449
Views
6
CrossRef citations to date
0
Altmetric
Original Article

Connected minimum secure-dominating sets in gridsFootnote

, , , , &
Pages 216-223 | Received 24 Feb 2017, Accepted 06 Mar 2017, Published online: 10 Jun 2020

Abstract

For any (finite simple) graph G the secure domination number of G satisfies γs(G)|V(G)|2. Here we find a secure-dominating set S in G such that |S|=|V(G)|2 in all cases when G is a grid, and in the majority of cases when G is a cylindrical or toroidal grid. In all such cases, S satisfies the additional requirement that G[S] is connected. We make note that the concept of secure-dominating sets considered in this paper is quite different from the other secure domination currently of interest.Footnote1

1 Introduction

All graphs will be finite and simple. The vertex and edge sets of a graph G are denoted V(G) and E(G), respectively. An edge joining vertices u and v may be denoted uv or vu. If xV(G), the open neighbor set of x in G is NG(x)={uV(G):uxE(G)}. If G is the only graph in the discussion, the subscript G may be omitted, in this notation. For instance, we define the closed neighborhood of xV(G) to be N[x]=N(x){x}. If SV(G), N(S)=xSN(x) and N[S]=N(S)S. A set SV(G) is dominating in G if and only if N[S]=V(G). A dominating set SV(G) is called a perfect dominating set if every vertex in V(G)S has exactly one neighbor in S.

Fig. 7 A connected secure-dominating set in both P9P10 and P9C10, on 45 vertices. Edges are omitted.
Fig. 8 A connected secure-dominating set on 18 vertices in C6C6. Edges are omitted.
Fig. 9 A connected secure-dominating set on 25 vertices in C7C7. Edges are omitted.
Fig. 10 A connected secure-dominating set C7C9. Edges are omitted.

In an attack on SV(G), in G, each vertex in N[S]S may expend its one unit of aggression by attacking one vertex of S to which it is adjacent (joined to by an edge of G). In a defense of S, each vertex of S may expend its one unit of defense by defending either itself or one of its neighbors in S. A defense of Sthwarts or overcomes an attack on S if and only if each vertex of S has at least as many defenders as attackers. S is secure in G if and only if for every attack on S there is a defense which thwarts that attack.

Suppose XSV(G) and |N(X)S|>|N[X]S|; that is, X has more potential attackers than potential defenders. If every vertex in N(X)S attacks one of its neighbors in X, then no matter how the defensive capabilities of S are allocated, at least one vertex of X will have more attackers than defenders. These observations disclose a necessary condition for security that turns out to be, in the fundamental theorem of security, sufficient.

Theorem 1.1

Brigham, Dutton, and Hedetniemi 2007; see also [Citation1]

Suppose G is a graph. A set SV(G) is secure in G if and only if for any XS, |N[X]S||N(X)S|.

A set SV(G) is secure-dominating if it is both secure and dominating. As observed in [Citation2], if S is secure-dominating in G, then |S||V(G)|2; otherwise, if |S|<|V(G)|2, then |N(S)S|=|V(G)S|>|S|, and so S has more potential attackers than defenders. Therefore, if we define the secure-domination number of G, γs(G), to be the minimum cardinality of a secure-dominating set in G, then γs(G)|V(G)|2.

The interesting evidence from [Citation2] is that for a great many common graphs such as paths, most cycles, complete multipartite graphs, n-cubes, the Petersen graph, and many more, γs(G)=|V(G)|2. The only exceptions to this statement on the list above are the cycles Cn on n2(mod4) vertices (n=6,10,), for which γs(Cn)=n2+1.

A grid is a Cartesian product PmPn of non-trivial (m,n>1) paths; a cylindrical grid is a Cartesian product PmCn of a non-trivial path with a cycle (m>1 and n>2); and a toroidal grid is a Cartesian product CmCn of cycles (m,n>2). We will show that if G is any of the first two types of grid, then γs(G)=V(G)2, and more: a secure-dominating set SV(G) can be found such that |S|=V(G)2 and, in addition, the subgraph induced by S in G, usually denoted G[S], is connected.

In such a case we say that S is a connected secure-dominating set. For arbitrary connected graphs G the minimum cardinality of a connected secure-dominating set is denoted γsc(G). Clearly, γsc(G)γs(G)V(G)2. Our aim here is to show that equality holds throughout, in this last string of inequalities, when G is one of these blankety-blank grids. We have not achieved this aim for the toroidal grids, and perhaps it is not achievable. But γsc(G)=V(G)2 does hold for better than 34 of the toroidal grids, in a sense that will be clear.

