D2) ⊆ Ecc(D1 ⊠D2). Thus, [⋃ ecc(ui)≥rad(D2) Ecc(ui)× V (D2) ] ∪ [V (D1)× Ecc(D2)] ⊆ Ecc(D1 ⊠D2). Conversely, let (uk, vr) ∈ Ecc(D1 ⊠ D2). Then there exists a vertex (uj , vs) ∈ V (D1⊠D2) such that ecc(uj , vs) = d((uj , vs), (uk, vr)) = max{d(uj , uk), d(vs, vr)} = max{ecc(uj), ecc(vs)}. If vr ∈ Ecc(D2), we get (uk, vr) ∈ V (D1)× Ecc(D2). Hence suppose that (uk, vr) ∈ Ecc(D1 ⊠ D2) and vr /∈ Ecc(D2). Then for all vs ∈ V (D2), ecc(vs) > d(vs, vr). Thus, ecc(uj , vs) = ecc(uj) = d(uj , uk). If possible, suppose that uk /∈ A = ⋃ ecc(ui)≥rad(D2) Ecc(ui). Thus, there is no vertex uj in D1 such that ecc(uj) = d(uj , uk) and ecc(uj) ≥ rad(D2). Hence if uk is an eccentric vertex of uj in D1, then d(uj , uk) < rad(D2). We have, rad(D1⊠D2) = max{rad(D1), rad(D2)} = rad(D2). Thus, (uk, vr) cannot be the eccentric vertex of any vertex (uj , vs) ∈ D1 ⊠D2, since d((uj , vs), (uk, vr)) = max{d(uj , uk), (vs, vr)} ≠ ecc(uj , vs) in this case. This is a contradiction, and hence uk ∈ A. Hence (uk, vr) ∈ ⋃ ecc(ui)≥rad(D2) Ecc(ui)× V (D2). Hence Ecc(D1⊠D2) ⊆ [⋃ ecc(ui)≥rad(D2) Ecc(ui)×V (D2) ] ∪ [V (D1)×Ecc(D2)]. Theorem 3.8. Let D1 and D2 be two strongly connected digraphs. Then Ct(D1 ⊠D2) = A1 ∪A2 ∪A3, where A1 = [Ct(D1)× Ct(D2)], A2 = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vq) ≤ ecc(ui) for all vq ∈ N [vr]}, A3 = {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ Ct(D1), vr ∈ Ct(D2), and ecc(uk) ≤ ecc(vr) for all uk ∈ N [ui]}. Proof. (ui, vr) ∈ Ct(D1 ⊠D2) if and only if ecc(ui, vr) ≥ ecc(uk, vq) for all (uk, vq) ∈ N [(ui, vr)]; if and only if max{ecc(ui), ecc(vr)} ≥ max{ecc(uk), ecc(vq)} for all uk ∈ N [ui] and vq ∈ N [vr]; if and only if one of the following three cases holds. 1. max{ecc(ui), ecc(vr)} = ecc(ui) = ecc(vr). Then, ecc(ui) ≥ ecc(uk) and ecc(vr) ≥ ecc(vq) for all uk ∈ N [ui] and vq ∈ N [vr]. 2. max{ecc(ui), ecc(vr)} = ecc(ui) > ecc(vr). Then, ecc(ui) ≥ ecc(uk) for all uk ∈ N [ui] and ecc(vr) < ecc(ui), ecc(vq) ≤ ecc(ui) for all vq ∈ N(vr). 286 Ars Math. Contemp. 20 (2021) 275–288 3. max{ecc(ui), ecc(vr)} = ecc(vr) > ecc(ui). Then, ecc(vr) ≥ ecc(vq) for all vq ∈ N [vr] and ecc(ui) < ecc(vr), ecc(uk) ≤ ecc(vr) for all uk ∈ N(ui). In case 1, (ui, vr) ∈ Ct(D1 ⊠D2). In case 2, (ui, vr) ∈ {(ui, vr) ∈ V (D1 ⊠ D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vq) ≤ ecc(ui) for all vq ∈ N [vr]}. In case 3, (ui, vr) ∈ {(ui, vr) ∈ V (D1 ⊠ D2) : ui /∈ Ct(D1), vr ∈ Ct(D2), and ecc(uk) ≤ ecc(vr) for all uk ∈ N [ui]}. Thus we get, Ct(D1 ⊠D2) = A1 ∪A2 ∪A3. Consider the contour set of the strong product of two connected undirected graphs. As in the case of the boundary set, the result holds even when the undirected graphs are not simple. Corollary 3.9. Let D1 and D2 be two connected undirected graphs. Then Ct(D1 ⊠D2) = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vr) < ecc(ui)} ∪ {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ Ct(D1), vr ∈ Ct(D2), and ecc(ui) < ecc(vr)} ∪ [Ct(D1)× Ct(D2)]. Proof. By Theorem 3.8, when D1 and D2 are two strongly connected digraphs, Ct(D1 ⊠ D2) = A1 ∪A2 ∪A3. Since D1 and D2 are given to be undirected graphs, eccentricity of two adjacent vertices differ by atmost one. Hence A1 = Ct(D1)× Ct(D2), A2 = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vq) ≤ ecc(ui) for all vq ∈ N [vr]} = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vr) + 1 ≤ ecc(ui)} = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vr) < ecc(ui)}, A3 = {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ Ct(D1), vr ∈ Ct(D2), and ecc(ui) < ecc(vr)}, since maxuk∈N [ui] ecc(uk) = ecc(ui) + 1. Hence it follows that Ct(D1 ⊠D2) = {(ui, vr) ∈ V (D1 ⊠D2) : ui ∈ Ct(D1), vr /∈ Ct(D2), and ecc(vr) < ecc(ui)} ∪ {(ui, vr) ∈ V (D1 ⊠D2) : ui /∈ Ct(D1), vr ∈ Ct(D2), and ecc(ui) < ecc(vr)} ∪ [Ct(D1)× Ct(D2)]. We have examined the boundary-type sets of the strong product of two strongly con- nected digraphs D1 and D2. Now suppose that at least one of D1 and D2, say