Home Mathematics Connected domination game played on Cartesian products
Article Open Access

Connected domination game played on Cartesian products

  • Csilla Bujtás , Pakanun Dokyeesun , Vesna Iršič and Sandi Klavžar EMAIL logo
Published/Copyright: November 8, 2019

Abstract

The connected domination game on a graph G is played by Dominator and Staller according to the rules of the standard domination game with the additional requirement that at each stage of the game the selected vertices induce a connected subgraph of G. If Dominator starts the game and both players play optimally, then the number of vertices selected during the game is the connected game domination number of G. Here this invariant is studied on Cartesian product graphs. A general upper bound is proved and demonstrated to be sharp on Cartesian products of stars with paths or cycles. The connected game domination number is determined for Cartesian products of P3 with arbitrary paths or cycles, as well as for Cartesian products of an arbitrary graph with Kk for the cases when k is relatively large. A monotonicity theorem is proved for products with one complete factor. A sharp general lower bound on the connected game domination number of Cartesian products is also established.

MSC 2010: 05C57; 05C69; 91A43

1 Introduction

Connected domination game on a graph G is played by two players, usually named Dominator and Staller. They play in turns, at each move selecting a single vertex of G such that it dominates at least one vertex that is not yet dominated with the previously played vertices and such that at each stage of the game the selected vertices induce a connected subgraph of G. If Dominator has the first move, then we speak of a D-game, otherwise they play an S-game. When the game is finished, that is, when there is no legal move available, the players have determined a connected dominating set D of G. The goal of Dominator is to finish with |D| being as small as possible, the goal of Staller is just the opposite. If both players play optimally, then |D| is unique. In the D-game it is called the connected game domination number γcg(G) of G, while when S-game is played, the corresponding invariant is denoted by γcg (G). The connected domination game is thus defined as the standard domination game [1] with the additional requirement that the players maintain connectedness of the selected vertices at all times. Other games closely related to the domination game are the total domination game [2, 3, 4, 5], the disjoint domination game [6], the transversal game on hypergraphs [7, 8], Maker-Breaker domination game [9, 10], Maker-Breaker total domination game [11], as well as the whole variety of domination games as described in [12].

The connected domination game was introduced by Borowiecki, Fiedorowicz, and Sidorowicz in [13]. Among other results they proved a sharp upper bound on γcg of 2-trees, studied the game on the Cartesian product, and considered the S-game and Staller-pass games. The game does not admit the so-called Continuation Principle [14], which is an utmost useful tool for the standard game, cf. [15, 16, 17, 18, 19]. Instead, [13] brings the so-called game with Chooser (to be described below) which plays a similar role for the connected game. The game was further investigated in [20] where a question from [13] on a relation between the D-game and the S-game is answered, the relation of the game with the diameter investigated, the game on the lexicographic product completely explained, and the effect of a vertex predomination studied.

In this paper we focus on the connected domination game played on Cartesian products. In Section 2 we give a general upper bound on it that in particular extends the upper bound on γcg(GKm) from [13]. The sharpness of the bound is demonstrated by several earlier exact results. In the subsequent section we determine the connected game domination number of the Cartesian product of stars with paths or with cycles. This yields additional families for which the upper bound from the earlier section is sharp. We continue by determining γcg(P3Pn) and γcg(P3Cn) in Section 4. In the subsequent section we consider products in which one factor is complete. We first prove that γcg(GKk) = 2n(G) – 1, provided that k is not too small. Then we obtain a monotonicity theorem asserting that γcg(Kn+1G) ≥ γcg(KnG) holds for n ≥ 1. In Section 6 we prove a general lower bound on the connected game domination number of Cartesian products and establish its sharpness on several infinite families. We conclude with some general thoughts about the connected domination game and pose several open problems.

In the rest of the introduction we present the connected domination game with Chooser and give further definitions needed later on.

The connected domination game with Chooser [13] has similar rules as the normal game, except that there is another player, Chooser, who can make zero, one, or more moves after any move of Dominator or Staller. The conditions for his move to be legal are the same as for Dominator and Staller. Chooser has no specific goal, he can help either Dominator or Staller. We recall the Chooser Lemma from [13] to be used later on.

Lemma 1.1

(Chooser Lemma). Consider the connected domination game with Chooser on a graph G. Suppose that in the game Chooser picks k vertices, and that both Dominator and Staller play optimally. Then at the end of the game the number of played vertices is at most γcg(G) + k and at least γcg(G) – k.

The Cartesian product GH of graphs G and H has the vertex set V(G) × V(H), vertices (g, h) and (g′, h′) being adjacent if either gg′ ∈ E(G) and h = h′, or g = g′ and hh′ ∈ E(H). If hV(H), then the subgraph of GH induced by the vertex set {(g, h) : gV(G)} is isomorphic to G, called a G-layer, and denoted with Gh. Analogously the H-layers are defined and denoted with gH for a fixed vertex gV(G). For more on the Cartesian product of graphs see [21].

The closed neighborhood of a vertex v in a graph G is denoted by N[v]. The closed neighborhood of a set of vertices SV(G) is N[S] = ⋃vS N[v]. Similarly, the open neighborhood of v is N(v) = N[v] ∖ {v}. A set SV(G) is a dominating set of the graph G if N[S] = V(G). The minimal cardinality of a dominating set is the domination number γ(G) of G. A connected dominating set is a set S that is both connected and dominating. The smallest cardinality of such set is the connected domination number γc(G) of G. The largest size of an independent set of vertices in G is the independence number α(G) of G. Additionally, we denote [n] = {1, …, n}, Δ(G) as the maximum degree of vertices of G, and n(G) as the number of vertices of G.

2 A general upper bound

In this section we prove a general upper bound on the connected game domination number of Cartesian products and demonstrate its sharpness by several earlier results. A key for the bound is the following theorem from the seminal paper [13]. We reprove the result here in more detail, to give additional insight.

Theorem 2.1

[13,Theorem 1] If G is a connected graph, then γcg(G) ≤ 2γc(G) – 1.

Proof

Let = γc(G) and let D = {x1, …, x} be a connected dominating set in G, where for every i ∈ [], the set {x1, …, xi} induces a connected subgraph of G. As D is a connected dominating set, such an ordering of the vertices of D is clearly possible.

The strategy of Dominator is to consecutively play the vertices x1, …, x, in particular, his first move is on the vertex x1. Suppose now that xi, i ≤ 2, is the next vertex to be played by Dominator according to his strategy and that he is not able to play it.

