Litherland R. - Introduction To Graph Theory, Lecture Notes-Louisiana State Univ. (2009)
Litherland R. - Introduction To Graph Theory, Lecture Notes-Louisiana State Univ. (2009)
R.A. Litherland
Contents
0 Preliminaries 2
1 Graphs 5
2 Degree sequences 11
3 Connectedness 15
5 Blocks 24
6 Connectivity 29
7 Trees 34
8 Counting trees 36
9 Permutation groups 39
11 Eulerian graphs 45
12 Hamiltonian graphs 48
1
14 Planar graphs 52
15 Kuratowski’s Theorem 56
16 Crossing number 63
17 Vertex colorings 68
19 Edge colorings 75
21 Moore graphs 80
22 Kneser graphs 86
A More topology 92
0 Preliminaries
We assume that the reader is familiar with the rudiments of set theory,
including the notations x ∈ X to indicate that x is an element of the set
X and x ∈ / X to indicate that it is not, and the use of expressions such as
{x ∈ R | x > 1}, which denotes the set of all real numbers greater than
1. We recall here some (but not all) of the ideas the reader should have
encountered. Two sets are equal iff they have exactly the same elements.
There is a unique set with no elements, the empty set ∅. We use the notation
X ⊆ Y to indicate that X is a subset of Y ; that is, that x ∈ X implies x ∈ Y .
We have X = Y iff X ⊆ Y and Y ⊆ X. We use X ⊂ Y to indicate that X
is a proper subset of Y ; that is, X ⊆ Y and X 6= Y . The intersection, union
and difference (or relative complement) of sets X and Y are defined by
X ∩ Y = {x | x ∈ X and x ∈ Y },
X ∪ Y = {x | x ∈ X or x ∈ Y },
and X − Y = {x | x ∈ X and x ∈
/ Y },
2
with finite sets. The number of elements, or cardinality of a finite set X will
be denoted by |X|.
We also assume familiarity with the idea of a function, and a few related
notions. The notation f : X → Y means that f is a function from the set X
to the set Y ; that is, to each element x of X is associated a unique element
f (x) of Y . Functions f : X → Y and f ′ : X ′ → Y ′ are equal iff X = X ′ ,
Y = Y ′ and f (x) = f ′ (x) for all x ∈ X. For any set X, the identity function
on X is the function idX (or just id) from X to X defined by idX (x) = x
for all x ∈ X. If f : X → Y and A ⊆ X, the image of A under f is the
subset f (A) = {f (a) | a ∈ A} of Y . The function f is injective (or an
injection) if f (x1 ) = f (x2 ) implies x1 = x2 for x1 , x2 ∈ X. It is surjective
(or a surjection) if f (X) = Y , and it is bijective (or a bijection) if it is both
injective and surjective. If f is a bijection from X to Y , there is an inverse
function f −1 : Y → X defined by f −1 (y) = x iff y = f (x) for x ∈ X and
y ∈Y.
If f : X → Y and g : Y → Z, the composite g ◦ f : X → Z is defined by
(g ◦ f )(x) = g(f (x)). We also use gf to denote the composite. If f : X → Y
is a bijection we have f −1 ◦ f = idX and f ◦ f −1 = idY .
The rest of this section discusses equivalence relations and partitions,
which the reader may not have met before. A binary relation on a set X
is a subset R of X × X. If (x, y) ∈ R, we say that x is related to y by R,
and write x R y. Typically, specific binary relations are denoted by symbols
such as =, ≤, > and ∼ rather than letters. Thus, formally, the relation >
on the set {1, 2, 3} is the set {(2, 1), (3, 1), (3, 2)} of ordered pairs. Let R be
a binary relation on X.
Example 0.2. On the set R of real numbers, the relation ≤ is reflexive and
transitive, but not symmetric.
3
Example 0.3. On the set R, the relation < is transitive, but neither reflex-
ive nor symmetric.
Proof. First, x ∈ [x] because R is reflexive. Second, if [x] = [y] then y ∈ [x],
so x R y. Finally, suppose x R y. If z ∈ [y] then y R z, so x R z by transitivity.
This gives z ∈ [x], and so [y] ⊆ [x]. By symmetry of R, y R x, so also
[x] ⊆ [y]; that is, [x] = [y].
