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.