It is possible that Dominator cannot play xi because xi was already played by Staller in the course of the game. If this is the case, Dominator simply plays xi+1, provided this is a legal move. If it is not and also xi+1 was already played by Staller, we continue this process until we reach a vertex xj, ji, that was neither played earlier by Staller nor it is playable by Dominator. Because of the latter fact, all the vertices from N[xj] are already dominated at this stage of the game. If each of the vertices xj+1, …, x was either already played by Staller or is unplayable by Dominator, then the set of vertices selected so far constitutes a connected dominating set of G. Indeed, all the vertices from each of N[x1], …, N[x] are dominated, hence G is dominated because {x1, …, x} is a dominating set. Moreover, it is a connected dominating set because the strategy of Dominator preserves the connectedness (and Staller must follow it anyway).

Assume therefore that there exits a vertex xk, j + 1 ≤ k, that is undominated and let k be the smallest possible index of such a vertex. If xk is adjacent to some vertex already played, then Dominator plays xk. Otherwise, consider an arbitrary neighbor xs of xk from the sequence x1, …, xk–1. If xs was played in the game played so far, then Dominator can extend the game by playing xk. So suppose that xs was not played before. This means that xs was (and is) an unplayable vertex. In particular, its neighbor xk must have been dominated by some vertex already played. But this means again that Dominator can play xk because it is adjacent to an already played vertex.

By the above strategy of Dominator, the game is finished no later than at the moment when Dominator plays x. If this happens and if Dominator played all the vertices x1, …, x, then the number of vertices played is exactly 2 – 1 = 2γc(G) – 1. If, however, some of the vertices xi were played by Staller or became unplayable during the game, the number of vertices played is even smaller.□

We can now quickly derive the announced upper bound.

Theorem 2.2

If G and H are connected graphs, then

γcg(GH)min{2γc(G)n(H),2γc(H)n(G)}1.

Proof

Let XG be a connected dominating set of G with |XG| = γc(G). Since XG induces a connected subgraph of G, the set of vertices XG × V(H) induces a connected subgraph of GH and hence it is a connected dominating set of GH. Therefore, γc(GH) ≤ 2γc(G)n(H) and thus γcg(GH) ≤ 2γc(G)n(H) – 1 by Theorem 2.1. By a parallel argument (or by referring to the commutativity of the Cartesian product) we also have γcg(GH) ≤ 2γc(H)n(G) – 1.□

Before listing known results that demonstrate the sharpness of the bound of Theorem 2.2, we note that the following result is a direct consequence of the theorem.

Corollary 2.3

[13, Theorem 5] If G is a connected graph, then γcg(GKm) ≤ min {2c(G)) – 1, 2n(G) – 1}.

In [13, Theorem 6, Theorem 7] it was proved that if 2 ≤ mn ≥ 4, then

γcg(K1,n1Km)=min{2n1,2m1},

and if m, n ≥ 4, then

γcg(PnKm)=2n1.

These results demonstrate that the bound of Theorem 2.2 is sharp.

Let next r ≥ 2 and n1n2 ≥ ⋯ ≥ nr ≥ 2. By the associativity of the Cartesian product we have Kn1Kn2 ◻ ⋯ ◻ Knr = Kn1 ◻ (Kn2 ◻ ⋯ ◻ Knr), hence applying Theorem 2.2 we get

γcg(Kn1Kn2Knr)2n2nr1.

Since it is proved in [20,Proposition 2.3] that the equality holds here provided that n1 ≥ 2n2nr, this gives another example for the sharpness of Theorem 2.2.

3 Products of stars with paths or cycles

In this section we determine the connected game domination number for the Cartesian product of stars with paths or with cycles. In both cases the bound of Theorem 2.2 is sharp.

Theorem 3.1

If n ≥ 3 and m ≥ 1, then

γcg(K1,nPm)=2m1.

Proof

The statement is clear for m = 1 and easily verified for m = 2. For m = 3 it holds γcg(K1,nP3) = 5. Indeed, the upper bound follows from Theorem 2.2. The lower bound can be shown with a simple case analysis. In the remaining part of the proof, we consider m ≥ 4.

Consider a path Pm = u1um (m ≥ 4) and a star K1,n on the vertex set v1, …, vn+1 (n ≥ 3) where vn+1 is the center. In K1,nPm, the layer K1,nuj will be denoted simply by K1,nj . In a connected domination game on K1,nPm, let Ai (resp. Ai ) denote the set of integers j ∈ [m] for which at least one vertex from the layer K1,nj was played within the moves d1, s1, …, di (resp. d1, s1, …, di, si). By connectivity, both Ai and Ai are sets of consecutive integers. We will use the notation Ai = [ai, bi] and Ai=[ai,bi]. By Theorem 2.2, we have γcg(K1,nPm) ≤ 2m – 1. To prove the lower bound we describe a strategy of Staller that maintains the following property for every i.

  1. After the move si of Staller, either at least two vertices are played from every layer K1,nj with j[ai,bi] {1, m}, or there is only one layer K1,nj (j* = ai > 1 or j* = bi < m) among them which is played only once. In the latter case, the vertex played in the layer K1,nj is a leaf of the layer, and there are at least two leaves played from the neighboring layer.

Property (⋆) and the strategy of Staller will ensure that from every K1,nj , with 2 ≤ jm – 1, at least two vertices will be played during the game. The below rule (S1) in Staller’s strategy clearly establishes (⋆) after s1. Then, we will suppose that it is true before Dominator’s move di (i ≥ 2). Note that, under this condition, we may have at most two layers K1,nj with 2 ≤ jm – 1 which are played exactly once after the move di.

  1. If Dominator plays the center of a layer K1,nj as his first move d1, Staller replies by playing the leaf (v1, uj) of the same star; if Dominator plays a leaf from K1,nj , Staller picks the center (vn+1, uj) as s1.

Later, if Dominator plays the first vertex from K1,nbi as di, that is, if bi = bi1 , and bi < m, then Staller replies according to the following rules:

  1. If at least two vertices from the neighboring layer K1,nbi1 have been played in the game so far, Staller chooses a vertex from K1,nbi . This clearly can be done as at most one vertex from K1,nbi+1 has been dominated so far.

  2. If only one vertex from K1,nbi1 has been played so far, Staller chooses a leaf from K1,nbi1 . Such a legal move exists as (⋆) was true after the move si–1.

The rules are analogous (and we do not repeat them) for the case when Dominator plays the first vertex from K1,nai as di (and ai > 1). If Dominator plays a vertex from K1,nm or K1,n1 as di, or he plays a vertex from a layer K1,nj which has been played earlier (i.e.,j[ai1,bi1]), Staller replies as follows:

  1. If bi < m and K1,nbi is played only once, Staller chooses a vertex from K1,nbi . This clearly can be done as at most one vertex from K1,nbi+1 has been dominated so far. The rule is similar for the case when ai > 1 and K1,nai is played only once.

  2. If both K1,nbi and K1,nai are played at least twice, Staller plays a vertex from K1,nbi or K1,nai , if such a legal move exists.

  3. If there are no playable vertices in K1,nbi (and bi < m), then all vertices of the layer have been played. If the situation is similar for K1,nai , Staller plays the leaf (v1, ubi+1) if bi < m – 1, and she plays (v1, uai–1) if bim – 1 and ai > 2.