4
Figure 1: The Petersen graph
classes for an equivalence relation R then x R y iff hxi = hyi, which proves
the uniqueness part. To prove existence, define a relation R by x R y if
hxi = hyi. It is trivial to check that R is an equivalence relation. We must
show that the equivalence class [x] is equal to hxi. Suppose y ∈ [x]. Then
hxi = hyi, and since y ∈ hyi by definition, y ∈ hxi. Conversely, if y ∈ hxi,
then since also y ∈ hyi, hxi and hyi are not disjoint. Hence hxi = hyi, so
x R y and y ∈ [x].
1 Graphs
Definition 1.1. A graph G is a pair (V, E), where V is a non-empty finite
set, and E is a set of unordered pairs of elements of V . The elements of V
are called the vertices of G, and the elements of E are the edges of G. We
always assume that V ∩ E = ∅.
5
Figure 2: The complete graph on 5 vertices
u 6= v and uv ∈
/ E(G). If u and v are adjacent we shall also say that u and
v are neighbors.
6
G1 G2 G3
4, 11 and 34, respectively. (For order 5, particularly, you may want to use
the following definition to reduce the tedium.)
Proof. We can count the number p of pairs (v, e) where v is a vertex of the
edge e in two ways. First, there are m edges, each with two vertices,
P so p =
2m. Second, the vertex v is incident to deg v edges, so p = v∈V (G) deg v.
7
The minmum degree of the vertices of a graph is denoted by δ(G), and
the maximum by ∆(G). Clearly 0 ≤ δ(G) ≤ ∆(G) ≤ n(G) − 1. A vertex
of degree 0 is an isolated vertex, while one of degree 1 is an end-vertex. If
δ(G) = ∆(G) then G is regular ; if the common value is d, G is d-regular
or regular of degree d. A 3-regular graph is also called a cubic graph. The
complete graph Kn is (n − 1)-regular, and the Petersen graph of Figure 1
is cubic. For a regular graph, any two of the order n, the size m and the
degree d determine the third by the equation 2m = nd.
A subgraph of a graph G is a graph H with V (H) ⊆ V (G) and E(H) ⊆
E(G). We write H ⊆ G. If U is any proper subset of V (G), G − U is the
subgraph obtained by deleting all vertices of U and all edges with at least
one vertex in U . When U = {u}, we abbreviate G − {u} to G − u. Similarly,
if F is a subset of E(G), G−F is the subgraph obtained by deleting the edges
of F , and G−{e} is written as G−e. (This notation is logically indefensible,
since an edge is a set of vertices. Nevertheless, for an edge e = uv = {u, v},
we adopt the convention that G − e and G − uv denote G with just an edge
deleted, and G − {u, v} denotes G with two vertices and all incident edges
deleted.) Note that V (G − F ) = V (G), and F = E(G) is allowed, with
G − E(G) being the empty graph on V (G). A spanning subgraph of G is
a subgraph of the same order (i.e., containing all the vertices of G). These
are precisely the subgraphs of the form G − F for F ⊆ E(G). If u and v are
independent vertices of G and f = uv, the graph G + f has vertex set V (G)
and edge set E(G) ∪ {f }.
If U is a non-empty set of vertices of the graph G, the subgraph hU i
induced by U is the subgraph with vertex set U and edge set consisting of
all edges of G with both vertices in U . A vertex-induced subgraph (or just
an induced subgraph) is a subgraph of this form. If F is a non-empty set
of edges of G, the subgraph hF i induced by F is the subgraph with vertex
set all vertices of edges of F and edge set F . An edge-induced subgraph is a
subgraph of this form.
