, H '. _ .~. ._ . . ., _,,,.Ar..r...;...”n.A._;--‘:-x,.'. ..;.‘_..,_;__ ‘_._-.. .- ‘ —‘ . ‘ “ ' —,”‘ , 0 7 lul‘ . ._ .u.’ j. .‘ E" - .f —_;_ GRAPHS AND THEiR SUBDIVISIONS Thesis for the «Degree of Ph. D. WCHIGAN STATE UNIVERSITY M. IAMES STEWART 1972 LIBRARY Michigan State University This is to certify that the thesis entitled Graphs and their Subdivisions presented by M. James Stewart has been accepted towards fulfillment of the requirements for El; . 12, degree in Mathematics XJZM/W Major professor 0-7639 800K NNDERYINB.' ‘ LIBRARY 8mm; R51 SPRINGPORI. MlClllBt-tfli It. I! J- . kn erILIr. ABSTRACT GRAPHS AND THEIR SUBDIVISIONS By M. James Stewart The concept of a subdivision of a graph is an old concept which is widely used in various graph-theoretic constructions and underlies the notion of graph-homeomorphism. A particular type of subdivision of a graph G, denoted by Sn(G) and referred to as the nth subdivision graph of G, is obtained by replacing every edge of G by a path of length n + l. The nth subdivision graph of G is related to three other graph-valued functions: the nth power G", the total graph T(G), and the line graph L(G). Although the graphs G", T(G), and L(G) have received a great deal of attention in the literature, surprisingly little research has been devoted to the investigation of Sn(G). The main purpose of this thesis is to add to the knowledge of subdivisions of graphs and, in particular, to study the properties of the nth subdivision graph and to obtain a characterization of it. The definitions of terms referred to in this thesis are presented in Chapter l, as well as much of the notation that will M. James Stewart be employed throughout. A few well known graph-theoretic results useful in this thesis are also stated here. The first section of Chapter 2 traces the development of the graph-homeomorphism concept from its toplogical origins to the equivalence classes of the relation "is homeomorphic with," where particular attention is paid to the canonical representatives of these classes, the homeomorphically irreducible graphs. In addition to the multiple characterization of these homeomorphically irreducible graphs in terms of paths, triangles, and matrices, in sections 2.2 and 2.3 several graph-theoretic properties, including colorability, traversibility, and automorphism groups, and the way these relate to homeomorphically irreducible graphs, are discussed. In section 2.4 the hamiltonian index of graphs contained in certain homeomorphism classes of nonhamiltonian graphs is investigated. A complete characterization of the nth subdivision graph involving lengths of trails is presented in the first section of Chapter 3, followed by some considerations of the structural rela- tionship of a graph G with its nth subdivision graph Sn(G)‘ Except for section 3.4, which deals with the enumeration of trees having a l-factor, the remainder of this chapter is devoted to the develop- ment of various salient characteristics of nth subdivision graphs. In particular, for any graph G, necessary and sufficient conditions for the existence of a l-factor of Sn(G), Gn (n 3_2), and T(G) are obtained in section 3.3 and generalizations of these results utilizing the edge independence number are discussed in the final section 3.5 GRAPHS AND THEIR SUBDIVISIONS By M. James Stewart A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mathematics 1972 O9 DEDICATION 1 (9 a“ To Gary, Shashi, Tricia's Mother, and the Local Leo. ii ACKNOWLEDGMENTS Words cannot suffice to thank my good friends Gary Chartrand and Shashi Kapoor, who are in large part responsible for giving me the motivation to initiate research of my own in graph theory. A special debt of gratitude is owed to my “academic father," Professor E. A. Nordhaus, without whose inspiration, guidance, and psychological support this thesis could not exist. I thank him for his concern, his knowledge, and the time and effort he has cheerfully spent on my behalf. My gratitude goes also to Professors E. M. Palmer and B. M. Stewart for the helpful suggestions and numerous kindnesses, often unsolicited, they have extended to me. I thank also Professors M. Fuchs and V. P. Sreedharan for serving on my advisory committee. Most importantly, my family has been an unseen but indis- pensable source of strength to me, especially Joyce 1. Stewart. I realize better in retrospect than I did at the time just how much her consideration, understanding, and support has contributed to the accomplishment this dissertation represents. I am very grateful. Over and above all else, I thank my wife, Mary-~without her there would be no point to this endeavor. Finally, I am indebted to the people of the Lansing Com- munity College district for their financial support through a sabbatical leave, and to the Department of Mathematics at Michigan State University for the gracious hospitality I enjoyed throughout the sabbatical year. iv LIST OF FIGURES .................... CHAPTER I ............................ CHAPTER 2 ............................ Section 1.1: Section l.2: Section 2. Section 2. Section 2. Section 2. Section 3. Section 3.2: Section 3.3: Section 3.4: Section 3.5: BIBLIOGRAPHY th—l TABLE OF CONTENTS Homeomorphic Graphs ........... Homeomorphically Irreducible Graphs . . . Some Properties of Irreducible Graphs Certain Homeomorphism Classes and the , Hamiltonian Index ............ CHAPTER 3 ............................ The Nth Subdivision Graph and Its Structure ................ Properties of Nth Subdivision Graphs . . . Factorization in Sn(G) .......... The Enumeration of Trees Having a l-Factor ................. The Edge Independence Number for Sn(G) . . . . Page vi 3O 47 57 59 LIST OF FIGURES Figure Page 2.1 Geometric l-complexes .................. 8 2.2 Homeomorphic graphs ................... 10 2.3 An example illustrating Corollary 2.4 .......... 19 2.4 Homeomorphs H1 and H2 of an irreducible graph G having |F(H])| < |F(G)| < |P(H2)| ........... 2l 2.5 H is randomly hamiltonian from v, but possesses a subdivision G which is not hamiltonian ........ 24 3.1 A multigraph M together with S(M) and 52(M) ....... 30 3.2 An outerplanar graph K4-x and its nonouterplanar subdivision graph .................... 42 3.3 Graphs G such that S(G) has no l-factor ......... 52 3.4 The 10 nonisomorphic rooted trees of order 6 or less having a l-factor ................ 57 3.5 A (6,7)-graph G along with S(G) and 53(6) ........ 63 vi CHAPTER 1 Section l.l Introduction One of the most celebrated theorems in graph theory is Kuratowski's theorem, which characterizes planar graphs as those graphs containing no subgraph isomorphic to a subdivision of the complete graph K5 or to the complete bipartite graph K(3,3). In addition to its central role in Kuratowski's theorem and in numerous other graph-theoretic constructions, the concept of the subdivision of a graph underlies the notion of graph-homeomorphism. This rela- tion is defined more precisely in Chapter 2 and there is shown to be an equivalence relation on the set of all graphs. A graph S(G), referred to as the subdivision graph of G, is also studied and is an example of a graph-valued function which is intimately related to a well known trio of graph-valued functions: the total graph T(G), the line graph L(G), and the nth power G". In l965, M. Behzad showed that the total graph T(G) of a graph G is isomorphic to the square of the subdivision graph S(G). The total graph T(G) of any graph G contains both S(G) and the line graph L(G) as subgraphs. The nth subdivision graph Sn(G), a generalization of S(G), is a graph-valued function which is associated with a graph G and is obtained by replacing every edge of G by a path of length n +l. Although the line graph, total graph, and, to a lesser extent, the nth power of a graph have received much attention in the literature, surprisingly little research has been devoted to the investigation of the nth subdivision graph function. One of the main purposes of this thesis is to study the properties of the nth subdivision graph Sn(G) of a graph G from the viewpoint that the pair [G, Sn(G)] is a specific instance of homeomorphic graphs. Definitions of terms used in this thesis are presented in Chapter l, along with much of the notation that will be employed. A few well known results are also stated here. The first section of Chapter 2 traces the development of the graph-homeomorphism concept from its topological origins to the equivalence classes of the relation "is homeomorphic with," where particular attention is paid to the canonical representatives of these classes, the homeomorphically irreducible graphs. In addition to the multiple characterization of these homeomorphically irreducible graphs in terms of paths, triangles, and matrices, in sections 2.2 and 2.3 several graph-theoretic properties, including colorability, traversibility, and automorphism groups, and the way these relate to homeomorphically irreducible graphs,are discussed. In section 2.4 the hamiltonian index of those graphs in certain homeomorphism classes of nonhamiltonian graphs is investigated. A complete characterization of the nth subdivision graph involving lengths of trails is presented in the first section of Chapter 3, followed by some considerations of the structural relationship of a graph G with its nth subdivision graph Sn(G). Except for section 3.4, which deals with the enumeration of trees having a l-factor, the remainder of this chapter is devoted to the development of various salient characteristics of nth subdivision graphs. In particular, for any graph G, necessary and sufficient conditions for the existence of a l-factor of Sn(G), Gn (n 3_2), and T(G) are obtained in section 3.3 and generalizations of these results utilizing the edge independence number are discussed in the final section 3.4. Section 1.2 Basic Terminology We set forth here some basic definitions and notation which will be used in the following chapters. For those terms not given here, the reader is directed to [2] or [9]. A graph_G is a finite, nonempty set V together with a set E of two-element subsets of V. Each element of V is referred to as a ygrtgx_or a point and V itself as the vertex set or point set of G; the members of the edge set E are called edges, In general, the vertex set and edge set of a graph G will be denoted by V(G) and E(G), respectively. There are variations of graphs to which we shall occasionally make reference. In a multigraph, we allow the presence of more than one edge joining a pair of distinct points; such edges are called multiple edges. If loops are also permitted, that is, edges which join a point to itself, we have a pseudograph. The order of a graph G, denoted [G], is the number of elements in V(G); if |G| = p and E(G) has q elements, we say G is a (p,q)-graph. A graph G is called gmpty_if E(G) is the empty set. The gggggg_of a point v of G is the number of edges of G incident with v and is denoted degGv, or simply deg v, if no confusion is likely to result by leaving the graph G unspecified. A point of degree l is called an endpoint of G. When every point of G has degree r, we say that G is regular of degree r, or r-regular. Two graphs G1 and G2 are called isomorphic, expressed by writing G1 = Gz, if there exists a one-to-one correspon- dence between V(G1) and V(Gz) which preserves adjacency and nonadjacency. The subgraph induced by a set U of points of G, denoted , is that subgraph having U as its point set and whose edge set consists of all edges of G incident with two points of U. In a completely analogous way, we speak of the subgraph induced by a nonempty subset of E(G). If v is a point of G, then G-v denotes the graph and, in general, if S is a proper subset of V(G), then G-S represents the graph . Similarly, G-F represents for any proper subset F of E(G), and, in particular, G-x represents for any edge x of G. A subset N of V(G) is independent if is empty. Two graphs are said to be disjoint if their vertex sets are disjoint. The ggjgg_of two graphs G1 and G2 is the graph G given by V(G) = V(G])\S}V(Gz) and E(G) = E(G])K,)E(GZ). -—For-points u and v of a graph G, a u-v trail is a finite alternating sequence of points and edges of G, in which no edge is repeated, beginning with u and ending with v, such that every edge is immediately preceded and succeeded by the two points with which it is incident (if we allow an edge to be repeated, we have a pplk_from u to v). A u-v path P is a u-v trail in which no point is repeated and the points of P distinct from u and v are called interior points of P. Two u-v paths in a graph G are vertex-disjoint (or simply disjoint) if they have no vertices in common, other than u and v. Edge-disjoint paths have no common edge. We say a u-v trail is glp§2g_or pp§p_depending on whether u = v or u f v, and a trivial trgjl_is one containing no edges. A nontrivial closed trail of G in which no points are repeated (except the first and last) is called a gyglg_of G. The number of edges in a u-v trail is called its lgpgtp, denoted as d(u,v), and a cycle of length n is an n-cycle. A graph of order n which is a path or a cycle is denoted by Pn or Cn, respec- tively. A graph G is connected if there exists a u-v path in G for every two points u and v of G. The relation "is connected to" is an equivalence relation on the vertex set of a graph G and the sub- graphs induced by the points in an equivalence class are called the components of G. The distance d(u,v) between points u and v of a connected graph is the length of the shortest u-v path. A point v of G is a cutpoint of G if G-v has more components than does G. A plgngOf a graph G is a maximal nontrivial connected subgraph of G containing no cutpoint of itself. For a connected graph G with at least three points, the following are equivalent (see Chapter 3): (l) G is a block., (2) Every two points of G lie on a common cycle. (3) For every three distinct points of G, there is a path joining any two of them which does not contain the third. A block of order 3 or more is called a cyclic block while the block of order 2 is called the acyclic block. An endblock of a connected graph G having more than one block is a block containing exactly one cutpoint of G. There are several special classes of graphs to which we will frequently make reference. An acyclic graph or forest is a graph with no cycles. A planar graph is one which can be embedded in the plane. A (planar) graph which can be embedded in the plane such that each of its points lies in the boundary of the exterior region is called. outerplanar. The complete graph Kp has every pair of its p points adjacent. A triangle is a subgraph isomorphic to K3. A bipartite graph G is a graph whose vertex set V(G) can be partitioned into two disjoint subsets V1 and V2 such that every edge of G is of the form v1v2 where viEE Vi, i = l, 2. If V] and V2 have m and n points and G has mn edges, we Say that G is a complete bipartite graph and write G = K(m,n). G. Chartrand and F. Harary ([9], p. 107) have shown that a graph is outerplanar if and only if it has no subgraph homeomorphic with K4 or K(2,3), except K4-x (where x is any edge of K4). To any given nonempty graph G, there are associated with G several special graphs which we will encounter. The nth power Gn of G is that graph with V(Gn) = V(G) such that an edge uv €5E(Gn) if and only if,l §_d(u,v) §_n in G. The line graph L(G) of G has vertex set V(L(G)) which can be placed in one-to-one correspondence with E(G) in such a way that two points of L(G) are adjacent if and only if the corresponding edges of G are adjacent. The total graph T(G) has vertex set V(T(G)) which can be placed in one-to-one cor- respondence with the points and edges of G in such a way that two points of T(G) are adjacent if and only if the corresponding elements of G are adjacent or incident. CHAPTER 2 Section 2.1 Homeomorphic Graphs A collection of p points (O-simplexes) and q arcs (l-simplexes) joining certain pairs of points which is embedded in 3-space in such a way that every intersection or arcs occurs only at some of the p points is a finite geometric simplicial l-complex, or simply a l-complex. Two l-complexes e] and éz are homeomorphic if there exists a one-to-one bicontinuous mapping from 61 onto 62. Since every graph is realizable as a l-complex and every l-complex can be embedded in 3-space, there is at least one associated geometric l-complex for every graph. This observation serves as the motivation for defining two graphs to be homeomopphic with each other (or simply homeomorphic) if their associated l-complexes are topologically (a) ' (b) ' 1c) Figure 2.l. Geometric l-complexes: (b) 15 homeomorphic With (C) but not with (a). equivalent, that is, homeomorphic. Each graph is then said to be a homeomorph of the other. It is a consequence of this definition (see [17]) that two homeomorphic graphs can differ only by the number of points of degree 2 that they possess. Another approach to the study of homeomorphic graphs involves the concept of a subdivision (see [2], Chapter 8). An elementary subdivision of a graph G is a graph obtained from G by the removal of a edge uv of G and the addition of a new point w together with the edges uw and vw. Occasionally we may describe an elementary sub- division of G by stating that a new vertex w is inserted on the edge uv of G, or that uv has been subdivided into the edges uw and vw. A subdivision of G is a graph obtained from G by a finite sequence of elementary subdivisions. Using this terminology, two graphs G1 and G2 are homeomorphic if either G1 = 62 or there exists a graph G such that both G1 and G2 are subdivisions of G . 3 Sometimes it is useful to be able to determine when a graph 3 G2 which is homeomorphic with a graph G] is a subdivision of G]. We say that G2 is homeomorphic from G1 if either G2 = G] or G2 is a subdivision of G]. In Figure 2.2, G1 and 63 are homeomorphic graphs, neither of which is homeomorphic from the other, but both are homeomorphic from G2. 10 Figure 2.2. Homeomorphic graphs. Although the relation ”homeomorphic from" is not in general symmetric, the slightly more general relation “homeomorphic with" is clearly both symmetric and reflexive. We next verify that it is also transitive, showing that this relation is an equivalence relation. Before proceeding, however, we find it helpful to introduce some additional notation. Whenever a pair of points u and v of a given graph, neither deg u nor deg v being 2, are joined by r distinct u-v paths having interior points only of degree 2, let these paths be denoted by 03:), for 1 §_i 5_r. To check the transitivity, suppose that the graph G1 is homeomorphic with G2 and also that G2 is homeomorphic with G3. This means, by definition, that there exist graphs G4 and G5 such that G1 and G2 are homeomorphic from G4 and that G2 and G3 are homeomorphic from Gs. Now construct G6 from G2 by (1) replacing each 051) by the edge uv if u and v are not adjacent, otherwise by a u-v path of length 2, and (2) replacing each 03;) by a u-v path of length 2, for i 3_2. Since applying a sequence of elementary subdivisions to a ll graph only increases the length of paths 05:), both G4 and G5 are homeomorphic from G6. This implies that both G1 and G3 are also homeomorphic from G6, hence G1 is homeomorphic with G3, and we have established the fact that the relation "homeomorphic with" is an equivalence relation on the set of all graphs. Section 2.2 Homeomorphically Irreducible Graphs We have seen in the preceding section that under the equivalence relation "homeomorphic with“ all graphs are partitioned into classes, with two graphs belonging to the same class if and only if they are homeomorphic with each other. In [2], it is shown that in each such equivalence class CE.there exists a unique graph H with the minimum number of points of all graphs in this class. For convenience, we include here a reproduction of that result and its proof. Theorem 2.1. In every class C§.of homeomorphic graphs, there exists a unique graph H such that if G66, then G is homeomorphic from H. Epggf. Let nonouterplanar, then u and v must belong to some cycle in not containing uv. Then the insertion of w produces a subgraph of which is homeomorphic from K(2,3) (see[9],|3.107). Con- versely, if u and v are situated as prescribed for every partition of G, then n(G') = n(G) + 1. We summarize this as: Propgsition 2.3. Let G' be the elementary subdivision of G produced by subdividing the edge uv. Then n(G') = n(G) +1 if and only if each minimal partition contains a subset in which u and v belong to a cycle not containing uv. The group T(G) of a graph G consists of the set of all iso- morphisms of G under the operation of composition. Many of the basic properties possessed by these groups can be found in [9, Chapter 14]. Here we discuss the relationship of the automorphism group of an irreducible graph G with the automorphism group of a homeomorph of G. We note that each point of a graph must map into a point of like degree under isomorphism. 