What remains is only the case when all the layers K1,nj with 2 ≤ jm – 1 have been played at least twice.

  1. Staller plays a vertex from K1,nm1 or from K1,n2 . Note that it is always possible if the game is not yet over.

One can check that property (⋆) is preserved if Staller applies the above strategy. Remark that, by (S5) and (S6), Staller plays a first vertex from a K1,nj only if all vertices of the neighboring layer have already been played. Moreover, if Staller plays the first vertex in a layer, then this vertex corresponds to a leaf of the star. Consequently, if we have two neighboring layers which are played only once, then these played vertices must be leaves. In particular, we may have the following two situations before the end of the game:

  1. If Staller finishes the game and in the last turn a vertex from K1,nm becomes dominated, then the center of this star is not played and, therefore, at least n + 1 vertices were played from these two stars. Moreover, every layer K1,nj , 2 ≤ jm – 2, is played at least twice and the domination of K1,n1 needs at least one further vertex. We may infer that at least n + 1 + 2(m – 3) + 1 ≥ 2m – 1 vertices were played during the game. (The argumentation is similar if a vertex from K1,n1 was dominated in the last turn.)

  2. If Dominator finishes the game, at least two vertices are chosen from every K1,nj , if 2 ≤ jm – 1. The domination of K1,n1 and K1,nm needs at least two further vertices. Hence, at least 2m – 2 vertices were played. On the other hand, the number of chosen vertices must be odd as, by our assumption, Dominator finishes the game. This proves γcg(K1,nPm) ≥ 2m – 1.

This finishes the proof of the equality.□

Theorem 3.2

If n ≥ 3 and m ≥ 3, then

γcg(K1,nCm)=2m1.

Proof

For m = 3, it can easily be checked that γcg(K1,nC3) = 5. Thus we only consider m ≥ 4.

By Theorem 2.2, γcg(K1,nCm) ≤ 2m – 1. To prove the lower bound, we use a similar strategy of Staller as in the proof of Theorem 3.1. We also use the same notation, but now the addition is done modulo m (with values in [m]), so the sets Ai and Ai are sets of consecutive integers modulo m. The main difference is only that in the strategy on K1,nPm layers K1,n1 and K1,nm are studied differently, but on the K1,nCm the special K1,n-layers are the ones which are played last or not played at all. Similarly, all the conditions like 1 < ai and bi < m must be replaced with |biai|m3. Apart from that, Staller’s strategy and the final counting of the moves remains the same.□

4 Products of P3 with paths or cycles

In Theorems 3.1 and 3.2 we have considered Cartesian products of stars K1,n, n ≥ 3, with paths or cycles. For the star K1,2 = P3 the situation is slightly different, in particular, the bound of Theorem 2.2 is no longer sharp.

Theorem 4.1

If m ≥ 4, then

γcg(P3Pm)=2m2.

Proof

Let P3 = v1v2v3 and Pm = u1um. We will also use the notations introduced in the proof of Theorem 3.1.

To prove that γcg(P3Pm) ≤ 2m – 2, we consider the following strategy of Dominator in a connected domination game with Chooser and apply the Chooser Lemma 1.1. Let d1 = (v2, u2). Then, Staller has four different legal replies.

  1. If Staller responds with the move s1 = (v1, u2), then Chooser picks the vertices (v2, u3), …, (v2, um–1). Dominator’s strategy is to play d2 = (v2, um). Then, only the vertex (v3, u1) remains undominated and the game finishes with a set of m + 1 played vertices. By Chooser Lemma, we have t ≤ (m + 1) + (m – 3) = 2m – 2, where t is the length of the connected domination game under the described strategies of the players.

  2. The reply s1 = (v3, u2) is treated analogously as the move s1 = (v1, u2).

  3. If s1 = (v2, u3), Chooser picks (v2, u4), …, (v2, um) and Dominator plays d2 = (v2, u1). By Chooser Lemma, we have tm + (m – 3) = 2m – 3.

  4. If s1 = (v2, u1), Chooser picks (v2, u3), …, (v2, um–1) and Dominator plays d2 = (v2, um). By Chooser Lemma, we have tm + (m – 3) = 2m – 3, again.

This proves that Dominator can ensure t ≤ 2m – 2 under any strategy of Staller, that is, γcg(P3Pm) ≤ 2m – 2.

To prove the other direction, that is, γcg(P3Pm) ≥ 2m – 2, we consider the strategy (S1)–(S7) of Staller and property (⋆) as these are given in the proof of Theorem 3.1. In fact, as P3 = K1,2, we have the same line of the proof except the final counting which is the following for the case of P3Pm. By Staller’s strategy, at least two vertices were played from every P3j with 2 ≤ jm – 1. Moreover, the domination of V(P31)V(P32) and V(P3m1)V(P3m) needs at least two further vertices. Hence, we get γcg(P3Pm) ≥ 2(m – 2) + 2 = 2m – 2 and the theorem follows.□

We remark in passing that γcg(P3P3) = 3, an optimal first move of Dominator being the vertex of P3P3 of degree 4.

Theorem 4.2

If m ≥ 4, then

γcg(P3Cm)=2m2.

Proof

Let m ≥ 4 and let V(Cm) = [m]. All the exponents of the P3-layers will be taken modulo m.

First, we show that γcg(P3Cm) ≤ 2m – 2. The basic strategy of Dominator is to play a legal move in the center of some P3-layer. Clearly this is possible in the D-game. Inductively, let P3i,,P3j,ij, be the P3-layers in which at least one vertex has been played so far. We may assume without loss of generality that Staller’s last move was in P3j . If this move was in the center of P3j , then Dominator replies with the move on the center of P3j+1 . Otherwise, Dominator selects the center of P3j . After m – 1 moves of Dominator we have two cases. If all the P3-layers were played, then by the strategy of Dominator, the whole graph is dominated, hence at most 2m – 3 moves were played. Otherwise, one P3-layer has not been played, say P3j . In this case Dominator has played all the centers of the remaining m – 1 P3-layers. But then Staller has played at least one non-center vertex from the layers P3j1 and P3j+1 . Consequently, at most one vertex from P3j is not yet dominated after the (m – 1)st move of Dominator, which means that in the next move Staller must finish the game. Hence the game finishes in no more than 2m – 2 moves.

Proving that γcg(P3Cm) ≥ 2m – 2 goes along the same lines as in the proof for P3Pm in Theorem 4.1, where the strategy of Staller is modified as in the proof of Theorem 3.2.□

The upper bound in Theorem 4.2 can also be proven using the game with Chooser as it is done in the proof of Theorem 4.1. We have given an alternative proof instead, because it yields an exact strategy of Dominator.