For those interested in applications, we note that a connected secure-dominating set is a connected dominating set, which is also a connected hub set [Citation3], and all of these are of interest in street design, transportation, and network communication problems, which often involve grid graphs with mutations (holes, loops, other deformities). Surely, in this day and age, security is a desirable property for connected dominating and connected hub sets. It is shown in [Citation3] that every minimum connected dominating set has at most one vertex more than a minimum connected hub set. The study of connected secure hub sets is wide open, at present. Our results for connected secure dominating sets in grids may be compared with the results in [Citation4], on connected dominating sets in grids.

2 Results

Throughout, the vertices of Pm or Cm will be u1,,um, along the path or around the cycle; the vertices of Pn or Cn will be v1,,vn; the ordered pair (ui,vj), a vertex of PmPn or PmCn or CmCn, will be denoted wi,j.

Theorem 2.1

If m,n2 are integers, then γsc(PmPn)=mn2.

Proof

Since γsc(G)|V(G)|2 for all graphs G, it suffices in each case to find a connected secure dominating set SV(PmPn) such that |S|=|V(G)|2.

Case 1: m=2. Take S={w1,j:j=1,,n}.

Case 2: m=4. Take S={wi,j:i=2,3,j=1,,n}.

In these first two cases, each vertex of V(PmPn)S has exactly one neighbor in S, so S is a perfect dominating set in G. A defensive strategy is easy: each wS defends itself. This opens an interesting question that we will leave for others. Which graphs admit connected secure-dominating sets, which are also perfect dominating sets? In the cases to follow, and in the proofs of Theorems 2.2 and 2.3, it will be self-evident that the proposed connected secure-dominating sets are connected and dominating, but verifying security will be more difficult. Airtight proofs based on appeals to Theorem 1.1 can be provided, but these would be lengthy and unintuitive. We prefer an informal approach, in which a defense thwarting the various possible attacks is indicated informally, and it is left to the reader to verify that no clever attack can be found. The task will be simplified by the circumstance that most of the vertices of V(G)S will have a unique neighbor in S; we may as well assume that, in any attack, each of those vertices will attack its unique neighbor in S. Then the question is, What will the vertices in V(G)S that have a choice do, and can their attack be thwarted? To illustrate, consider

Case 3: m=3,n3. Take S={w2,j:j=1,,n}{w1,j:j odd}. (See .)

The bottom row w3,1,,w3,n are all in V(G)S(G=P3Pn); each w3,j will attack w2,j. Each w1,j, j even, has a choice of 3 vertices of S to attack, unless n is even and j=n, in which case the choice is two vertices.

To see that S is secure, let us consider the first doubtful case, n=4, and some of 6 possible attacks. First, suppose that w1,4 attacks w1,3. If w1,2 does not attack w1,3 then let w1,3, w2,4, and w2,3 defend themselves; we leave it to the reader to verify that S={w1,1,w2,1,w2,2} can then defend itself whether w1,2 attacks w1,1 or w2,2. If, on the other hand w1,2 does attack w1,3, so that w1,3 has 2 attackers, w1,2 and w1,4, then obviously the only possible successful defense will start with w2,4 defending itself, and w1,3 and w2,3 defending w1,3. This leads to w2,2 defending w2,3,w2,1 defending w2,2, and w1,1 defending w2,1, and this defense defeats the attack.

In another attack, suppose that w1,4 attacks w2,4, which then has 2 attackers. The only possible way to save the day is for w2,3 to defend w2,4, which is, of course, defending itself. Then if w1,2 is not attacking w1,3, we simply let if w1,3 defend w2,3, and we are back to a case in which the rest of the attack can be thwarted, regardless of what w1,2 does. On the other hand, if w1,2 attacks w1,3 then w1,3 must defend itself, and, as before, w2,2 defends w2,3,w2,1 defends w2,2, and w1,1 defends w2,1.

We hope the reader, after working through these cases when n=4, will agree that, not only is S secure for all values of n3 (the argument is less delicate when n is odd), but also that there is a relatively simple set of military-style, battlefield instructions for each vertex in S, telling the vertex xS who to defend in each of a small number (6) of cases, defined by how many attackers each vertex in N[x]S has and also, for x=w2,j,1j<n, when x has only one attacker, whether or not w2,j+1 is defending itself, or being defended by w1,j+1(j even), or neither. It would be tedious to prove that such explicit battlefield instructions produce a successful defense, and we are not going to try, but we mention this because clearly such instructions bear on applications: a successful defense can be constructed by the troops by transmission of local intelligence (right-to-left along the columns of the grid, so that w2,j can be apprised of what w2,j+1 has been instructed to do), without each soldier in S having to know anything about the attack around vertices of S a distance >2 in the grid from the soldiers.

By Cases 1–3, we may now assume that both m and n are greater than 4.