21 Many possibilities arise when one considers the orders of the groups of an irreducible graph and one of its homeomorphs. For example, if G is a path P2 with two points and if H is homeomorphic with G, then |F(G)| = |F(H)| = 2. If G is a cycle of length 3 and H is homeomorphic with G, then H is a cycle of length n for some n 3_3. Hence, |F(G)[ = 6, while |F(H)| = 2n, so that |F(H)| 3_|F(G)|. For the irreducible graph G of Figure 2.3 and the graphs H1 and H2 homeo- morphic with G, we have [T(G)] = 2, |P(H])| = l, and |F(H2)| = 8, so that lr(e)l > Ir1H])I but [T(G)l < IF(H2)I. Figure 2.4. Homeomorphs H1 and H2 of an irreducible graph G having |F(H])| < |F(G)| < |P(H2)|. This uncertain situation can be remedied, however, by requiring that G have no points of degree 2. Theorem 2.5. If a graph G has no points of degree 2 (and so is irreducible) and if H is homeomorphic with G, then T(H) is isomorphic to a subgroup of T(G). 22 2599:, Since H is homeomorphic from G, the points of H consist of the points of G together with the points of degree 2 resulting from how- ever many elementary subdivisions are required to transform G into H. If a denotes any isomorphism of H, then a must map the points of G onto themselves, since the points of G have degrees other than 2. Now let o' be the restriction of a to the set of points of G (so a' is one-to-one since a is). To show that a' is an isomorphism of G, consider any two adjacent points u and v of G. If u and v are also adjacent in H, then d(u) is adjacent to d(v) so that o‘(u) is adjacent to a'(v). If u and v are not adjacent in H, then the edge uv of G has been subdivided, say u, u], ”2’ ..., un, v, in the process of producing H. However, d(u), d(ul), d(uz), ..., d(u ), d(v) is a n path in H in which d(u.), i = l, 2, ..., n, is a point of degree 2; 1 thus this path is a subdivision of the edge a(u)a(v). Hence, in this case, o'(u) is also adjacent to a'(v) so that a' is an isomorphism of G. Under any isomorphism of H, a path whose interior points have degree 2 can only be mapped into a like path and the endpoints of such a path must be mapped into the endpoints of the image path. Therefore, each element aEF(H) uniquely determines an element o'e T(G), and two distinct elements of T(H) produce two distinct elements of T(G). From these observations and the manner in which the mappings a' were defined, it follows that the set F' of all such mappings is a subgroup of T(G) isomorphic to T(H). 23 Corollary 2.5. If a graph G has no point of degree 2, and if H is homeomorphic with G, then |F(H)| 5.|F(G)|. Pertaining to the area of traversibility, we say that a graph G is eulerian if it contains a closed trail, called an eulerian trail, which includes every edge of G exactly once and every point at least once. A graph is hamiltonian if it has a cycle, which we refer to as a hamiltonian cycle, which encounters every point of the graph exactly once. A well-known theorem of Euler (see [9], p. 64) states that a graph is eulerian if and only if it is connected and every point has even degree. Thus for an irreducible graph H and any homeomorph G of H, we have Prpposition 2.4. If H is irreducible and G is homeomorphic from H, then G is eulerian if and only if H is eulerian. The statement obtained through the substitution of “hamil- tonian" for “eulerian" in Proposition 2.4 is false, as can be seen by taking H to be the graph made up of two 3-cycles sharing a com- mon edge x and G obtained from H by inserting a new point on x. Two rather more stringent traversibility requirements are for a graph G to be randomly eulerian from appoint v65 G and to be randomly hamiltonian from appoint vEEG. The first of these two requirements means that G contains a point v such that the following procedure always results in an eulerian trail: Begin at v and traverse any incident edge. On arriving at a point, choose any incident edge which has not yet been traversed. When no unused edges 24 are available, the procedure terminates. Analogously, G is randomly hamiltonian from one of its points v if a hamiltonian cycle always results from this procedure: Begin at v, proceed to any adjacent point. 0n arriving at a point, select any adjacent point not previously encountered. When no unused points remain, and an edge exists between the final point chosen and v, the process terminates. In 1951 Ore [12] characterized graphs which are randomly eluerian from a point v as those graphs in which every cycle contains v. The following result is an immediate corollary. Proposition 2.5. If H is irreducible and G is homeomorphic from H, then G is randomly eulerian from one of its points if and only if H is randomly eulerian from one of its points. As was the case with Proposition 2.4, the substitution of "hamiltonian" for “eulerian" in Proposition 2.5 produces a false statement, as illustrated by the example in Figure 2.5. Figure 2.5. H is randomly hamiltonian from v, but possesses a sub- division G which is not hamiltonian. The example of Figure 2.5 prompts an additional remark about homeomorphism equivalence classes. Although the irreducible graph of 25 a class containing infinitely many nonhamiltonian graphs may be hamiltonian, the existence of one hamiltonian graph in a class guarantees that the irreducible graph for that class is also hamil- tonian (and similarly for the property of being randomly hamiltonian from a point). Section 2.4 Certain Homeomorphism Classes and the Hamiltonian Index A related aspect of the hamiltonian property is the hamil- tonian index h(G) of a graph G. For any connected graph G, which is not a simple path, h(G) is defined as the smallest nonnegative integer n such that the iterated line graph Ln(G) is hamiltonian. It is well known (see [9], p. 66) that every hamiltonian graph is 2-connected and that every 2-connected nonhamiltonian graph contains a subgraph homeomorphic from K(2,3). Thus the graphs homeo- morphic from K(2,3), and, more generally, graphs homeomorphic from K(2,n) (for n 3_3), provide an especially notorious class of non- hamiltonian graphs. Consequently, for such graphs G, the value of h(G) is positive. A natural question, then, to ask of a homeomorph of K(2,n) is this: ls l the best possible lower bound, in general, for h(G), and can the precise value of h(G) be determined? This question is answered in the affirmative by our next result. Theorem 2.6. Let G be any graph homeomorphic from K(2,n), n 3_3, and let u and v denote the points of G having degree n. Then 26 1 if n is even. h(G) = d(u,v) - 1 if n is odd. Proof. If G is homeomorphic from K(2,n) with n even, then u and v are joined by an even number of paths, say P1, P2, ..., Pn’ so that every edge of G appears in the trail constructed by travelling P1 from u to v, then P2 from v back to u, and so forth, exhausting all n paths. In [3], a graph with the property that every edge of a graph can be ordered in a sequence such that consecutive edges are adjacent is called sequential and it is shown there that the line graph of a graph is hamiltonian if and only if the original graph is sequential. Thus h(G) = 1. For G homeomorphic from K(2,n) with n odd, we proceed via induction on d(u,v), where u and v denote the points of G with degree n. If d(u,v) = 2, then at least one of the n paths Pi (i = l, 2, ..., n) joining u to v in G, say P], consists of only two edges x1 and x2. The line graph L(G) consists of two copies 8], B2 of Kn whose points are pairwise joined by n disjoint paths Pi' cor- responding to the n paths Pi in G such that each path P1' in L(G) has length one less than its counterpart Pi in G (see the following diagram). Thus the path P]' in L(G) consists of a single edge x. A hamiltonian cycle is then obtained for L(G) as follows: u], ”lu2’ P . u2, 2 , v2, vzv], v], v1v3, P3', u3, u3u4, U4, P4', v4, ..., vn, Pn un, unu], u]. Hence h(G) = l = d(u,v) - l. 9 27 ,.. P' _.. I61- \\ X 1 // cv]\\ ‘ 'V ‘. P' ' I 1u2°\'_/\/g—\_!/°v2 1 l ' ' I U ' 1: l l l 2 1 l 1 l I PI | I 1 1 1 {UDM /—--O' Vn,‘ \ l \’/ \_’l G L(G) Assuming that h(G) = d(u,v) - l for all homeomorphs G with d(u,v) < k, we seek to verify that h(G) = d(u,v) - 1 when d(u,v) = k. From G, construct G* by removing one edge from exactly one shortest path of G; this yields d(u,v) = k - l in G*, so that G* falls under the induction hypothesis, so that h(G*) = k - 2, implying that Lk-2(G*) is hamiltonian. Now consider the k - 2nd line graphs of G and G*. Since the shortest u-v paths present in G and 6* have lengths k-2( k and k-l, respectively, we are assured that Lk-2(G*) and L G) each consist of identical subgraphs B1 and 82 joined by n disjoint paths, the only distinction between them being that precisely one k-2 path of L (G*) is a single edge x while the corresponding shortest k-2 path in L (G) consists of two edges x1, x2 (see the figure below). ”~\l\.,___._.—-—-————O—/-o \\ I ""+ —————— fl; \ / ‘ ’_»_’_,_.—1._. \ |l ._\_._ ________ \‘ I! B 1 " ”””” ’+'. B A I B f—‘L‘ ------ E B I 1 I ' ' i 1 : 2./ 1, 1: I I 1 E 2; \ T J V 28 Although Lk‘2( G*) is hamiltonian, no hamiltonian cycle of it contains the edge x since such a cycle must also contain the other n-l paths (possessing interior points) connecting B1 and 82, making an odd number of disjoint paths that such a cycle must contain, which is k-2( impossible. Similar reasoning precludes L G) being hamiltonian, as it contains an odd number of paths with interior points connecting B1 to 82. However, Lk'2(G*) and Lk'2(G) can now be redrawn as shown in the following figure. Lk-2(G*) Lk-2(G) Another result of G. Chartrand ([3] p. 562) states that if a graph H contains a cycle C with the property that every edge of H is incident with at least one point of C, then L(H) is hamiltonian. The graph Li‘-2 (G) satisfies this condition, thus insuring that L(Lk'2(e)) = Lk"(o) is hamiltonian. Hence h(G) = k - 1 = d(u,v) - 1. This completes the proof. To illustrate Theorem 2.6, consider the graph G1 composed of 2 points joined by 3 disjoint paths of length 3 and the graph 62 29 consisting of 2 points joined by 4 disjoint paths of length 3. Then h(G1) = 2 while h(Gz) = 1. If G]' and 62' represent the graphs obtained from G1 and G2 by the insertion of an additional point into each edge, then h(Gl') = 5 while h(GZ') = h(Gz) = 1. CHAPTER 3 Section 3.1 The Nth Subdivision Graph and Its Structure In 1965 Harary and Nash-Williams [10] introduced the concept of the nth subdivision graph Sn(M) of a multigraph M. By definition, this is a graph obtained from M by replacing every edge uv of M by a u-v path of length n + l, where n 3_l. The graph S](M) is denoted simply S(M) and is called the subdivision graph of M. Behzad [1], also in l965, utilized the subdivision graph S(G) of a graph G in order to characterize the total graph T(G) as isomorphic to the square of S(G). While M ordinarily is not a graph, Sn(M) is always a graph; and if M has p points and q edges, then Sn(M) is a (p + nq, (n + l)q)-graph. The above definition for the nth subdivision graph could be extended to the case where G is a pseudograph, but we shall not consider this possibility since when n = l, S(G) may be a multigraph and not a graph. 