We have thus seen that in many cases the upper bound of Theorem 2.2 is sharp or close to be sharp. On the other hand, the bound can also be arbitrary weak. For this sake let G be a graph with γc(G) ≫ n(G)/2 and consider GG. Then a rough estimate for the the right hand side of the inequality of Theorem 2.2 is that it is ≫ n(G)2 = n(GG).

5 Products with complete graphs

A large class of graphs which are extremal with respect to the general upper bound in Theorem 2.2 can be obtained when we consider the Cartesian products KkG with sufficiently large k. In fact, for every graph G, there exists a threshold k0(G) such that kk0(G) ensures

γcg(KkG)=2n(G)1=min{2γc(G)n(Kk),2γc(Kk)n(G)}1.

It is not difficult to prove that k0(G) = 2n(G) – 1 is always appropriate, see the proof of Theorem 5.1. On the other hand, we also establish a threshold k0(G) in terms of the maximum degree Δ(G) and independence number α(G) of G.

Theorem 5.1

If G is a connected graph and k ≥ min {4Δ(G) + α(G), 2n(G) – 1}, then

γcg(KkG)=2n(G)1.

Proof

Let G be a connected graph on n vertices with maximum degree Δ and with independence number α. During the game, a Kk-layer will be called played if at least one vertex has been already played from it. A Kk-layer will be called dominated if all of its vertices have been dominated, otherwise, it is undominated. Clearly, if a Kk-layer is played, then it is dominated. Observe that, by Theorem 2.2, we have γcg(KkG) ≤ 2n – 1.

First suppose that k ≥ 2n – 1 and consider the following simple strategy of Staller: she always chooses a (legal) vertex from a dominated Kk-layer. While the game is not over, Staller can follow this strategy. Indeed, by the connectivity of G, if there is an undominated Kk-layer in KkG, then we have two neighboring layers Kkg and Kkg (i.e., gg′ ∈ E(G)) such that Kkg is dominated but a vertex (h, g′) ∈ V( Kkg ) is undominated. Then, Staller may play (h, g). Hence, after the (n – 1)st move of Staller, at most n – 1 Kk-layers are played. Each remaining unplayed layer is certainly not dominated since it would need at least k ≥ 2n – 1 vertices played in the neighboring layers. This proves γcg(KkG) ≥ 2n – 1 under the condition k ≥ 2n – 1.

Now, suppose that k ≥ 4Δ + α and consider a connected domination game on KkG. We assign blue, green or red color to those vertices which were played in the game. If Dominator plays a vertex di = v from a Kk-layer such that the layer was undominated before his move, the vertex di is colored blue. If di is blue and Staller plays si from the same Kk-layer as her next move, then si is blue, otherwise si is green. If Dominator plays di from a dominated layer, then both di and si are red. We say that a Kk-layer is a blue layer if it contains a blue vertex.[1] A Kk-layer is white if it remains unplayed at the end of the game and also if it becomes dominated before it is played. When the game finishes, each Kk-layer is either blue or white and, denoting the number of blue or white layers by b or w respectively, we have b + w = n.

Consider the following strategy of Staller: After Dominator’s move di, Staller plays a legal vertex from the same Kk-layer if possible; otherwise, Staller plays a legal vertex from any dominated layer. As we have seen earlier, the latter choice is always possible (if the game is not over) as there must be an undominated layer which has a neighboring dominated layer.

Now, we prove the following claims.

Claim 1

At the end of the game, the number of green vertices is at most α.

Proof

Suppose that si, which is the ith move of Staller, is a green vertex. Then di, the previous move of Dominator, was the first vertex played from the undominated layer Kkg and no further vertex could be played from Kkg after his move. The latter property means that every neighboring layer was already dominated (either played or not) right after the move di. Consequently, no blue vertex can be played from the neighboring layers in the game after di. Therefore, if every green vertex sj of the product is associated with the vertex f(sj) from V(G) such that djV( Kkf(sj) ) holds for the previous move of Dominator, the set F = {f(sj) : sj is green} is an independent vertex set in G and |F| ≤ α follows.(◻)

Claim 2

If Kkg is a white layer, then at least 2Δ red vertices were played from the neighboring layers.

Proof

Since Kkg is white, its k vertices are dominated by k different (blue, green, red) vertices of the neighboring layers. (This remains true, even if later some vertices from Kkg are played.) By definition, every neighboring Kk-layer may contain at most two blue vertices and there are at most Δ neighboring layers. By Claim 1, there exist at most α green vertices. Hence, the number of red vertices in the neighboring layers is at least k – 2Δα. Since k ≥ 4Δ + α, the number of such red vertices is at least 2Δ.(◻)

Now, denote by r and p the number of red vertices and played vertices, respectively, at the end of the game. The number of edges between the red vertices and the white layers is at least 2Δw by Claim 2. On the other hand, since Δ is the maximum vertex degree, the number of these edges is at most Δr. This yields 2wr. Moreover, pr, which is the total number of blue and green vertices, equals either 2b or 2b – 1. (The latter occurs if Dominator finishes the game by playing a blue vertex.) Hence, 2b – 1 ≤ pr. We have the following inequalities:

2b1prp2w

and, since b + w = n, we may conclude 2n – 1 ≤ p. This finishes the proof of γcg(KkG) = 2n(G) – 1 under the condition k ≥ 4Δ + α.□

In the following theorem we prove that if γcg(Kk0G) = 2n(G) – 1 holds for some k0, then γcg(KkG) = 2n(G) – 1 holds for every kk0. This result might appear obvious, however, it does not generalize to general subgraphs. If G is a graph with a large γcg(G) and H is a graph obtained from G by adding a universal vertex, then it holds 1 = γcg(H) ≪ γcg(G), even though GH.

Theorem 5.2

If n ≥ 1, then

γcg(Kn+1G)γcg(KnG).

Proof

We prove the theorem using imagination strategy. Two games will be played in the proof, one on Kn+1G and one on KnG. In order to avoid confusion between vertices of those two graphs, we introduce the notation x[n] = {x1, …, xn}.

The real game is played on Kn+1G with vertex set v[n+1] × V(G), while Staller imagines a game on KnG with vertex set u[n] × V(G). Define a function f : v[n] × V(G) → u[n] × V(G) by f : (vi, g) ↦ (ui, g) for every i ∈ [n] and gV(G), with the inverse f–1 : (ui, g) ↦ (vi, g). Even though f(vn+1, g) is not formally defined, we will simply say that it is not a legal move in the imagined game on KnG. Additionally, in the real game the label vn+1 is given to a vertex of Kn+1 such that the underlying G-layer was played last among all G-layers (or not played at all).

Dominator plays optimally in the real game and Staller plays optimally in the imagined game. Let DR and DI denote the set of already played vertices in the real and in the imagined game, respectively. Staller’s strategy is to ensure that until the imagined games ends it holds (i) |DR| = |DI| and (ii) f(N[DR] ∩ (v[n] × V(G))) ⊆ N[DI]. If this is true, then the number of moves needed to finish the real game is at least the number of moves needed in the imagined game. As Staller plays optimally in the imagined game and Dominator plays optimally in the real game, we have γcg(KnG) ≤ γcg(Kn+1G).