Case 4: m,n>4 and m0(mod4): Take S={wj,2:1jm}{wj,n1:1jm}{wi,j:i2,3(mod4),1im, and 3jn2}.

We illustrate in , with m=8 and n=7, and S indicated by darkening n=7.

It is helpful to realize that for any values of m,n>4, m0(mod4), the set S is obtainable by repeating the pattern in the first 4 rows of the m×n grid in each of the m4 blocks of 4 rows each into which the grid can be partitioned.

With this aid to understanding, it is clear that S is connected and dominating, with mn2 vertices. The security of S is much easier to see than the corresponding claim in Case 3; in the battlefield instructions to the vertices of S for defending against an attack, many (must, if n is large) vertices will be instructed to defend themselves, in all circumstances, and the others will have to defend a neighbor under only one local attack variation. For instance, w2,2,w2,3S will each defend themselves unless and only unless w1,3 attacks w1,2, giving w1,2 two attackers. When this happens, w2,2 will defend w1,2 and w2,3 (which has no attacker if w1,3 is attacking w1,2) defends w2,2.

Case 5: m2(mod4),m,n>4: Take S={wi,2:1im}{wi,n1:1im}{wi,j:i2,3(mod4),2im,3jn2}.

This amounts to adding two rows to the pattern for Pm2Pn that look like (with elements of S darkened):

We illustrate with m=6, and n unspecified. (See .)

From Cases 1–5 we are left with the cases m,n5, both odd.

Case 6: m5, m1(mod4),n5, odd: Take S={wi,2:1im}{wi,n1:1im}{wi,j:i1,0(mod4),1im1,3jn2}S0, where S0={wm,j:j0,3(mod4),3jn2} if n1(mod4), and S0={wm,j:j2,3(mod4),3jn4}{wm,n2} if n3(mod4). See .

Case 7: m7, m3(mod4),n5, odd: Take S={wi,2:1im}{wi,n1:1im}{wi,j:i2,3(mod4),1im1,3jn2}S0, where S0 is as in Case 6. See . □

Fig. 1 Case 3. Vertices in S are blacked-in.
Fig. 2 A connected minimum secure dominating set of vertices in P8P7 with 28=8×72 vertices.
Fig. 3 P6Pn, a minimum secure dominating set which is also connected.
Fig. 4 A connected secure dominating set in P9P11.
Fig. 5 A connected secure dominating set in P11P9 with 50=992 vertices.

Theorem 2.2

If m2 and n3 are integers, then γsc(PmCn)=mn2 except, possibly, in the cases m3(mod4), m7, when n2(mod4), and m1(mod4), m9, when n=6.

Proof

Suppose that SV(PmPn) is secure and dominating in PmPn and that, for each i{1,,m},wi,1S if and only if wi,nS. It then follows that S is secure and dominating in PmCn, which is obtained from PmPn by adding the edges wi,1wi,n,i=1,,m; obviously S is dominating in the larger graph, and it is secure because each uS has the same set of potential attackers in PmCn as in PmPn, and its neighbor set in S, in PmCn, contains its neighbor set in S in PmPn; that is, it has the same set of potential attackers and has lost no potential defenders in the transition from PmPn to PmCn.

Therefore the claim of Theorem 2.2 follows from the proof of Theorem 2.1 in all cases except, possibly, in the cases

(i)

m=3, n4, n even;

(ii)

m4, n=3;

(iii)

m odd, m>3, n2(mod4), n6.

The reduction to these cases from the proof of Theorem 2.1 uses, of course, the obvious isomorphism of CnPm and PmCn.

Since we are giving up on the cases m3(mod4) under (iii), and m1(mod4), m9, n=6, we need only verify the claim of the theorem in cases (i), (ii), and (iii): m1(mod4) and n2(mod4), (n10), or m=5 and n=6. For practical purposes it is worth noting that in case (iii) the secure-dominating sets to be given are also secure-dominating in PmPn, adding to the supply of minimum secure-dominating sets in plain grids. The new secure-dominating sets in case (ii), for PmC3, are not dominating for PmP3.

In the cases m=3, n even (n4), it turns out that the secure dominating set in P3Pn given in the proof Theorem 1, Case 3, is also secure and dominating in P3Cn, even though the variety of attacks to be thwarted has increased with the addition of the edges wi,1wi,n,i=1,2,3. As in the proof of Theorem 2.1, we leave it to the reader to verify this claim, most usefully by providing battlefield instructions to the vertices of S for reacting to any attack. Anyone who found such instructions for the P3Pn case (n even) will find that they have not much changed, in fact, they will become simpler, in that instructions to the vertices w2,j,j odd, will be pretty much the same, for different j.

For PmC3, m4, let S be as follows.