0‘ M S(M) 52(M) Figure 3.1. A multigraph M together with S(M) and 52(M). 30 31 We next consider some elementary properties of the subdivision graph S(M) of a multigraph M. If we regard S(M) as obtained from M by the insertion of one new point on every edge of M, then the original points of M are nonadjacent in S(M) and form an independent set in S(M). Since each new point is adjacent to exactly two old points, the set of all the new points is also independent in S(M) and contains only points of degree 2. It is clearly necessary, then, that for a graph G to be a subdivision graph of a multigraph M, G must be bipartite, and one of the two independent sets which partition the point set of G contains only points of degree 2. We prove later that this restriction is also a sufficient condition for a graph to be a subdivision graph. Thus the only regular graphs which are subdivision graphs are those that are 2-regular, namely those graphs which are unions of disjoint cycles of even length. Of course, for n 3_l, Sn(M) may contain odd cycles. We next consider the question of deciding when a given graph G is the nth subdivision graph of some multigraph M, and examine a few simple choices for G. Certainly, (a) no nonempty complete graph can be an nth subdivision graph, (b) every cycle Ck(n + 1), for k 3_3, is isomorphic to the nth subdivision graph of Ck (for k = 2, 02(n + 1) = Sn(90) where 90 is the multigraph consisting of two points joined by a pair of edges), (c) every path P(k _ l)(n + 1) + 1 is isomorphic to Sn(Pk)’ and (d) the complete bipartite graph K(m,n) is a subdivision graph if and only if at least one of m and n is 2. 32 Since we have already characterized 2-regular subdivison graphs as unions of disjoint cycles, we now examine graphs containing at least one point whose degree is different from 2. An uncomplicated criterion for identifying those graphs which are nth subdivision graphs is pre- sented as the following theorem. Theorem 3.1. Let G be a graph which is not 2-regular. Then G is h the nt subdivision graph of a multigraph if and only if no block of G is a cycle of length n + l and every ul-uz trail in G has a length which is a multiple of n + l for all points u1 and u2 with deg ”i f 2, for i = l, 2. Epggf. Necessity. Let M be any multigraph and consider G = Sn(M). In obtaining G from M, every edge uv of M is replaced by a path of length n + 1. Thus every ul-u2 trail in G joining points u1 and u2 in M has length (k - l)(n + l), where k is the number of points of M that appear in this trail. Since the only points of G which are not of degree 2 are also points of M, then every u1-u2 trail in G joining points u1 and u2 which are not of degree 2 has length that is a multiple of n + 1. However, no block composed entirely by a ul-u1 cycle of length n + 1, deg u.l f 2, can occur in G, since such a cycle corre- sponds to a loop in M. Sufficiency. Now let G be any graph, not 2-regular, with the property that every u1-u2 trail joining points u1 and ”2’ which are not of degree 2, has length a multiple of n + 1. We must show the existence of a multigraph M such that Sn(M) = G. We may assume G 33 is connected, for otherwise the ensuing argument can be applied to each component of G in order to construct the corresponding component for M. Select any point v1 in G which does not have degree 2. At least one such point exists, since G is not 2-regular. Let V be the set of all points in G whose distance from v1 is a multiple of n + l and let U be the set of all other points. Thus, by hypothesis, all points of G which are not of degree 2 belong to V so that U contains only points of degree 2. Every pair of points Vi’ vj of V are joined only by trails whose length is a multiple of n + 1, for if Vi were joined to Vj by a trail T0 of length d which is not a multiple of n + 1, then To, together with trails T1, T2 from v1 to vi and from v1 to vj, forms a v1-v1 trail whose length is not a multiple of n + 1. Now let P1 and P2 be shortest paths from any point uEEU to the set V, say to points Vi and vj in V. Inasmuch as the vi-vj trail T formed by P1 and P2 has no interior point that is a point of V, the length of T must be exactly n + 1. Furthermore, Vi # v\j because if vi = Vj’ then all points of T except Vi have degree 2, making T a block of G, a contradiction. 34 Thus every point u of U lies on some vi-vj path P1 j of length n + 1, all of whose interior points are in U, joining two points of V. Each point of U has degree 2, making it impossible for any point of U to simultaneously lie on two such paths; thus each point uEEU uniquely determines one of these v1-vj paths P. 1,1. If we label all paths of length n + l joining each pair of points v1, vj of V by P113, Pizi, ..., P(t), the desired multigraph 1sJ19J M can now easily be described. The point set of M is the set V and two points v1 and vj of M are joined by one edge x(kg for each path 19”}, k= 1, 2, ...,t. All that remains is to verify that Sn(M) = G, that is, to exhibit an isomorphism ¢ between Sn(M) and G. The point set of Sn (M) is VLJV], where V1 consists of n points of degree two "E5; 1, "(5i 2, ..., wi53,n for each edge xffg of M, with "(5; r( ) k adjacent to w. for r = l, 2, ..., n - 1. In G each path P1 1 i,i M of length n + 1 from v. to vj contains precisely n points of U, say 1 k k k k k "(,3 1, u(,; 2, ..., u§,g n’ where u(,; r (’3 r+1 r = l, 2, ..., n - 1. Thus the mapping ¢ of S n(M) onto G defined is adjacent to u by the equations ¢(v.) = v. for v1€EV, ¢(w(k3 r) = u(kg r for Wékg rEEV1s r = 1,2, ..., n. is one-to-one and adjacency-preserving. This completes the proof of Theorem 3.1. Corollary 3.1. Let G be a graph which is not 2-regular. Then G is the subdivision graph of a multigraph if and only if every ul-u2 trail in G has even length for all points u1 and ”2 with deg U1? 2, i = l, 2. 35 3399:, This corollary follows immediately from Theorem 3.1 with the observation that when n = l, the hypothesis forbidding the presence of a cycle of length n + l as a block of G is superfluous. When a graph G is known to be an rnfllsubdivision graph of some multigraph M, one might wonder if M is unique up to isomorphism; that is, can G be the nth subdivision graph of two or more noniso- morphic multigraphs? This question is settled in the negative by the following: Theorem 3.2. If M1 and M2 are connected multigraphs whose nth sub- division graphs are isomorphic, then M1 and M2 are isomorphic. Proof. It is convenient to deal first with connected multigraphs M1 and M2 having 3 or fewer points and which have isOmorphic nth subdivision graphs. If we temporarily agree to call the multigraph Go a cycle, then whenever Sn(Ml) = 5 (M2) is 2-regular, namely a n cycle, then M1 and M2 must also both be cycles, necessarily with the same number of points, and hence isomorphic. If Sn(M]) = Sn(M2) is not 2-regular, then it contains a point of degree k f 2. Since the only points of Sn(M]) and Sn(M2) not of degree two are those points which are in the original multigraph, consequently M1 and M2 must pggp consist of one point v joined by k edges to precisely those points connected to v in the nth subdivision graph by a path of length n + 1. Next we consider connected multigraphs M1 and M2 having more than 3 points with Sn(Ml) = Sn(M2)‘ The remainder of this proof was 36 suggested by an elegant analogous proof for line graphs given by Jung [12]. For any isomorphism ¢1 : Sn(M]) + Sn(M2), we shall exhibit an induced isomorphism ¢ : M1 +-M2. Suppose v is any point of M1, and let E(v) be the set of all edges incident to v. We claim that to every véEM1 there is exactly one point v' in M2 such that E(v) + E(v') under ¢]- If deg v 3_2, let x1 and x2 be edges at v and let v' be the common point of edges ¢](x1) and ¢1(x2). Then for each edge x at v, v' is incident with ¢](x); and for each edge x' at v', v is incident with ¢'1(x'). If deg v = 1, let x = uv be the edge at v. Then deg u 3_2, hence E(u) corresponds to E(u') and ¢](x) = u'v', where v' = ¢](v). Since for every edge x' at v', the edges o'}(x') and x have u as a common point and u' is on x', therefore x' = ¢1(x) and deg v' = 1. Thus in all cases there is precisely one v' in M2 for each v in M1 such that E(v) + E(v') under 6]; call this induced mapping ¢- Then ¢ is evi- dently one-to-one on the point sets of M1 and M2 since E(u) = E(v) only if u = v. Also o is onto, since to any point v' in M2 there exists an incident edge x' which determines an edge o'}(x') in M1, one of whose endpoints maps into v'. The adjacency-preserving property of t is inherited from ¢], and the proof is complete. One useful benefit which we can derive from Theorem 3.2 is that since each nth subdivision graph G is obtained from only one multigraph M (up to isomorphism), then we can always regard M as obtainable from G by the procedure used in the proof of Theorem 3.1. Furthermore, if G is an nth subdivision graph of a graph M, we can 37 regard G as composed of an independent subset V of points v1, v2, ..., v together with certain v1-vj paths P11 of length n + l, t where each edge in G belongs to some P11. and no more than one path . “oins v. and v.. J J 1 J A related question is whether there exist nonisomorphic multi- P. 1 graphs M1 and M2 with Sm(M]) = Sn(M2) for m f n. This can occur as is seen by taking m = 2, n = 3, M1 = P5, and M2 = P4 (or M1 = C4, and M2 = C3). A more general example of this is obtained by taking M1 = Sn(M) and M2 = Sm(M) for any multigraph M and distinct integers m and n; then Sm(M]) = Sm(sn(M)) = Smn+m+n(M) = Sn(Sm(M)) = S (M n 2) and M1 and M2 are nonisomorphic graphs. Although each acyclic block of a graph G gives rise to n + l acyclic blocks in Sn(G), the next proposition shows that G and Sn(G) have the same number of cyclic blocks. Proposition 3.1. The set of cyclic blocks of Sn(G) is {Sn(B1)} where {Bi}’ 1 = i, 2, ..., k, is the set of cyclic blocks of G. 3529:, Suppose that B' is any cyclic block of Sn(G) and denote those points of B' which are also points of G by v1, v2, ..., Vr' We claim that the induced subgraph = B of G is a cyclic block whose nth subdivision Sn(B) equals 8'. One way to show that B is a cyclic block is, for every three distinct points of B, to exhibit a path joining any two of them which does not contain the third (B surely contains three or more points since 3' cyclic implies that 8' contains a cycle which is an nth subdivision of a corresponding 38 cycle in G). For each trio of points u, v, w of B, there is a path P' in B' joining u to v which does not contain w, by virtue of B' being a cyclic block. Since B' is a subgraph of an nth subdivision graph, P' has