Consider the strategy of Staller. She always replies optimally in the imagined game and tries to copy her move to the real game. By the property (ii), this move surely dominates new vertices in the real game, but might not be connected to a previous move in the real game. Let x = si be the optimal move of Staller in the imagined game. If f–1(x) is a legal move in the real game, then Staller copies it to the real game (clearly preserving properties (i) and (ii)). But if f–1(x) is not a legal move in the real game, then let y be the move in the imagined game that first dominated x. Now set x = y and iteratively repeat the above procedure until Staller reaches a vertex x such that f–1(x) is a legal move in the real game. Clearly, the procedure stops (as the graph is finite) and properties (i) and (ii) are preserved.

Let di = (vk, g), k ∈ [n + 1], gV(G), be an (optimal) move of Dominator in the real game. Staller copies his move to the imagined game by the following rules.

  1. If f(di) is a legal move in the imagined game, then Staller imagines Dominator played f(di) = (uk, g) in the imagined game.

  2. If f(di) is not a legal move, but there exists a playable vertex p in Kng in the imagined game, then Staller imagines Dominator played p.

  3. If neither (D1) nor (D2) can be applied, then Staller imagines Dominator played any legal move in the imagined game.

All the rules clearly preserve (i) and the rule (D1) also preserves (ii).

If k ∈ [n], but f(di) is not a legal move, then either f(di) dominates no new vertex in the imagined game or it is not connected to a previously played vertex. In the first case, the property (ii) still holds after Staller imagines the ith move of Dominator by the rule (D2) or (D3). We now explain that the second case is not possible. Let x be the already played neighbor of di in the real game. By the property (ii), N[x] and f(N[x] ∩ (v[n] × V(G))) are both dominated, in particular, di and f(di) are both dominated. Thus a neighbor of f(di) was already played in the imagined game.

Next, consider the case k = n + 1. By the definition of the vertex vn+1, the move di cannot be the first move of the game. By induction on the number of moves on vn+1G, we will prove that the property (ii) is preserved and also that if Staller cannot apply (D2) at Dominator’s move di, then the whole set u[n] × N[g] is already dominated in the imagined game.

Firstly, consider the first move di0 = (vn+1, g0) played on vn+1G in the real game. Property (ii) is preserved no matter where Staller imagines his move. Suppose that Staller cannot apply (D2) after the move di0. If a vertex was already played in Kng0 in the imagined game, then Staller cannot apply (D2) only if u[n] × N[g0] is already dominated (otherwise she could play another move on Kng0 ). So from now on, suppose also that no vertex was already played in Kng0 in the imagined game. As di0 is a legal move of Dominator in the real game, it is adjacent to an already played vertex x = (vk, g0), k ∈ [n], in the real game. Then x was a move of Dominator in the real game, N[(uk, g0)] is already dominated in the imagined game, in particular Kng0 is already dominated. Thus all vertices in Kng0 are adjacent to an already played vertex in the imagined game. As none of them is a legal move, it means that u[n] × N[g0] is already dominated.

Next, consider the case when di is some later move on vn+1G. If Staller can apply (D2), then the property (ii) is clearly satisfied. Suppose now that Staller cannot apply (D2). If di is adjacent to an already played vertex in v[n] × {g} in the real game, then the proof goes along the same lines as above for di0. So the only remaining case is if di is not adjacent to an already played vertex in v[n] × {g} in the real game. This means it is adjacent to an already played vertex d′ = (vn+1, g′), gg′ ∈ V(G), in the real game and no vertex was played so far in v[n] × {g} in the real game. We distinguish two cases.

The first case is that di dominates no new vertex in v[n] × {g}. Then this set is already dominated in the real game and u[n] × {g} is dominated in the imagined game, thus the connectedness condition is satisfied for every vertex in u[n] × {g}. As Staller cannot apply (D2), u[n] × N[g] is already dominated and property (ii) holds.

The second case is that di dominates a new vertex in v[n] × {g}. If u[n] × {g} is already dominated in the imagined game, then Staller could apply rule (D2), unless u[n] × N[g] is already dominated. Otherwise, there is an undominated vertex in u[n] × {g} in the imagined game. This means that a move (ul, g′) was played in the imagined game on u[n] × {g′} (otherwise u[n] × {g} ⊆ u[n] × N[g′] would be already dominated). But then Staller can imagine Dominator plays (ul, g) in the imagined game to dominate the undominated vertex in u[n] × {g}, hence applying (D2).□

6 A lower bound

In [13] an upper bound but no lower bound is given on γcg(GKn). To give (general) lower bounds on γcg(GH) appears more difficult than giving upper bounds. An intuitive reason for this is the intrinsic structure of the Cartesian product that, for instance, makes a possible proof of the lower bound on the domination number of GH as suggested by Vizing a notorious open problem [22]. Nonetheless, we can give the following general lower bound.

Theorem 6.1

If G and H are connected graphs on at least two vertices, then

γcg(GH)2γc(G);n(H)=2,2γc(G)+1;n(H)3.

Proof

The vertices of V(GH) will be denoted as {(g, h); gG, hH}. Let Ai (resp. Ai ) denote the set of all gG such that at least one vertex in the layer gH has been played within moves d1, s1, …, di (resp. d1, s1, …, di, si). Let BiAi be the smallest connected subset of Ai that dominates N[Ai]. We similarly define Bi Ai . Note that at each step (and at the end of the game) it holds |Ai| ≥ |Bi| and | Ai | ≥ | Bi |.

Staller’s strategy is to play si on a vertex above Bi, meaning that siBi × V(H). If this is not possible, she plays any legal move. We prove that this strategy assures that after each move of Staller it holds i ≥ | Bi |.

As n(H) ≥ 2, Staller can make her first move in the same H-layer where Dominator played, thus 1 ≥ | B1 |. Suppose the property holds for the first i – 1 moves of Staller and consider the situation after Dominator’s ith move, say di = (g, h). We now distinguish two cases.

  1. The vertex di newly dominates only vertices above Ai (i.e. diAi × V(H)).

    In this case, g is not added to Bi, so Bi = Bi1 . No matter where Staller replies, we have |Bi| ≤ | Bi | ≤ |Bi| + 1 = | Bi1 | + 1. But then i – 1 ≥ | Bi1 | ≥ | Bi | – 1, which implies i ≥ | Bi |.

  2. The vertex di dominates at least one vertex not above Ai, say a vertex (g′, h).

    1. Bi1 dominates N[Ai].

      In this case Bi = Bi1 and i ≥ | Bi | as in the previous case.