8
Figure 4: The complete bipartite graph K3,3
G1 G2 G1 ∗ G2
G1 × G2
9
Let G1 and G2 be graphs. Their union G1 ∪ G2 is the graph with vertex
set V (G1 )∪V (G2 ) and edge set E(G1 )∪E(G2 ). Their disjoint union G1 ⊔G2
is any graph of the form G′1 ∪ G2 where G′1 ∼ = G1 and V (G′1 ) is disjoint from
V (G2 ); it is well-defined up to isomorphism. The disjoint union of k copies
of a graph G is denoted kG. The join of G1 and G2 is the graph G1 ∗ G2
obtained from G1 ⊔ G2 by adding an edge v1 v2 for each vertex v1 of G1 and
v2 of G2 ; see Figure 5. Note that Ḡ1 ∗ Ḡ2 ∼ = G1 ⊔ G2 . The complete bipartite
graph Kr,s can be described as either K̄r ∗ K̄s or Kr ⊔ Ks . The Cartesian
product G1 × G2 has vertex set V (G1 ) × V (G2 ), and there is an edge with
vertices (u1 , u2 ) and (v1 , v2 ) iff either u1 v1 ∈ E(G1 ) and u2 = v2 , or u1 = v1
and u2 v2 ∈ E(G2 ). Figure 6 shows the product of the graphs G1 and G2 of
Figure 5.
Two classes of graphs we shall meet frequently are the paths and cycles.
The path Pn (n ≥ 1) of order n has vertices v1 , . . . , vn and edges vi vi+1
for 1 ≤ i < n. Of course, P1 ∼ = K1 and P2 ∼ = K2 . For n ≥ 3, the cycle
Cn is obtained from Pn by adding the edge vn v1 . It can be represented
as a regular n-gon. For small values of n we refer to Cn as a triangle,
quadrilateral, pentagon, . . . . Another special class of graphs consists of the
cubes. The n-cube Qn (n ≥ 0) is defined recursively by setting Q0 = K1
and Qn+1 = Qn × K2 . It may also be defined as the graph with vertices all
sequences of zeroes and ones of length n, with two sequences being adjacent
iff they differ in exactly one place.
Exercises for §1
1.2. Show that in any non-trivial graph there are two distinct vertices with
the same degree.
10
1.3. Let G be a self-complementary graph, and let G′ be obtained from
G ⊔ P4 by adding edges from every vertex of G to the two end-vertices
of P4 . Show that G′ is self-complementary, and deduce that there are
self complementary graphs of order n for every positive integer n with
n ≡ 0 or 1 (mod 4).
1.6. For any integer k ≥ 2, show that there are k non-isomorphic regular
graphs, all of the same order and degree. [Hint: the disjoint union of
regular graphs of the same degree is regular.]
1.7. Show that the minimum order of a 3-regular graph containing the graph
G below as an induced subgraph is 6. Also draw the 3-regular graph
constructed from G as in the proof of Theorem 1.6.
2 Degree sequences
Let G be a graph with vertices v1 , . . . , vn , and set di = deg vi . The sequence
(d1 , . . . , dn ) is called a degree sequence of G. A sequence (d1 , . . . , dn ) of
integers is graphical if it is a degree sequence of some graph. Necessary
conditions for the sequence to be graphical are that 0 ≤ di ≤ n − 1 for all i
and (by Theorem 1.4) that d1 + · · · + dn is even. These are not sufficient; it
is easy to see that (3, 3, 3, 1) is not graphical.
is graphical.
11
Evidently this theorem gives an algorithm for determining whether a
sequence d of integers is graphical. The only graphical sequence of length
1 is (0), any sequence consisting entirely of zeroes is graphical, and any
sequence with negative terms is not graphical. For any other sequence,
arrange it in descending order. If d1 ≥ n, it is not graphical. Otherwise,
form the sequence d′ as above, and repeat the process until a decision is
reached. For d = (3, 3, 3, 1), we get d′ = (2, 2, 0) and then d′′ = (1, −1), so
indeed d is not graphical.
Proof. We first show the necessity of the conditions. That ni=1 di must be
P
even has already been remarked. Let G be a graph of order n with vertices
v1 , . . . , vn for which deg vi = di , 1 ≤ i ≤ n, and let 1 ≤ k ≤ n − 1. Consider
all ordered pairs (i, j) with 1 ≤ i ≤ k and 1 ≤ j ≤ n for which vi vj ∈ E(G).
12
For 1 ≤ i ≤ k there are di such pairs, so the total number of pairs is ki=1 di .
P
The number of such pairs with j ≤ k is at most k(k − 1), and for each j > k
the number of such pairs is at most min{dj , k}, so (EG) holds.