D1, is not strongly connected. Then the eccentricity of every vertex in D1 is infinity, and hence the B. S. Anand et al.: Boundary-type sets of strong product of directed graphs 287 eccentricity of every vertex in D1 ⊠ D2 is infinity. Thus, we have ∂(D1) = Per(D1) = Ecc(D1) = Ct(D1) = V (D1), and ∂(D1 ⊠D2) = Per(D1 ⊠D2) = Ecc(D1 ⊠D2) = Ct(D1 ⊠ D2) = V (D1) × V (D2). Since rad(D1) = diam(D1) = ∞, the expression for Per(D1 ⊠D2) in Theorem 3.6, and the expression for Ecc(D1 ⊠D2) in Theorem 3.7 gives V (D1) × V (D2). Since ecc(ui) = ∞ for all ui ∈ V (D1), the expression for ∂(D1 ⊠ D2) in Theorem 3.3, and the expression for Ct(D1 ⊠ D2) in Theorem 3.8 also gives V (D1)× V (D2). Similar is the case when D2 and both D1 and D2 are not strongly connected. Thus, the results derived for the boundary-type sets of the strong product of two strongly connected digraphs D1 and D2 hold also when the digraphs D1 and D2 are not even weakly connected. So the results for the boundary-type sets of the strong product of two connected undirected graphs hold for any two arbitrary undirected graphs. 4 Conclusion In this article, the relationship between the boundary-type sets of the strong product of two digraphs, and that of its factors is derived. As ‘maximum distance’ is the generalization of the usual distance in an undirected graph, these results hold for undirected graphs also. The results for the periphery and eccentricity sets of the strong product of two undirected graphs turn out to be the same as the results in [4]. The results for the boundary and contour sets in the undirected case, as described in [4], are also derived as special cases. ORCID iDs Bijo S. Anand https://orcid.org/0000-0002-7221-904X Manoj Changat https://orcid.org/0000-0001-7257-6031 Prasanth G. Narasimha-Shenoi https://orcid.org/0000-0002-5850-5410 Mary Shalet Thottungal Joseph https://orcid.org/0000-0001-6350-7106 References [1] J. Bang-Jensen and G. Gutin (eds.), Classes of Directed Graphs, Springer Monographs in Math- ematics, Springer, Cham, 2018, doi:10.1007/978-3-319-71840-8. [2] A. Broido and K. C. Claffy, Internet topology: Connectivity of IP graphs, in: S. Fahmy and K. Park (eds.), Scalability and Traffic Control in IP Networks, SPIE, volume 4526 of Con- ference Proceedings of SPIE, 2001 pp. 172–187, doi:10.1117/12.434393, proceedings of the International Symposium on the Convergence of IT and Communications (ITCom 2001) held in Denver, Colorado, United States, August 20 – 24, 2001. [3] J. Cáceres, C. Hernando, M. Mora, I. M. Pelayo, M. L. Puertas and C. Seara, On geodetic sets formed by boundary vertices, Discrete Math. 306 (2006), 188–198, doi:10.1016/j.disc.2005. 12.012. [4] J. Cáceres, M. del C. Hernando Martín, M. Mora Giné, I. M. Pelayo Melero and M. L. Puertas González, Boundary-type sets and product operators in graphs, in: VII Jornadas de Matemática Discreta y Algorítmica, Castro Urdiales, 2010 pp. 187–194, http://hdl. handle.net/2117/8383. [5] J. Cáceres, A. Márquez, O. R. Oellermann and M. Luz Puertas, Rebuilding convex sets in graphs, Discrete Math. 297 (2005), 26–37, doi:10.1016/j.disc.2005.03.020. 288 Ars Math. Contemp. 20 (2021) 275–288 [6] M. Changat, P. G. Narasimha-Shenoi, M. S. Thottungal Joseph and R. Kumar, Boundary ver- tices of Cartesian product of directed graphs, Int. J. Appl. Comput. Math. 5 (2019), Article No. 19, doi:10.1007/s40819-019-0604-4. [7] G. Chartrand, D. Erwin, G. L. Johns and P. Zhang, Boundary vertices in graphs, Discrete Math. 263 (2003), 25–34, doi:10.1016/s0012-365x(02)00567-8. [8] G. Chartrand and S. Tian, Distance in digraphs, Computers Math. Applic. 34 (1997), 15–23, doi:10.1016/s0898-1221(97)00216-2. [9] R. Hammack, W. Imrich and S. Klavžar, Handbook of Product Graphs, Discrete Mathematics and its Applications (Boca Raton), CRC Press, Boca Raton, FL, 2nd edition, 2011, doi:10. 