    2. Bi1 does not dominate N[Ai].

      This means that at least the vertex g′ ∈ N[Ai] is not dominated by Bi1 , so no vertex in gH, except for (g′, h), is dominated after Dominator’s ith move. We have Bi = Bi1 ∪ {g} and di is the only already played vertex in gH. As n(H) ≥ 2, Staller can reply in the same H-layer to newly dominate a vertex in gH. Thus | Bi | = |Bi| = | Bi1 | + 1 ≤ (i – 1) + 1 = i.

If Dominator is the one who finishes the game in his jth move, then his last move is not as in the case 2.(b) above (otherwise Staller has a legal reply and the game is not over). Thus j – 1 ≥ | Bj1 | = |Bj| ≥ γc(G), which means jγc(G) + 1. Hence the number of moves is at least j + (j – 1) ≥ 2 γc(G) + 1.

On the other hand, if Staller ends the game with her jth move, then her strategy assures that j ≥ | Bj | ≥ γc(G), thus the number of moves is at least 2 γc(G). This completes the proof in the case n(H) = 2.

If n(H) ≥ 3, consider again the situation where Staller ended the game. {We distinguish two cases. If the last vertex was added to the set Bj not in the last two moves of the game but earlier, then clearly the number of moves is greater than 2 γc(G) + 1. Otherwise, in the last two steps, one of the players added the last vertex to the set Bj . If the move dj added the last vertex to Bj = Bj, then this situation is as in case 2.(b). Thus after the moves dj, sj, at most 2 < n(H) vertices in the layer gH are dominated. But then the game is not yet finished and at least one more move is needed. On the other hand, if the move sj added a vertex to Bj Bj (and it follows from the above cases that this can only occur if Dominator did not add a vertex to Bj), then in the newly dominated H-layer only 1 < n(H) vertex is dominated, hence at least one more move is needed. It follows that if n(H) ≥ 3, the game ends after at least 2 γc(G) + 1 moves.□

If G and H are graphs of order at least 3, then Theorem 6.1 can be rephrased as follows:

γcg(GH)max{2γc(G),2γc(H)}+1.

Particular examples on which the bound of Theorem 6.1 is sharp are P3P3, C3K2, and C3C3. The bound is also sharp for PmK2 and PmK3, m ≥ 4 [13,Theorem 7], and for Circular ladders CmK2, m ≥ 4 [20, Theorem 5.3]. Using similar reasoning as in the proof of [13, Theorem 7], we obtain another equality case

γcg(CmK3)=3;m=3,2m3;m4.

7 Concluding remarks

As far as we can recall, the (total) domination game was studied on the Cartesian product only in [1], where a connection with Vizing’s conjecture was established and a lower bound in terms of 2-packings proved; in [23] where Cartesian products of complete graphs were investigated; and in [24], where (with a considerable effort) the total domination game critical graphs were characterized among the products K2Cn. Relating this small number of known results to the waste bibliography on the (total) domination game indicates that the game is intrinsically difficult on the Cartesian product. On the other hand, the results from this paper as well as from [13, 20] demonstrate that on Cartesian products it is somehow easier to investigate the connected domination game. An intuitive reason for this could be that the players have more restrictive possibilities for their moves than in the standard game. This intuition does not hold in general when comparing the connected domination game with the domination game. For instance and as already noted earlier, the domination game admits Continuation Principle while the connected domination game does not.

The results of this paper lead to several open problems, we list some of them.

Problem 7.1

Find additional infinite families of graphs which attain the upper bound of Theorem 2.1.

In view of Theorems 4.1 and 4.2 we pose:

Problem 7.2

Determine γcg(PnPm), γcg(PnCm), and γcg(CnCm).

For Theorem 5.1 we strongly feel that the condition 4Δ + α is replaceable with some smaller value. More precisely:

Problem 7.3

Does there exist a constant c ∈ ℝ+, such that Theorem 5.1 remains valid provided that 4Δ + α is replaced with ?

If the answer is positive for Problem 7.3, one may be interested in finding the smallest coefficient c = c(𝓖) over the class 𝓖 of all connected graphs or over any specified subclass of it. In view of the monotonicity stated in Theorem 5.2, this question can be formulated as follows.

Problem 7.4

Given a graph class 𝓖 that does not contain disconnected graphs, determine the constant

c(G)=supGGminkΔ(G):γcg(GKk)=2n(G)1,kZ+.

For the lower bound of Theorem 6.1 we do not know many sharpness examples, hence we also pose:

Problem 7.5

Do there exist families of graphs for which the lower bound of Theorem 6.1 is sharp and none of the factors is K2 or K3?

Finally, in this paper we have focused on the D-game played on Cartesian products. It seems likewise interesting to investigate the S-game on Cartesian products.

Acknowledgements

We acknowledge the financial support from the Slovenian Research Agency (research core funding No. P1-0297 and projects J1-9109, J1-1693, N1-0095, N1-0108). Pakanun Dokyeesun was also supported by the Development and Promotion of Science and Technology Talents Project (DPST) from Thailand.

References

[1] Brešar B., Klavžar S., Rall D.F., Domination game and an imagination strategy, SIAM J. Discrete Math., 2010, 24, 979–991.10.1137/100786800Search in Google Scholar

[2] Bujtás Cs., On the game total domination number, Graphs Combin., 2018, 34, 415–425.10.1007/s00373-018-1883-ySearch in Google Scholar

[3] Henning M.A., Klavžar S., Rall D.F., Total version of the domination game, Graphs Combin., 2015, 31, 1453–1462.10.1007/s00373-014-1470-9Search in Google Scholar

[4] Henning M.A., Klavžar S., Rall D.F., The 4/5 upper bound on the game total domination number, Combinatorica, 2017, 37, 223–251.10.1007/s00493-015-3316-3Search in Google Scholar

[5] Henning M.A., Rall D.F., Progress towards the total domination game 3/4-conjecture, Discrete Math., 2016, 339, 2620–2627.10.1016/j.disc.2016.05.014Search in Google Scholar

[6] Bujtás Cs., Tuza Zs., The disjoint domination game, Discrete Math., 2016, 339, 1985–1992.10.1016/j.disc.2015.04.028Search in Google Scholar

[7] Bujtás Cs., Henning M.A., Tuza Zs., Transversal game on hypergraphs and the 34 -conjecture on the total domination game, SIAM J. Discrete Math., 2016, 30, 1830–1847.10.1137/15M1049361Search in Google Scholar

[8] Bujtás Cs., Henning M.A., Tuza Zs., Bounds on the game transversal number in hypergraphs, European J. Combin., 2017, 59, 34–50.10.1016/j.ejc.2016.07.003Search in Google Scholar

[9] Duchêne E., Gledel V., Parreau A., Renault G., Maker-Breaker domination game, arXiv:1807.09479.10.1016/j.disc.2020.111955Search in Google Scholar