For the converse, we prove the stronger result that if ni=1 di is even
P
and (EG) holds for all k with 1 ≤ k ≤ dk then G is graphical. (It is easy
to see that these cases of (EG) imply the remainder, but we do not need
this.) If n = 2, the only such sequences are (0, 0) and (1, 1), which are
graphical. Suppose then that n > 2 and the result holds for sequences of
length n − 1. If d1 = 0 then d is the degree sequencePof an empty graph.
Suppose that d1 > 0. The case k = 1 of (EG) is d1 ≤ ni=2 min{di , 1}, and
the right-hand side of this inequality is the number of i ≥ 2 with di 6= 0.
Thus dˆ = dd1 +1 is positive. Let a and c be the least and greatest elements
ˆ We have a ≤ d1 + 1 ≤ c, so
of the set of those i ∈ {2, . . . , n} with di = d.
setting b = a + c − d1 − 1 we have a ≤ b ≤ c. Partition the set {2, . . . , n}
into R = {i | 2 ≤ i < a or b ≤ i ≤ c} and S = {i | a ≤ i < b or c < i ≤ n},
and note that |R| = d1 . For 2 ≤ i ≤ n, let d′i = di − 1 if i ∈ R and
d ′ = d if i ∈ S. The sequence d′ = (d′ , . . . , d′ ) has d′ ≥ · · · ≥ d′ ≥ 0 and
i i 2 n 2 n
P n ′
Pn
i=2 i =
d i=1 di −2d1 . (It is just the sequence of Theorem 2.1 arranged in
descending order.) We show that it satisfies the inequalities corresponding
to (EG), namely
k
X n
X
d′i ≤ (k − 1)(k − 2) + min{d′i , k − 1},
i=2 i=k+1
for 1 ≤ k −1 ≤ d′k . Once this has been done, it will follow from the inductive
hypothesis that d′ is graphical, and then that d is graphical as in the earlier
proof. For i ≥ c we have
d′i ≤ d′c = dˆ − 1 ≤ d1 − 1 ≤ c − 2 ≤ i − 2,
so k − 1 ≤ d′k implies k < c. Suppose first that k < a. If k ≤ d, ˆ then
′
min{di , k − 1} = k − 1 = min{di , k − 1} for i ∈ R, while clearly min{d′i , k −
1} = min{di , k − 1} for i ∈ S. Hence
k
X k
X k−1
X
d′i = di − k + 1 ≤ di − k + 1
i=2 i=2 i=1
Xn
≤ (k − 1)(k − 2) + min{di , k − 1} − k + 1
i=k
Xn
= (k − 1)(k − 2) + min{d′i , k − 1}.
i=k+1
13
If k ≥ dˆ + 1, then min{d′i , k − 1} = di = min{di , k} for i ∈ S, and of course
min{d′i , k − 1} = min{di , k} − 1 for i ∈ R. Therefore
k
X k
X
d′i = di − d1 − k + 1
i=2 i=1
n
X
≤ k(k − 1) + min{di , k} − d1 − k + 1
i=k+1
Xn
= k(k − 1) + min{d′i , k − 1} + (d1 − k + 1) − d1 − k + 1
i=k+1
n
X
= (k − 1)(k − 2) + min{d′i , k − 1}.
i=k+1
We have
k
X k
X
d′i ≤ (a − 2)(d1 − 1) + d′i
i=2 i=a
= (a − 2)(d1 − 1) + (k − a + 1)dˆ − xk ,
14
where yk is 0 in cases 1 and 3, and c − b + 1 in case 2. It is therefore enough
to prove that zk ≥ 0, where
zk = (a − 2)(c − d1 − 1) + (k − a + 1)(c − dˆ − 2) − (c − b + 1)
= (a − 2)(c − d1 − 1) + (k − a)(c − dˆ − 2) + b − dˆ − 3.
zk = (a − 2)(c − d1 − 1) + (k − a + 1)(c − dˆ − 2) + k − b + 1,
Exercises for §2
2.1. Let n and d be integers with 0 ≤ d < n and dn even. Prove that there
is a d-regular graph of order n.