length a multiple of n + 1 and is of the form u, u0,1, 110,2, ..., ”0,", U], U],-l, 111,2, ..., “1,11, U2, ..., utfll’ Ut+1= V. Thus in B, the path P : u, u], ”2’ ..., ut, u = v joins u and v, t+l omitting w. Now let B be any cyclic block of G and consider the subgraph B' = Sn(B) of Sn(G). To demonstrate that B' is a cyclic block of Sn(G), it suffices to show that each pair of points of B' lie on a common cycle. Letting u and v be any two points of B', we distinguish three cases concerning them. CASE 1. If both u and v are points of B, then since B is a cyclic block, there exists a cycle C containing both u and v; thus Sn(C) is a cycle in 8' containing both u and v. CASE 2. If exactly one of u and v is not a point of B, say u, then u is an interior point of a ul—u2 path P' of length n + l in B' where u1 and u2 are points of 8. Since u], v, and u2 all belong to the block B, there exists a ul-v path P] not containing u2 and a v-u2 path P2 not containing u]. Thus in B', the paths P', Sn(P1), and Sn(P2) form a cycle containing u and v. CASE 3. If neither u nor v is a point of B, then in a manner almost identical to the approach employed in CASE 2, a cycle can be con- structed in B' which is common to u and v. 39 Thus every cyclic block B1 of G gives rise to a cyclic block Sn(Bi) of Sn(G) and, conversely, each cyclic block Sn(B1) of Sn(G) corresponds to precisely one cyclic block B1 of G. If we restrict our attention to those graphs G for which G = S(H) for some graph H, we obtain several additional equivalent conditions. Theorem 3.3. Let G be a graph which is not a union of disjoint cycles. Then the following statements are equivalent: (1) G is the subdivision graph of a graph. (2) G does not contain K(2,2) as a subgraph and every ul-u2 trail in G has even length for all points u1 and u2 with deg u1 # 2, i = 1, 2. (3) G is bipartite having a point set which is a dis- joint union of independent sets U and V such that U contains only points of degree 2 and no two points in U are mutually adjacent to the same two points of V. 3599:, Throughout the proof we shall assume G is connected since if the theorem holds for every component of G, it holds for G itself. (2)==>(l). Since a connected graph which is not a cycle cannot be 2-regular, G satisfies the hypothesis of Corollary 3.1. It remains to be checked that by forbidding the presence of K(2,2) as a subgraph of G we guarantee that the multigraph M constructed in the proof of Theorem 3.2 is actually a graph. Referring to the 40 construction of M in the sufficiency portion of the proof of Theorem 3.1, we find that the only way that one pair of points v1, v. can be joined by two lines x! ., x? . in M is by the occurrence J 1 J 1:3 of at least two paths P11}, P12; of length 2 from v1 to vj in G. But these paths Pili’ P§?g together form the subgraph K(2,2) in G. Thus every pair of edges in M with common endpoints corresponds to a K(2,2) subgraph of G. (1)==>(3). For any point v1 of G which is not of degree 2, we define V as the set of all points of G having even distance from v1 and U as the set of all other points of G. Then, as in the proof of Theorem 3.1, all points of U have degree 2 and both U and V are independent sets, implying that G is bipartite. Again referring to the construction of M such that S(M) = G detailed in the proof of Theorem 3.1, no two points of U can be interior points of two paths P11}, P§?; joining the same pair of points v1, vj of V since this would mean that in the graph M v1 and vj are joined by 2 edges. Hence no two points of U are mutually adjacent to the same pair of points of V. (3)==>(2). If G is a bigraph whose point set is partitioned into subsets U and V such that U contains only points of degree 2, then every vl-v2 trail, where v1 and v2 are not 2-valent, joins points in V and hence has even length. Thus by Corollary 3.1, G is the sub- division graph of a multigraph M. A repetition of the same argument given in the preceding paragraph suffices to show that no multiple edges can occur in M; that is, M is a graph. 41 Corollary 3.3a. A graph G is the subdivision graph of Kn’ n 3_2, if and only if G is a bipartite graph, containing no 4-cycle, and whose point set can be partitioned into two independent subsets composed of n points of degree~n - l and (3) points of degree 2. 3599:, First we show that the above condition is sufficient. By Theorem 3.3(3), G = S(H) for some graph H. Assuming that the above condition holds, it remains to be shown that H = Kt for some integer t 3_2. If H is a (p, q)-graph, then G has p + q points and 2q edges. Since each edge of G is incident with exactly one of the n points of the independent set V, we have that 2q = n(n - l), or that q = 91575—11 = (3). All of the p + q points of G lie in U V, thus p + q = (g)+ n implies that p = n. Hence H is the complete graph on n points. The necessity follows immediately. Corollary 3.3b. A graph G is the subdivision graph of the complete bipartite graph K(m,n) if and only if G contains no 4-cycle and the point set of G is the disjoint union of independent sets U, V, and W, composed respectively of mn points of degree 2, m points of degree n, and n points of degree m. Egggj, The necessity is immediate. To show that the above condition is sufficient, we note that Theorem 3.3(3) guarantees that G = S(H) for some graph H; it remains to be seen that H = K(s,t) for some pair of positive integers s and t. If H is a (p, q)-graph, then G has p + q points and 2 q edges. Summing the degrees of all the points of G, we 42 find that 2q = l/2(2mn + mn + nm), or that q = mn. Then p + q = mn + m + n implies that p = m + n. Hence H = K(m,n). Section 3.2 Properties of Nth Subdivision Graphs Since a graph G and its nth subdivision graph Sn(G) belong to the same homeomorphism equivalence class, they possess many common characteristics. Although such concepts as order or the number of edges are obviously not preserved, planarity and the invariance of the number of connected components is maintained. Outerplanarity, on the other hand, is not in general inherited by Sn(G) from G, as shown by K4-x and it subdivision graph. S(K4-x) f Figure 3.2. An outerplanar graph K4-x and its nonouterplanar sub- d1v151on graph. It is possible to completely delineate those nth subdivision graphs which are outerplanar, as shown by our next theorem; before Proceeding, however, we present a definition. A gpgtg§_is a connected graph every block of which is either a (zycle or K Within the context of nth subdivision graphs, the 2. 43 concepts of cacti and outerplanar graphs coincide, as we now see. Theorem 3.4. Let G be an nth subdivision graph. Then G is outer- planar if and only if every component of G is a cactus. Erggf, Since the sufficiency is evident, we proceed directly to the necessity portion of the proof. We may assume that G is a connected outerplanar nth subdivision graph, for otherwise we may consider the individual components of G. Moreover, we may also assume that G contains at least one point not of degree two since the theorem holds for 2-regular connected outerplanar nth subdivision graphs, i.e., disjoint unions of cycles. To verify that G is a cactus, we proceed via induction on the order of G. The theorem is immediate for connected nth subdivision graphs on three or fewer points. Assuming that every outerplanar connected nth subdivision graph with fewer than p points is a cactus, we consider such a graph G on p points. The blocks of G having 3 points or less are either single edges or 3-cycles, and both types are cacti. Accordingly, we let 8 be any block of G having 4 or more points. We may regard the nth subdivision graph G, by means of the construction given in the proof of Theorem 3.1, as having a point set made up of two subsets U and V, where all noncarriers are contained in V and all points of U have (k degree 2, and where all the edges of G occur in paths P1 1 of length n + l joining distinct points v1, v1 in V. Since 8 is a cyclic block, 44 it contains a point of U which is an interior point of some such path P of length n + l joining two points u and v of V; u and v are also points of B. Denote the subgraph obtained from G by deleting the n interior points of degree 2 of the path P as 6*. Note that G* is an nth subdivision graph since all ul-u2 trails in G* joining points u1 and u2 not of degree 2 (which are elements of the set V) have length a multiple of n + 1. To be sure, G outerplanar implies the subgraph G* is outerplanar; thus by our induction hypothesis, 6* is a cactus. In the block B, the points u and v belong to a common cycle C. If we can demonstrate that C comprises all of B, then we will have shown that every block of the graph G is either a cycle or a single edge, so that G is indeed a cactus. For every three distinct points of B, there exists a path joining any two of them which excludes the third. Thus B contains a u-v path P' which is disjoint from the path P. No two points of P' lie on a cycle in 6* because this would generate a subgraph homeomorphic from K(2,3) in B, contradicting the outerplanarity of G. Thus the cycle C is composed of the paths P and P', so that we can label C by u, u], ”2’ ..., u , v = wo, w], "2’ ..., wk_], wk = u, n where the u., l 5_i 5_n, are carriers in P and the w1, O §_i §_k, 1 are the points of P'. If 8 contains a point w not belonging to C, then again by utilizing the criterion of the first sentence in the preceding para- graph, we can find paths P", P"' joining w to distinct points W1. wj of C (if every path from w to C joins w to the same point w1, then w1 is a cutpoint of B). 45 -- - Q But this implies the existence of a cycle in G* containing the two points w1 and "j of P', which we've already seen is impossible. Therefore C passes through every point of B. This completes the proof. Section 2.3 of Chapter 2 contained a discussion of various characteristics of a graph G and any subdivision H of G, in particular their chromatic numbers and their automorphism groups. The remainder of this section is devoted to obtaining more precise results in these two areas when H = Sn(G). Proposition 2.2 of Chapter 2 states that if G is irreducible, with chromatic number x(G) 3_3, then for any graph H in the same homeo- morphism class with G, we have 2 §_X(H) 5_x(G). In particular, for all positive integers n, we have 2 §_ X(5n(G)) 5_X(G). This inequality can in general be greatly improved, as our next result shows. 46 2 if n is odd. Theorem 3.5. Let x(G) 3.3. Then X(5n(G)) = 3 if n is even. Proof. Suppose the point set of G is partitioned into x(G) = k color classes C], C , Ck, where k 3_3. 