[10] Gledel V., Iršič V., Klavžar S., Maker-Breaker domination number, Bull. Malays. Math. Sci. Soc., 2019, 42, 1773–1789.10.1007/s40840-019-00757-1Search in Google Scholar

[11] Gledel V., Henning M.A., Iršič V., Klavžar S., Maker-Breaker total domination game, arXiv:1902.00204.10.1016/j.dam.2019.11.004Search in Google Scholar

[12] Brešar B., et al., The variety of domination games, Aequat. Math., 2019, https://doi.org/10.1007/s00010-019-00661-w.10.1007/s00010-019-00661-wSearch in Google Scholar

[13] Borowiecki M., Fiedorowicz A., Sidorowicz E., Connected domination game, Appl. Anal. Discrete Math., 2019, 13, 261–289.10.2298/AADM171126020BSearch in Google Scholar

[14] Kinnersley W.B., West D.B., Zamani R., Extremal problems for game domination number, SIAM J. Discrete Math., 2013, 27, 2090–2107.10.1137/120884742Search in Google Scholar

[15] Brešar B., Dorbec P., Klavžar S., Košmrlj G., Domination game: effect of edge- and vertex-removal, Discrete Math., 2014, 330, 1–10.10.1016/j.disc.2014.04.015Search in Google Scholar

[16] Dorbec P., Košmrlj G., Renault G., The domination game played on unions of graphs, Discrete Math., 2015, 338, 71–79.10.1016/j.disc.2014.08.024Search in Google Scholar

[17] James T., Klavžar S., Vijayakumar A., The domination game on split graphs, Bull. Aust. Math. Soc., 2019, 99, 327–337.10.1017/S0004972718001053Search in Google Scholar

[18] Košmrlj G., Domination game on paths and cycles, Ars Math. Contemp., 2017, 13, 125–136.10.26493/1855-3974.891.e93Search in Google Scholar

[19] Xu K., Li X., On domination game stable graphs and domination game edge-critical graphs, Discrete Appl. Math., 2018, 250, 47–56.10.1016/j.dam.2018.05.027Search in Google Scholar

[20] Iršič V., Connected domination game: predomination, Staller-start game, and lexicographic products, arXiv:1902.02087.Search in Google Scholar

[21] Imrich W., Klavžar S., Rall D.F., Topics in Graph Theory: Graphs and their Cartesian Product, A K Peters, Ltd., Wellesley, MA, 2008.10.1201/b10613Search in Google Scholar

[22] Brešar B., et al., Vizing’s conjecture: a survey and recent results, J. Graph Theory, 2012, 69, 46–76.10.1002/jgt.20565Search in Google Scholar

[23] Brešar B., Dorbec P., Klavžar S., Košmrlj G., How long can one bluff in the domination game?, Discuss. Math. Graph Theory, 2017, 37, 337–352.10.7151/dmgt.1899Search in Google Scholar

[24] Henning M.A., Klavžar S., Infinite families of circular and Möbius ladders that are total domination game critical, Bull. Malays. Math. Sci. Soc., 2018, 41, 2141–2149.10.1007/s40840-018-0635-8Search in Google Scholar

Received: 2019-04-12
Accepted: 2019-09-30
Published Online: 2019-11-08

© 2019 Bujtás et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 Public License.