1201/b10959. [10] R. H. Hammack, Digraphs products, in: J. Bang-Jensen and G. Gutin (eds.), Classes of Directed Graphs, Springer, Cham, pp. 467–515, 2018, doi:10.1007/978-3-319-71840-8_10. [11] M. Hellmuth and T. Marc, On the Cartesian skeleton and the factorization of the strong product of digraphs, Theoret. Comput. Sci. 565 (2015), 16–29, doi:10.1016/j.tcs.2014.10.045. [12] M. W. Massing, G. A. Robinson, C. E. Marx, O. Alzate and R. D. Madison, Applications of proteomics to nerve regeneration research, in: O. Alzate (ed.), Neuroproteomics, CRC Press, Boca Raton, FL, Frontiers in Neuroscience, pp. 289–313, 2009. [13] H. E. Robbins, Questions, discussions, and notes: A theorem on graphs, with an application to a problem of traffic control, Amer. Math. Monthly 46 (1939), 281–283, doi:10.2307/2303897. [14] J. J. Stemley, One-way streets provide superior safety and convenience, ITE Journal 68 (1998), 47–50. 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). xv 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 2021. 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 xvi SIGMAP 2022 Workshop – Announcement and Call for Papers The Symmetries in Graphs, Maps, and Polytopes Workshop (SIGMAP) has been held every four years since Steve Wilson organized the first one in Flagstaff, Arizona in 1998. The workshop is devoted to the exploration of the symmetries of discrete objects, and has been an opportunity to share recent advances, discuss open problems, and start new col- laborations. In addition to two daily sessions of talks, SIGMAP 2022 will set aside time each day for researchers to gather and meet to explore questions in the field. The plenary speakers for SIGMAP 2022 are: • Gabriel Cunningham (University of Massachusetts Boston, USA) • Maria Elisa Fernandes (University of Aveiro, Portugal) • Gareth Jones (University of Southampton, United Kingdom) • Klavdija Kutnar (University of Primorska, Slovenia) • Primož Šparl (University of Ljubljana and University of Primorska, Slovenia) • Pablo Spiga (Università degli Studi di Milano-Bicocca, Italy) • Klara Stokes (Umeå University, Sweden) • Gabriel Verret (University of Auckland, New Zealand) • Jinxin Zhou (Beijing Jiaotong University, China) xvii The workshop will be held at the University of Alaska Fairbanks campus. UAF is Alaska’s flagship research university, with an enrollment of over 7 000 students. The campus is situated on 2 250 acres amidst the boreal forest, and features 41.6 km of trails, the Museum of the North, botanical gardens, an experimental farm, and a viewing area for our herds of musk ox and reindeer. If you’re lucky, you’ll spot some of the campus’s wild moose during a visit to the campus. More information about the workshop is available at https://www.alaska.edu/sigmap This is also a call for papers for a special issue of the journal The Art of Discrete and Applied Mathematics (ADAM). Papers submitted for this special issue should be on top- ics presented or discussed at the workshop, or closely related to them. The Art of Discrete and Applied Mathematics (ADAM) is a modern, dynamic, platinum open access, electronic journal that publishes high-quality articles in contemporary discrete and applied mathemat- ics (including pure and applied graph theory and combinatorics), with no costs to authors or readers. To be considered for inclusion in this special issue, papers should be submitted by December 31, 2022, via the ADAM website https://adam-journal.eu/. A template and style file for submissions can be downloaded from that website, or obtained from one of the guest editors on request. The ideal length of papers is 5 to 15 pages, but longer or shorter papers will certainly be considered. Papers that are accepted will appear on-line soon after acceptance, and papers that are not processed in time for the special issue may still be accepted and published in a subsequent regular issue of ADAM. Leah Berman and Gordon Williams Guest Editors xviii Printed in Slovenia by IME TISKARNE