2, ... For odd values of n, say n = 2r - l with r = l, 2, 3, ..., Sn(G) is obtained from G by inserting new points u], u2, ..., ”Zr-l into each edge uv of G. A minimal 2-coloring of Sn(G) is then achieved by assigning color a to the points of the type ”2’ U4, ..., ”Zr-2 and also to the points of Sn(G) which are in G, and color 8 to points of type u], u3, ..., ”Zr-1' When n is even, say n = 2r with r = l, 2, 3, ..., X(G) > 2 implies the existence of at least one cycle of odd length 5 in G (since every pair of color classes are joined by at least one edge). Thus in Sn(G) there exists at least one cycle of odd length s(2r + 1), so that X(5n(G)) > 2. Now Sn(G) is obtained from G by inserting 2r points u], "2’ ..., ”2r into each edge uv of G; if the points of Sn(G) which are in G are assigned color a, points of the type u], u3, ..., ”Zr-l are colored 3, points of the type ”2’ U4, ..., ”Zr-2 are colored a, and u2 is colored y, then we have produced a 3- r coloring of Sn(G). Hence X(Sn(G)) = 3. For any irreducible graph G, Theorem 2.5 guarantees that T(Sn(G)) is isomorphic to a subgroup of T(G). In the specific case concerning nth subdivision graphs, this result can be strengthened considerably. 47 Theorem 3.6. For any graph G, no component of which is a cycle, F(Sn(G)) is isomorphic to T(G) for every positive integer n. Proof. As in the proof of Theorem 3.1 and the remarks following Theorem 3.2, both G and Sn(G) can be regarded as consisting of an independent set V of points v1, v2, ..., v with one path P1. of t J length n + l joining v1 to Vj in Sn(G) whenever the edge Vivj is j has interior points of degree 2 only, so that under any automorphism a of Sn(G), each path Pij must map present in G. Each path P1 into a like path and the endpoints of P11 must be mapped into the endpoints of the image path. (If any interior point u1 of some path P11 maps into a point vjEEV, then deg Vj = 2 and since the component of G containing vj is not a cycle, there exists a point vkEEV with deg vk f 2. Then dSn(G)(Vj’ vk) E O mod(n + 1) and dSn(G)(ui’ Vj) z 0 mod(n + 1). Hence dSn(G)(ui’ vk) i 0 mod (n + l), but dsn(G)(o(u1), d(vk)) = dSn(G)(vj’ a(vk)) E 0 mod(n + l), inasmuch as d(vk)€EV, contradicting the assumption that o is an automorphism.) Thus each element a€F(Sn(G)) uniquely determines an element a'E I‘(G), and conversely. From this observation and the way in which these automorphisms are defined, it follows that T(G) and P(Sn(G)) are isomorphic. Section 3.3 Factorization in Sn(G) An r-factor of a graph G is a nonempty spanning subgraph of G which is regular of degree r 3.1. If a graph G can be expressed 48 as an edge-disjoint union of r-factors, we say that G is r-factorable. Clearly a graph G may have an r-factor without being r-factorable. Inasmuch as Sn(G) always contains points of degree two, we shall deal only with the question of the existence of a l-factor, or a 2-factor, for Sn(G) and certain other related graphs. It is immediate that a necessary condition for a connected graph G to contain a l-factor is that G have even order. It is easily seen that this simple condition in general is not sufficient; for example, the complete bipartite graph K(l, n-l), for n an even integer (n > 2), does not contain two nonadjacent edges, much less n/2 such edges. We continue our discussion of factorization with an elementary observation which will be used in the sequel. A path or cycle on n points is denoted by Pn or C", respectively. Proposition 3.2. For k = l, 2, 3, ..., the path P2k has a unique 1-factor while the cycle C2k has exactly two 1-factors. Proof. For any path P2k : v1, v2, ..., V2k’ the unique l-factor must consist of the edges V1V2’ v3v4, ..., V2k-lv2k' In the case of the cycle C2k : v1, v2, ..., v2k, v], in addition to the aforementioned set of edges, the edges v2kv], v2v3, ..., V2k-2v2k-l also form a l-factor. Obviously P2k is not l-factorable, while C2k is l-factorable. Whenever a graph G possesses a l-factor F, it follows that Sn(G) also has a l-factor provided that n is an even positive integer. 49 To verify this assertion, it is helpful to observe that if the point set of G is V = {v1, v2, ..., vp}, then Sn(G) consists of the set V together with q disjoint vivj paths P11 of length n + l, where each path Pij 1n Sn(G) corresponds to the edge Vivj in G. In $2k(G)’ each path which corresponds to an edge x1 of G in F has odd length, hence possesses a unique l-factor F1, for i = l, 2, 3, ..., p/2. Letting F6 denote the set of all edges of the form u21_]u21, for 1 = l, 2, 3, ..., 2k, 1n all paths P11 : v1, u], ”2’ ..., u2k, vj wh1ch cor- P/Z respond to an edge of G not in F, K.) F1 is the desired l—factor of i=0 $2k(G). Conversely, if $2k(G) possesses a l-factor F, a l—factor for G is formed by all edges Vivj corresponding to those Vi‘vj paths P11 in S2k(G) which contain an edge x1€EF incident with v1 (notice that two paths Pij’ Pik cannot both contain an edge from F which is incident with v1). Thus we have shown Theorem 3.7. A graph G has a 1-factor if and only if $2k(G) has a l-factor, where k = l, 2, 3, .... We next consider the question of deciding when an odd sub- division graph 52k-1(G) has a 1-factor, for k = 1, 2, 3, .... The following lemma reduces this question to a simpler one. Lemma 3.1. For any graph G and positive integer k, 52k-1(G) has a l-factor if and only if S(G) has a l-factor. 50 23921. Let V = {v1, v2, ..., vp} be the point set of G. Then S(G) and 52k-1(G) each consist of the set V together with exactly one v1-vj path of length 2 or 2k, respectively, whenever v1vj is an edge of G. To any l-factor of S(G), there is an associated l-factor F' of 52k-1(G) obtained in the following way. In S(G), each point v1 is incident with exactly one edge x1€EF and this edge lies in exactly one v1-vj path P13. Labelling the corresponding v1-vj path P..' in $2k-1(G) as v1, u], ”2’ ..., u2k_], vj, let F' be the set lJ {Viu'ls U2U3, ..., "ZR-ZUZk-1}‘ Then F' = .HIFi' 15 a 1-factor 0f 52k-1(G)' Of course, if $2k-1(G) has a l-factor for each positive integer k, then by taking k = l, we have that S(G) has a l-factor. One additional definition is needed in dealing with the ques- tion of the existence of al-factorfor 52k-1(G)° A connected graph is called unicyclic if it contains precisely one cycle. Theorem 3.8. For any graph G and positive integer k, 52k-1(G) has a l-factor if and only if each component of G is unicyclic. 3592:, By Lemma 3.1, it suffices to show that S(G) has a l-factor if and only if each component of G is unicyclic. In showing the necessity of this condition, we follow a suggestion of J. Zaks. If G has a component C which is not unicyclic, then the number of points p of C is not equal to the number of edges q of C (a connected (p, q)- graph is unicyclic if and only if p = q). All edges of the bipartite component S(C) of S(G) are of the form u1vj, where u], ”2’ ..., u P and v1, v2, ..., vq correspond to the points and edges, respectively, 51 of C. Thus any l-factor of S(C) is composed of nonadjacent edges "ivj’ implying that there is a one-to-one correspondence between the u. and the Vj' But this means that p = q, a contradiction. Hence 1 each component of G is unicyclic. Conversely, assume that each component of G is unicyclic and consider any component C' of S(G), where C' = S(C) for some component C of G. Then C unicyclic implies that C' is unicyclic; denote the unique cycle in C' by Z : v], u], v2, ”2’ ..., Vt, ut, v], where the points v1 correspond to points in C and deg u1 = 2, for i = l, 2, ..., t. As in Proposition 3.2, Z = C2t has a l-factor F0. C' - Z is a union of trees Tk, for k = l, 2, ..., r, where each tree Tk contains a point wk adjacent to some point vj of 2 with deg Vj # 2. The distance d(u,v) in C' = S(C) between points u and v, neither of whose degrees is 2, is always even; thus the distance d(wk,v) in Tk’ where deg v f 2, is always odd. This implies that Tk has a unique l-factor Fk composed of all edges uv in Tk’ where d(wk,u) is odd and r d(w .V) is even. Then F = \,)F k k=0 k is a l-factor for S(G). Corollary 3.8. Let G be a connected (p, q)-graph and k a positive integer. Then 52k-1(G) has a l-factor if and only if p = q. In general, the existence of a l-factor for a (p - q)-graph G is independent of the existence of a l-factor for S(G), even if we restrict our attention to connected graphs G having both p and q even. This lack of relationship is illustrated by the following four infinite classes of connected graphs. For each positive integer i, if G1 is 52 the cycle C 2, then G1 has a l-factor, as does S(G1), since G1 is 2i+ unicyclic. Taking G1 as the cycle C21+2 together with edges wv1 and wv2 joining any point w of C21+2 to two additional new points v1 and v2, we find that G1 has no l-factor, although S(G1) does. In Figure 3.3, the remaining two situations are illustrated. / / I K \ C2i+1 (a) Graphs 61 = W21+2 having (b) Graphs G1 having no a l-factor while S(Gi) l-factor and such that has no l-factor. S(G1) also has no 1-factor. Figure 3.3. Graphs G such that S(G) has no l-factor. Choosing 61 as the "wheel" W21+2 of Figure 3.5(a), it is easy to find a l-factor of G1, but S(Gi) has no 1-factor. If G1 consists of three copies of C 1 joined by single edges with a new point v, 21+ as in Figure 3.5(b), then neither 61 nor S(Gi) possess a l-factor. 