(1)

If m0(mod4), take S={wi,1:1im}{wi,2:i2,3(mod4)}.

(2)

If m1(mod4), S={wi,1:1im}{wi,2:i2,3(mod4)}{wm1,2}.

(3)

If m2(mod4), S={wi,1:1im}{wi,2:i2,3(mod4),1im2}{wm,2}.

(4)

If m3(mod4), S={wi,1:1im}{wi,2:i2,3(mod4),1im3}{wm2,2,wm1,2}.

The cases m1(mod4), n2(mod4), m=5 if n=6, divide into 2 subcases.

(1)

m=5, n=6:

S={wi,j:i{2,4},1j6}{w3,2,w3,4,w3,5} (see ).

(2)

m1(mod4), m5, n2(mod4), n10:

S={wi,j:1im,j{2,n1}}{wi,j:i2,3(mod4),2im3,3jn2}{wm1,j:3jn2}{wm2,j:j2,3(mod4),3jn7}{wm2,n4,wm2,n2}. □

Fig. 6 A connected secure-dominating set in both P5P6 and P5C6, on 15 vertices. Edges are omitted.

(See .)

Theorem 2.3

Suppose that m,n3 are integers. Let G=CmCn. Then γsc(G)=mn2 except, possibly, in the following cases:

(1)

{m,n}={6,z}, z1(mod4), z9 ;

(2)

mn2(mod4), and m,n10 ;

(3)

mn3(mod4), and m,n11 ;

(4)

{m,n}={4r+2,4s+3} for some r1, s0.

Proof

Let us say that a pair (m,n) is good if and only if γsc(CmCn)=mn2. Obviously (m,n) is good if and only if (n,m) is good.

As at the beginning of the proof of Theorem 2.2, we observe that if S is a secure-dominating set of vertices in PmCn such that w1,jS if and only if wm,jS then S is also secure-dominating in CmCn. Applying this observation to the secure-dominating sets in the proof of Theorem 2.2, some of which were imported from the proof of Theorem 2.1, we find that (m,n) is good in the following cases,

(i)

{m,n}={3,z}, z4, z2(mod4), Look at the cases PmC3, m4, in the proof of Theorem 2.2.

(ii)

{m,n}={4k,z} for some k1, z4. See the proof of Theorem 2.1 in the cases m=4 and m>4, m0(mod4).

(iii)

{m,n}={4r+1,4s+2}, r,s1, except for {4r+1,6}, r2. In the proof of Theorem 2.2, inspect the cases P5C6 and PmCn, m1(mod4), n2(mod4), n10.

Therefore, to finish the proof, it will suffice to show that the following pairs (m,n) are good:

(a)

(3,3);

(b)

m=6, n2(mod4), n6;

(c)

m=7, n3(mod4), n7;

(d)

m,n>4, m odd, n1(mod4).

Here are the secure-dominating sets that show that (m,n) is a good pair, in these cases.

(a)

m=n=3: S={wi,1:1i3}{w1,2,w2,2}

(b)

m=6, n2(mod4), n6:

S={wi,j:i{2,5},1jn}{wi,j:i{3,4},j2,3(mod4),2jn2,}{w3,n1,w4,n1}

(c)

m=7, n3(mod4), n7:

S={wi,j:i{2,6},1jn}{wi,j:i{3,4,5},j2,3(mod4),1jn1,}{w3,n2,w5,n2}

(d)

m,n>4, m odd, n1(mod4):

S={wi,j:i{2,m1},1jn}{wi,j:3im2,j2,3(mod4),1jn3,}{wi,n1:3im2}{wi,n2:3im2,i odd}.

Determining γsc(CmCn) in the cases left open by Theorem 2.3 is a worthy problem. We remark that it is easy to see that γsc(C3Cn)3n2+1, for all n. In the other cases we can show that γsc(CmCn)mn2+m42, but feel that some clever person will soon surpass this estimate. □

Notes

Peer review under responsibility of Kalasalingam University.

1 The other sort of secure domination: A dominating set SV(G) is secure dominating (in G) if and only if for every vV(G)S there is some uS such that (S{u}){v} is dominating, in G.

References

  • GarthIsaacPeterJohnsonCalebPetrieInteger and fractional security in graphsDiscrete Appl. Math.160201220602062
  • PeterJohnsonCadaviousJonesSecure dominating sets in graphsV.R.KulliAdvances in Domination Theory II2013Vishwa International Publications19 ISBN 81-900205-6-0
  • PeterJohnsonPeterSlaterMatthewWalshThe connected hub number and the connected domination numberNetworks582011171180
  • EgbertMujuniConnected dominating set problem for hypercubes and grid graphiCASTOR J. Math. Sci.7220138189