Articles in the same Issue

  1. Regular Articles
  2. On the Gevrey ultradifferentiability of weak solutions of an abstract evolution equation with a scalar type spectral operator of orders less than one
  3. Centralizers of automorphisms permuting free generators
  4. Extreme points and support points of conformal mappings
  5. Arithmetical properties of double Möbius-Bernoulli numbers
  6. The product of quasi-ideal refined generalised quasi-adequate transversals
  7. Characterizations of the Solution Sets of Generalized Convex Fuzzy Optimization Problem
  8. Augmented, free and tensor generalized digroups
  9. Time-dependent attractor of wave equations with nonlinear damping and linear memory
  10. A new smoothing method for solving nonlinear complementarity problems
  11. Almost periodic solution of a discrete competitive system with delays and feedback controls
  12. On a problem of Hasse and Ramachandra
  13. Hopf bifurcation and stability in a Beddington-DeAngelis predator-prey model with stage structure for predator and time delay incorporating prey refuge
  14. A note on the formulas for the Drazin inverse of the sum of two matrices
  15. Completeness theorem for probability models with finitely many valued measure
  16. Periodic solution for ϕ-Laplacian neutral differential equation
  17. Asymptotic orbital shadowing property for diffeomorphisms
  18. Modular equations of a continued fraction of order six
  19. Solutions with concentration and cavitation to the Riemann problem for the isentropic relativistic Euler system for the extended Chaplygin gas
  20. Stability Problems and Analytical Integration for the Clebsch’s System
  21. Topological Indices of Para-line Graphs of V-Phenylenic Nanostructures
  22. On split Lie color triple systems
  23. Triangular Surface Patch Based on Bivariate Meyer-König-Zeller Operator
  24. Generators for maximal subgroups of Conway group Co1
  25. Positivity preserving operator splitting nonstandard finite difference methods for SEIR reaction diffusion model
  26. Characterizations of Convex spaces and Anti-matroids via Derived Operators
  27. On Partitions and Arf Semigroups
  28. Arithmetic properties for Andrews’ (48,6)- and (48,18)-singular overpartitions
  29. A concise proof to the spectral and nuclear norm bounds through tensor partitions
  30. A categorical approach to abstract convex spaces and interval spaces
  31. Dynamics of two-species delayed competitive stage-structured model described by differential-difference equations
  32. Parity results for broken 11-diamond partitions
  33. A new fourth power mean of two-term exponential sums
  34. The new operations on complete ideals
  35. Soft covering based rough graphs and corresponding decision making
  36. Complete convergence for arrays of ratios of order statistics
  37. Sufficient and necessary conditions of convergence for ρ͠ mixing random variables
  38. Attractors of dynamical systems in locally compact spaces
  39. Random attractors for stochastic retarded strongly damped wave equations with additive noise on bounded domains
  40. Statistical approximation properties of λ-Bernstein operators based on q-integers
  41. An investigation of fractional Bagley-Torvik equation
  42. Pentavalent arc-transitive Cayley graphs on Frobenius groups with soluble vertex stabilizer
  43. On the hybrid power mean of two kind different trigonometric sums
  44. Embedding of Supplementary Results in Strong EMT Valuations and Strength
  45. On Diophantine approximation by unlike powers of primes
  46. A General Version of the Nullstellensatz for Arbitrary Fields
  47. A new representation of α-openness, α-continuity, α-irresoluteness, and α-compactness in L-fuzzy pretopological spaces
  48. Random Polygons and Estimations of π
  49. The optimal pebbling of spindle graphs
  50. MBJ-neutrosophic ideals of BCK/BCI-algebras
  51. A note on the structure of a finite group G having a subgroup H maximal in 〈H, Hg
  52. A fuzzy multi-objective linear programming with interval-typed triangular fuzzy numbers
  53. Variational-like inequalities for n-dimensional fuzzy-vector-valued functions and fuzzy optimization
  54. Stability property of the prey free equilibrium point
  55. Rayleigh-Ritz Majorization Error Bounds for the Linear Response Eigenvalue Problem
  56. Hyper-Wiener indices of polyphenyl chains and polyphenyl spiders
  57. Razumikhin-type theorem on time-changed stochastic functional differential equations with Markovian switching
  58. Fixed Points of Meromorphic Functions and Their Higher Order Differences and Shifts
  59. Properties and Inference for a New Class of Generalized Rayleigh Distributions with an Application
  60. Nonfragile observer-based guaranteed cost finite-time control of discrete-time positive impulsive switched systems
  61. Empirical likelihood confidence regions of the parameters in a partially single-index varying-coefficient model
  62. Algebraic loop structures on algebra comultiplications
  63. Two weight estimates for a class of (p, q) type sublinear operators and their commutators
  64. Dynamic of a nonautonomous two-species impulsive competitive system with infinite delays
  65. 2-closures of primitive permutation groups of holomorph type
  66. Monotonicity properties and inequalities related to generalized Grötzsch ring functions
  67. Variation inequalities related to Schrödinger operators on weighted Morrey spaces
  68. Research on cooperation strategy between government and green supply chain based on differential game
  69. Extinction of a two species competitive stage-structured system with the effect of toxic substance and harvesting
  70. *-Ricci soliton on (κ, μ)′-almost Kenmotsu manifolds
  71. Some improved bounds on two energy-like invariants of some derived graphs
  72. Pricing under dynamic risk measures
  73. Finite groups with star-free noncyclic graphs
  74. A degree approach to relationship among fuzzy convex structures, fuzzy closure systems and fuzzy Alexandrov topologies
  75. S-shaped connected component of radial positive solutions for a prescribed mean curvature problem in an annular domain
  76. On Diophantine equations involving Lucas sequences
  77. A new way to represent functions as series
  78. Stability and Hopf bifurcation periodic orbits in delay coupled Lotka-Volterra ring system
  79. Some remarks on a pair of seemingly unrelated regression models
  80. Lyapunov stable homoclinic classes for smooth vector fields
  81. Stabilizers in EQ-algebras
  82. The properties of solutions for several types of Painlevé equations concerning fixed-points, zeros and poles
  83. Spectrum perturbations of compact operators in a Banach space
  84. The non-commuting graph of a non-central hypergroup
  85. Lie symmetry analysis and conservation law for the equation arising from higher order Broer-Kaup equation
  86. Positive solutions of the discrete Dirichlet problem involving the mean curvature operator
  87. Dislocated quasi cone b-metric space over Banach algebra and contraction principles with application to functional equations
  88. On the Gevrey ultradifferentiability of weak solutions of an abstract evolution equation with a scalar type spectral operator on the open semi-axis
  89. Differential polynomials of L-functions with truncated shared values
  90. Exclusion sets in the S-type eigenvalue localization sets for tensors
  91. Continuous linear operators on Orlicz-Bochner spaces
  92. Non-trivial solutions for Schrödinger-Poisson systems involving critical nonlocal term and potential vanishing at infinity
  93. Characterizations of Benson proper efficiency of set-valued optimization in real linear spaces
  94. A quantitative obstruction to collapsing surfaces
  95. Dynamic behaviors of a Lotka-Volterra type predator-prey system with Allee effect on the predator species and density dependent birth rate on the prey species
  96. Coexistence for a kind of stochastic three-species competitive models
  97. Algebraic and qualitative remarks about the family yy′ = (αxm+k–1 + βxmk–1)y + γx2m–2k–1
  98. On the two-term exponential sums and character sums of polynomials
  99. F-biharmonic maps into general Riemannian manifolds
  100. Embeddings of harmonic mixed norm spaces on smoothly bounded domains in ℝn
  101. Asymptotic behavior for non-autonomous stochastic plate equation on unbounded domains
  102. Power graphs and exchange property for resolving sets
  103. On nearly Hurewicz spaces
  104. Least eigenvalue of the connected graphs whose complements are cacti
  105. Determinants of two kinds of matrices whose elements involve sine functions
  106. A characterization of translational hulls of a strongly right type B semigroup
  107. Common fixed point results for two families of multivalued A–dominated contractive mappings on closed ball with applications
  108. Lp estimates for maximal functions along surfaces of revolution on product spaces
  109. Path-induced closure operators on graphs for defining digital Jordan surfaces
  110. Irreducible modules with highest weight vectors over modular Witt and special Lie superalgebras
  111. Existence of periodic solutions with prescribed minimal period of a 2nth-order discrete system
  112. Injective hulls of many-sorted ordered algebras
  113. Random uniform exponential attractor for stochastic non-autonomous reaction-diffusion equation with multiplicative noise in ℝ3
  114. Global properties of virus dynamics with B-cell impairment
  115. The monotonicity of ratios involving arc tangent function with applications
  116. A family of Cantorvals
  117. An asymptotic property of branching-type overloaded polling networks
  118. Almost periodic solutions of a commensalism system with Michaelis-Menten type harvesting on time scales
  119. Explicit order 3/2 Runge-Kutta method for numerical solutions of stochastic differential equations by using Itô-Taylor expansion
  120. L-fuzzy ideals and L-fuzzy subalgebras of Novikov algebras
  121. L-topological-convex spaces generated by L-convex bases
  122. An optimal fourth-order family of modified Cauchy methods for finding solutions of nonlinear equations and their dynamical behavior
  123. New error bounds for linear complementarity problems of Σ-SDD matrices and SB-matrices
  124. Hankel determinant of order three for familiar subsets of analytic functions related with sine function
  125. On some automorphic properties of Galois traces of class invariants from generalized Weber functions of level 5
  126. Results on existence for generalized nD Navier-Stokes equations
  127. Regular Banach space net and abstract-valued Orlicz space of range-varying type
  128. Some properties of pre-quasi operator ideal of type generalized Cesáro sequence space defined by weighted means
  129. On a new convergence in topological spaces
  130. On a fixed point theorem with application to functional equations
  131. Coupled system of a fractional order differential equations with weighted initial conditions
  132. Rough quotient in topological rough sets
  133. Split Hausdorff internal topologies on posets
  134. A preconditioned AOR iterative scheme for systems of linear equations with L-matrics
  135. New handy and accurate approximation for the Gaussian integrals with applications to science and engineering
  136. Special Issue on Graph Theory (GWGT 2019)
  137. The general position problem and strong resolving graphs
  138. Connected domination game played on Cartesian products
  139. On minimum algebraic connectivity of graphs whose complements are bicyclic
  140. A novel method to construct NSSD molecular graphs
Downloaded on 24.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/math-2019-0111/html?lang=en
Scroll to top button