53 A criterion for Sn(G) to be l-factorable is much easier to establish than a test for the existence of a l-factor. For Sn(G) to be l-factorable, it must be an edge-disjoint union of l-factors F1, F2, ..., Ft’ so that Sn(G) is regular of degree t. Thus t = 2, imply- ing that Sn(G), and G as well, is a disjoint union of cycles. We summarize this as Proposition 3.3. For any graph G, Sn(G) is l-factorable, for n 3_l, if and only if each component of G is a cycle. For any graph G, the total graph T(G) is related to the sub- division graph S(G) by the equation T(G) = 52(6). We shall make use of this to develop necessary and sufficient conditions for T(G) to possess a l-factor. In order for the nth power Gn (n 3_2) of any graph G to have a l-factor, Gn must possess an even number of points. The partner- ship of G. Chartrand, A. Polimeni, and the author was able to prove that this simple necessary condition is also sufficient for Gn to have a l-factor. Theorem 3.9. Let G be an arbitrary graph. Then Gn (n 3_2) has a l-factor if and only if each component of G has even order. Proof. It suffices to show that G2 has a l-factor if and only if G has even order, inasmuch as Gn (n 3_2) and G share the same point set. We may further assume that G is a connected graph, for otherwise we may apply the following argument to each component of G. 54 The necessity of the given condition being obvious, we proceed to the sufficiency portion of the proof. Letting the order of G be 2n, the proof is by induction on n. For connected graphs on 2 points, the assertion is obvious. Assuming the assertion is true for all appropriate graphs on fewer than 2n points, we consider a connected graph G with order 2n. We may further assume that G consists of two or more blocks, for if G is a single cyclic block, then G2 has a hamiltonian cycle (see [8]), alternate edges of which furnish a 1- factor. It is convenient to distinguish two cases. CASE 1. G contains a cyclic endblock 8. Suppose that v is the cut- point of G in B. In [6], it is shown that for any cyclic block B with 2 at least 4 points, 8 - u is hamiltonian for any u B. Thus if B contains an odd number p of points, P 2.5, then B2 - v is hamiltonian and has an even number of points; hence there is a l-factor F0 for 2 B2 - v. If B has exactly three points, then B - v is simply a copy of K2, and therefore is a l-factor, say F0. Also the connected graph G' = G - (B - v) has an even number of points, so that by the induction assumption there is a l-factor F1 of (G')2. of 62. Then FOUF1 is a l-factor CASE 2. Each endblock of G is acyclic. Let v denote a cutpoint of G such that all blocks containing v, except at most one, are endblocks. If there are at least two endblocks vv1 and W2 containing v, then the connected graph G' = G - {v],v2} has an even number of points and by the induction assumption, (G')2 has a l-factor F. Since the edge vlv2 55 is present in Gz, FL}{v]v2} is a l-factor of G2. In the event that there is just one acyclic endblock vv1 containing v, let 8 be the other block containing v. Then G' = G - {v,v]} is connected and has 2n - 2 points, so that the induction hypothesis implies that there is a l-factor F for (G')2. Then FlJ{vv]} is a l-factor of G2. This com- pletes the proof of the theorem. Corollary 3.9a. Let G be a connected (p,q)-graph. Then T(G) has a l-factor if and only if p + q is even. T(G) Corollary_3.9b. Let G be a connected graph. Then n T(G) G (n 2_ 2) has a l-factor if and only if has even order. Gn (n 3.2) An interesting special case of Corollary 3.9a occurs when G ' = M 2: +1 15 the complete graph Kp. Here we have q 2 and p + q E-(Ez—A, so that we may conclude that T(Kp) has a l-factor if and only if p E O or 3 (mod 4). Whenever a graph H has a 2-factor, by definition it contains a collection of disjoint cycles C1, C2, ..., Ck which span H. If H is an nth subdivision graph Sn(G) of a graph G, then it turns out that Sn(G) is composed of these cycles only. Theorem 3.10. For any graph G, Sn(G) has a Z-factor if and only if each component of G is a cycle. Proof. We may assume G is connected, for otherwise we may work with each component separately. Suppose first that Sn(G) has a 2-factor 56 F; then F is a collection of disjoint cycles C1, C2, ..., Ck which span 511(6). Now if any point u EC] is adjacent to a point VEC], then deg v 313 implies that v corresponds to a point of G, so that deg u = 2. But then the cycle C1 containing u must also contain v, which means that C1 = C], violating the choice of u. Thus k = 1, that is, C1 is a spanning subgraph of Sn(G). Furthermore, if two points v1 and v2 in C1 are joined by a diagonal edge (an edge not belonging to C] which joins two points of C1), then deg v1 f 2 for i = l, 2; this is a contradiction since by Theorem 3.1, every vl-v2 path has length at least n + 1. Thus C = Sn(G) implies that G is 1 also a cycle. The converse is immediate. Corollary 3.10. The following statements are equivalent for any graph G: (l) Sn(G) is l-factorable. (2) Sn(G) is 2-factorable. (3) Each component of G is a cycle. Section 3.4 The Enumeration of Trees Having a l-Factor In the course of this research a considerable number of enumeration problems have arisen. Some of these problems, for example the enumeration of homeomorphically irreducible trees, have been dealt with elsewhere (see [11]). Although extensive considera- tion of the area of graphical enumeration is beyond the focus of this dissertation, we do present one such problem as an example: the number of trees of given order which possess a l-factor. 57 Before proceeding further, we require some additional definitions. A rooted graph is a graph in which one of its points, called the root, is distinguished from the others. Thus two rooted graphs are isomorphic if there exists a one-to-one correspondence between their vertex sets which preserves not only adjacency, but also the roots. I h—z Figure 3.4. The 10 nonisomorphic rooted trees of order 6 or less having a l-factor. The generatinggfunction T(x) for rooted trees possessing a l-factor is defined as T(x) = iaZkXZk, where a2k is the number of nonisomorphic rooted trees of :rJer 2k having a l-factor. The generating function t(x) for (unrooted) trees possessing a l-factor is similarly defined ad t(x) = ib2kx2k, where b2k is the number of nonisomorphic trees of orderk2A having a l-factor. Figure 3.4 displays the nonisomorphic rooted trees of orders 2, 4, and 6 which 58 have a l-factor--each tree has the edges of its l-factor darkened and its root encircled. Thus by direct observation we conclude that T(x) = x2 + 2x4 + 7x6 + .... E. M. Palmer has suggested an approach originated by P61ya [l6] and refined by Otter [14], details of which can be found in Chapter 3 of [11], which shows that T(x) and t(x) satisfy the follow- ing functional equations: 21 T(x) V x), and 11x) - 1/2111x) - V(le) w k T(x where V(x) = x exp Z—E—A . k=l From these equations we are able to obtain the following by direct t(x) computatibn. V(x) = x + x3 + 3x5 + 10x7 + 39x9 + 160xn + 702x13 + 3177x15 + ... T(x) = x2 + 2x4 + 7x6 + 26x8 + 107x10 +458x12 + zossx14 + 9498x16 + ... t(x) = x2 + x4 + 2x6 + 5x8 + 15x10 + 49x12 + 180x14 + 7le16 + ... If desired, the correctness of the first 5 coefficients of t(x) can easily be checked firsthand by means of the tree diagrams found in [9], Appendix 3. 59 Section 3.5 The Edge Independence Number for Sn(G) For any graph G, the edge independence number B](G) of G is the maximum number of edges among all the l-regular subgraphs of G, or equivalently, the maximum cardinality of any independent set of __ edges in G. For example, 81(K(m,n)) = min (m,n). The observation néw~1 that a graph G has a l-factor if and only if 81(6) = [GI/2 suggests - that the parameter 6] can be viewed as a generalization of the l-factor concept. In general, |Gl/2 is an upper bound for 31(9), but when G j has no l-factor, 81(G) may be much less than this. For example, L“‘b K(l,p-l) is connected and has arbitrarily large order p, yet 81(K(1.p-1)) =1. If, for any real number r, we denote by [r] the greatest integer not exceeding r, then without specializing the graph G under consideration, the best possible upper bound for 81(G) is [|G|/2]. Letting f(g) denote either of the graph—valued functions T(G) or Gn (n 3_2), we next show that 81(f(G)) invariably attains this upper bound, thereby generalizing Corollary 3.9b. Theorem 3.11. Let G be a connected graph and f(G) denote either one of the graph-valued functions Gn (n 3_2) or T(G). Then 81(f(G)) = [lf(G)|/21- Epggf, Suppose that G is a connected (p,q)-graph and f(G) is either Gn (n 3_2) or T(G). In the event that |f(G)| is even, the assertion reduces to Corollary 3.9b. Assume, then, that |f(G)| is odd. 60 For f(G) = Gn (n 3_2), |f(G)| = p, and since G2 is a subgraph of G", it suffices to show that 81(62) = 9—5—1 = [43-]. Let v be any point in G which is not a cutpoint of G (this is impossible only if G = K], in which case there is nothing to prove). Since G-v is connected and (G-v)2 has even order, by Corollary 3.9b, 81((G-v)2) = E—%—l. Since (G-v)2 is a subgraph of G2, 81(G2) = E—%—l- TEEIi also. ' I For f(G) = T(G) = 52(G), since S(G) is a connected graph of _ order p+q, 81(T(G)) = 81(52(G)) = [E—g—H] = [AI%?11] follows from the : j result shown in the preceding paragraph. i-ww Corollary 3.11. Let G be a graph having components C1, C2, ..., Ck, and let f(G) denote either one of the Tr2ph57alued functions Gn k f C. (n a 2) or T(G). Then 31mm) = Z ———2-‘—. i=1 As regards the nth subdivision graph S11(G), Theorem 3.7 and Lemma 3.1 generalize to the following. Theorem 3.12. Let G be a connected graph and k any positive integer. Then, (1) 81(82k(G)) = [|52k(G)|/2] if and only if 81(G) = [|G|/2], and (2) 81(52k_](G)) = 1152k_,(e)1/21 if and oniy ii 81(S(G)) = [|s1e)|/21. Proof. Let G be a connected (p,q)-graph, so that |S(G)| = p + q, ISZk-l(G)| = p + (2k - l)q, and |52k(G)| = p + 2kq. Each of the graphs 51(6), 1 = 1, 2k - l, or 2k, is a union of q disjoint v1-vj 61 paths P11 where each path P11 corresponds to the edge Vivj in G and V = {v1, v2, ..., vp} is the set of points in S1(G) corresponding to the points of G. For part (1). When p is even, (1) is a restatement of Theorem 3.7; we therefore assume that p is odd. For the sufficiency portion of (l), we assume B](G) = (p - 1)/2 and seek to prove that 81(52k(G)) = (p + 2kq - l)/2. Let F be a set of (p - l)/2 independent edges in G. Each of the (p - l)/2 v1-vj paths of length 2k + l in $2k(G) corresponding to edges of F yield k + 1 independent edges; denote the set of these (p - l)(k + l)/2 independent edges by F'. Each of the remaining q - (p - l)/2 v1-vj paths yield k additional independent edges which are mutually nonadjacent with the edges in F', so that altogether $2k(G) contains (p — l)(k + l)/2 + (q - (p - l)/2)k = (p + 2kq - l)/2 independent edges. Conversely, if 81(52k(G)) = (p + 2kq - l)/2, to show that 81(6) = (p - l)/2, assume to the contrary that 81(6) §_(p - 3)/2. Then no more than (p - 3)/2 of the v1-vj paths of length 2k + 1 in $2k(G) can fail to share a common point (otherwise the edges in G corresponding to these paths form a set of independent edges with cardinality greater than (p - 3)/2) and each such path yields at most k + 1 independent edges; denote the set of all such independent edges from these paths by F'. Each of the remaining v1-vj paths of length 2k + 1 contains at least one point common to an edge of F', so that each such path yields at most k additional independent edges which are mutually nonadjacent with the edges in F'. Thus the 62 number of independent edges in $2k(G) is no more than (p - 3)(k + l)/2 + (q - (p - 3)/2)k = (p + 2kq - 3)/2, a contradiction. Hence 81(6) = (p -1)/2. For part (2). When p + q is even, (2) becomes Lemma 3.1; we therefore assume p + q is odd (so that p + (2k - l)q is also odd). Supposing that 81(S(G)) = (p + q - l)/2, we must check that 81(52k-l(6)) = (p + (2k - l)q - l)/2. All edges in S(G) have the form ”ivj’ where vjEV and u1. GEV, so that for any set F of (p + q - l)/2 independent edges in S(G), every point of S(G), except one, belongs to an edge in F. Furthermore, the edges of F determine (p + q - l)/2 v1-vj paths P15 of length 2 in S(G), each of which con- tributes one edge to F, and q - (p + q - l)/2 other vk-v1 paths Pkl’ each of which contains no edge of F; thus the points vk and v1 of such paths Pkl must belong to edges in F (otherwise S(G) contains more than (p + q - l)/2 independent edges). Consequently, in SZk-l(G) each path of length 2k corresponding to some Pij yields k independent edges and each path of length 2k corresponding to some Pkl yields (k - 1) additional independent edges, so that altogether we have a set of k(p + q - l)/2 + (k - l)(q - (p + q - l)/2) = l/2(p + (2k - l)q - 1) independent edges. The necessity portion of (2) follows immediately by taking k = 1. This completes the proof of Theorem 3.12. Theorem 3.12(2) is illustrated in Figure 3.5 with a graph G having order 6 and 7 edges for which 81(S(G)) = [13/2] = 6 and 31(s3(s)) = [1/2(5 + 3(7))1 = [27/2] = 13 (the edges in the independent 63 '0 3 3 (5 5(6) 53(6) Figure 3.5. A (6,7)-graph G along with S(G) and 53(3)- sets of maximum cardinality are darkened). Figure 3.5 also illustrates another result that applies to Sn(G), for an an odd integer, which we state next. Theorem 3.13. Let G be a connected (p,q)-graph. Then kq if G is a tree. 81(52k'](6)) = kq + p - q otherwise. Egggf, If the connected (p,q)-graph G is a tree, then we shall prove that 81(52k-1(G)) = kq by showing that for every tree with p points, a maximal independent set F of edges in 52k-1(G) has cardinality kq and that for any endpoint w in SZk-T(G)’ the edges u1v1, for l §_i §_kq, of F may be selected so that d(w,u1) is odd and d(w,v1) = d(w,u1) + 1. We proceed using induction on the order p of G. The assertion is immediate for trees of order 1 or 2. Now assume the assertion holds for trees on fewer than p points and consider a tree G with order p. For any endpoint v of G, 64 let G' denote the tree on p - 1 points obtained from G by deleting v and its incident edge uv; by the induction assumption, an independent set F' of maximum cardinality of edges in $2k_](G') has k(p - 2) members u1v1, for l g-i §;k(p - 2), which can be chosen in such a way that for any endpoint w in $2k_](G') which is distinct from u, d(w,u1) is odd and d(w, v1) = d(w,u1) + 1. Since 52k-1(G) can be regarded as formed from 52k-](G') by the addition of a u-v path P of even length 2k, it follows that the desired largest independent set F of edges for 52k-l(G) results by annexing the k alternate edges of P, beginning with the edge incident with v, to F'. Next we consider connected graphs which contain at least one cycle. Proceeding by induction on the number of edges q, we shall prove that for any such (p,q)-graph G, each largest independent set F of edges of 52k-1(G) has cardinality p + (k - l)q and contains an edge incident with each point of 52k-1(G) which corresponds to a point in G. In particular, we have that B](S2k_](G)) = kq + p - q. If G is unicyclic (so that p = q), then Theorem 3.8 shows that 52k-1(G) has a l-factor F, hence F is a largest independent set of edges of cardinality l/2|52k_](G)| = 1/2(p + (2k - l)q) = kq = kq+p-q. The induction begins with q = 3; the only appropriate graph having 3 edges is K3, a unicyclic graph, which therefore satisfies the induction assumption. So now let G be a connected (p,q)-graph with two or more cycles and let 6' denote the connected 03,q-l)-graph obtained from 65 G by deleting an edge uv not belonging to every cycle in G. Then G' contains at least one cycle and, by the induction assumption, every independent set F' of k(q - l) + p - (q - l) edges in $2k_1(G') contains an edge incident with each point of 52k_](G') corresponding to a point in G'. Since 52k-1(G) differs from $2k_](G') only by the presence of a u-v path P of length 2k and both u and v are incident with members of F', then an independent set of F edges with maximum cardinality for 52k-1(G) is obtained by adding to some set F' any collection of k - 1 mutually nonadjacent edges from P, none of which is incident with either u or v. Then |F| = k(q - l) + p - (q - l) + (k - l) = kq + p - q and, moreover, every point of SZk-l(G) cor- responding to a point in G is incident with an element of F. This completes the proof of Theorem 3.13. In the case of a connected (p,q)-graph G. Theorem 3.8 is a corollary to Theorem 3.13. For if the graph 52k-1(G) has a l-factor, then since its order is p + (2k - l)q, we have l/2(p + (2k - l)q) = 81(6) = kq + p - q (G is not a tree since that would imply that p + (2k - l)q is odd). Simplifying, we obtain p = q, which means that G is unicyclic. Conversely, if G is unicyclic, then 81(SZk-1(G)) = kq + p - q. But since p = q, this value is equal to l/2|52k_](G)|, implying that S G) has a l-factor. 2k-l( By combining Theorem 3.13, Theorem 3.7, and the well known facts that the complete graph Kp has a l-factor if and only if p is even and that the complete bipartite graph K(r,s) has a l-factor if and only if r = s, we obtain the following corollaries. 66 Corollary 3.13a. Sn(Kp) has a l-factor if and only if either p = 3 and n is odd or both n and p are even. Corollary 3.13b. Sn(K(r,s)) has a l-factor if and only if either r = s = 2 and n is odd or r = s and n is even. The analogous result to Theorem 3.13 for 81(Sn(G))s with n an even integer, is simpler to state but somewhat more difficult to apply, as the value of 81(6) is required. Theorem 3.14. Let G be a connected (p,q)-graph. Then 81(52k(G)) = kq + 81(6). ' Proof. Suppose that F0 is any independent set of edges of G having cardinality |F0| = 31(6) = r. Let both the points of G and the points of $2k(G) corresponding to the points of G be labeled as v1, v2, ..., vp; to each edge Vivj in G, denote the corresponding v1-vj path of length 2k + l in $2k(G) by Pij' From each Pij Vivj in F0, we select the available k + l mutually nonadjacent edges; corresponding to an edge from each P11 corresponding to an edge Vivj not in F0, we select the k mutually nonadjacent edges, none of which is incident with v1 or vj. Clearly this selection is always possible. Let F1 denote this collection of r(k + l) + (q - r)k = kq + r independent edges in $2k(G). All that remains is to verify that F1 is an independent set of edges in $2k(G) having maximum cardinality. For any independent set F2 of edges in $2k(G) with maximum cardinality |F2| 3.|F1|, we obtain another independent set of edges 67 F3 from F2 having the same cardinality in the following way. The edges of F2 are distributed among the q paths P11 of $2k(G), some of these paths--say s in number--containing k + l of these edges each, while the other q-s paths each contain only k edges of F2. Whenever a vivj path P1j contains exactly k edges of F2, at least one of v1 and vj must be incident with an edge of F2 not in P11, since otherwise k + 1 edges could be used in F2 for that path P11, contradicting the maximality of |F2|. Replacing this set of k edges by the set of k mutually nonadjacent edges of P11, no member of which is incident with v1 or vj (if this latter set is distinct from the former) yields a maximal independent set of edges having the same cardinality as F2. Repeating this process for every path P11 con- taining only k edges of F2, we obtain the promised independent set F3 where |F3| = IFZI. Now choose a new independent set of edges in G consisting of the edges Vivj corresponding to the s v1-vj paths Pij containing k + l edges of F3. Since |F3| = s(k + l) + (q - s)k 3_r(k + 1) + (q - r)k = |F1| implies s 3_r, and since FO has maximum cardinality, we have s = r. Therefore, the cardinality of F1 is equal to that of F3, and hence to |F2| as well. Thus F1 is a largest independent set of edges in S2k(G). For connected (p,q)-graphs, Theorem 3.7 follows directly from Theorem 3.14. For if G has a l-factor (so that 81(6) = p/2), then 81(S2k(G)) = kq + p/2 = l/2|SZk(G)| implies that $2k(G) has a l-factor. Conversely, if $2k(G) has a l-factor, then 81(52k(6)) = 68 l/2|52k(G)| = kq + p/2 = kq + 81(G) implies that 81(G) = p/2, so that G also has a l-factor. Another application of Theorem 3.14 is obtained by setting G = Kp or G = K(r,s). This procedure yields well known necessary and sufficient conditions that G have a l-factor, as noted in the remarks preceding Corollary 3.13a. The formulas presented in Theorems 3.13 and 3.14 are suffi- ciently similar so that one is tempted to try to combine them. One such formulation is the following. Corollary 3.14. Let G be a connected (p,q)-graph. Then. ! {gig + 12[(-l)n + l]B](G), if G is a tree, 81(5n(G)) =1 qu+0'14VNp-n+<‘+gUDamhomema, where curly brackets denote the greatest integer function. 10. ll. 12. 13. BIBLIOGRAPHY Mehdi Behzad. Graphs and Their Chromatic Numbers. Ph.D. Thesis, Michigan State University (1965). Mehdi Behzad and Gary Chartrand. An Introduction to the Theory of Graphs, Allyn and Bacon, Inc., 1972. Gary Chartrand. "0n Hamiltonian Line Graphs.‘I Transactions of the American Mathematical Society 3 (1968), 559-566. Gary Chartrand; Dennis Geller; Stephen Hedetniemi. "Graphs with Forbidden Subgraphs." Journal of Combinatorial Theory 1 (1971), 12-41. Gary Chartrand and Frank Harary. "Planar Permutation Graphs." Les Annales de l'Institut Henri Poincare (Section B), 3 (1967), 433-438. Gary Chartrand; A. M. Hobbs; C. St. J. A. Nash-Williams; S. F. Kapoor. "The Square of a Block Is Hamiltonian Connected." (To appear.) Gary Chartrand and H. V. Kronk. "Randomly Traceable Graphs." SIAM Journal of Applied Mathematics 4 (1968), 696-700. H. Fleischner. "The Square of Every Non-separable Graph Is Hamiltonian Connected." (To appear.) Frank Harary. Graph Theory. Addison-Wesley Co., 1969. Frank Harary and C. St. J. A. Nash-Williams. "On Eulerian and Hamiltonian Graphs and Line Graphs." Canadian Mathematical Bulletin 8 (1965), 701-709. F. Harary and E. M. Palmer. Graphical Enumeration. Academic Press, New York, 1973. (To appear.) H. A. Jung. "Zu Einem Isomorphiesatz von Whitney ffir Graphen." Math. Ann. 164 (1966), 270-271. 0. Ore. "A Problem Regarding the Tracing of Graphs." Elem. Math. 6 (1951), 49-53. 69 F 14. 15. 16. 17. 70 R. Otter. "The Number of Trees." Ann. of Math. 49 (1948), 583-599. Wayne S. Petroelje and Curtiss E. Wall. "Graph-Valued Functions and Hamiltonian Graphs.“ Recent Trends in Graph Theory (M. Capobianco, J. B. Frechen, and M. Krolik, editors). Springer- Verlag, Berlin (1971), 211-213. G. Polya. "Kombinatorische Anzahlbestimmungen ffir Gruppen, Graphen und Chemische Verbindungen.“ Acta. Math. 68 (1937), 145-254. M. James Stewart. "A Topological Influence: Homeomorphically Irreducible Graphs." The Many Facets of Graph Theory (G. Char- trand and S. F. Kapoor, editors). Springer-Verlag, Berlin (1969), 281-282. M1111111111111111111111111111111ES