3 Connectedness
Definition 3.1. A walk in a graph G is a sequence W = (u0 , u1 , . . . , ul )
(l ≥ 0) of vertices of G such that ui−1 and ui are adjacent for 1 ≤ i ≤ l.
The length of W is l (the number of edges on W rather than the number
of vertices). We say that W is a u0 -ul walk, or a walk from u0 to ul . If
15
u0 = ul , W is a closed walk; otherwise it is open. A trail is a walk with
no repeated edges (ui ui+1 6= uj uj+1 for i 6= j). A path is a walk with no
repeated vertices (ui 6= uj for i 6= j). Every path is a trail, and for any
vertex u the trivial u-u walk (u) is a path. A circuit is a non-trivial closed
trail, and a cycle is a closed walk (u0 , u1 , . . . , ul ) of length l ≥ 3 such that
ui 6= uj for 1 ≤ i < j ≤ l. Every cycle is a circuit, and every circuit has
length at least 3.
W = (u = u0 , u1 , . . . , ul = v)
is a u-v walk, and assume inductively that every u-v walk (if any) of length
less than l contains a u-v path. If W is a path, there is nothing to do.
Otherwise, there exist i and j with 0 ≤ i < j ≤ l and ui = uj . Now W
contains the u-v walk (u0 , u1 , . . . , ui , uj+1 , . . . , ul ) of length less than l, which
in turn contains a u-v path.
Note that by Theorem 3.2, u and v are connected iff there is a u-v
walk. If W = (u0 , u1 , . . . , ul ) is a walk, we denote by W r the reverse walk
(ul , ul−1 , . . . , u0 ). If W ′ = (v0 , v1 , . . . , vk ) is another walk with v0 = ul we
denote by W · W ′ the walk (u0 , u1 , . . . , ul , v1 , . . . , vk ).
Proof. Let u, v and w be vertices of G. The trivial u-u path shows that
u ∼ u. If u ∼ v, let P be a u-v path. Then P r is a v-u path, so v ∼ u. If
also v ∼ w, let Q be a v-w path. Then P · Q is a u-w walk, so u ∼ w.
16
The subgraphs of G induced by the equivalence classes for the relation
∼ are called the components of G. Each component is connected, and every
connected subgraph of G is contained in a component. The number of
components is denoted by k(G). Thus k(G) = 1 iff G is connected.
By Theorem 3.2, d(u, v) is also the minimum of the lengths of all walks
from u to v. In a disconnected graph, we may define d(u, v) as above for
vertices in the same component, and set it to ∞ for vertices in different
components.
Proof. (1) is obvious. The other parts follow by examining the lengths of
the walks used to show symmetry and transitivity in the proof of Lemma
3.4.
17
the same set of the partition; that is, with d(u, w) ≡ d(v, w) (mod 2). Let
(w = u0 , u1 , . . . , uk = u) and (w = v0 , v1 , . . . , vl = v) be geodesics, so that
k ≡ l (mod 2). For 0 ≤ i ≤ j ≤ k, d(ui , uj ) = j − i, since otherwise we
could replace the subpath (ui , . . . , uj ) by a shorter path to obtain a shorter
w-u walk. Similarly, d(vi , vj ) = j − i for 1 ≤ i ≤ j ≤ l. Hence if ui = vj then
i = d(w, ui ) = d(w, vj ) = j. Let i be the greatest integer for which ui = vi .
Since u 6= v, (k −i)+(l −i) ≥ 1, and since k ≡ l (mod 2), (k −i)+(l −i) ≥ 2.
Now (ui , . . . , uk−1 , u, v, vl−1 , . . . , vi ) is a cycle of odd length k +l −2i+1.
The radius rad G of G is the minimum of e(u) over all vertices u, while the
diameter diam G is the maximum of e(u). We also have
Proof. That rad G ≤ diam G is immediate from the definitions. For the
other inequality we must show that d(u, v) ≤ 2 rad G for any vertices u and
v. If w is a central vertex we have
Theorem 3.13 (Hedetniemi; see [5]). Any graph G is the center of some
connected graph.
18
Proof. Form a graph H by adding to G four new vertices v1 , v2 , w1 and w2 ,
and edges uvi and vi wi for u ∈ V (G) and i = 1 or 2. Then H is connected,
eH (u) = 2 for u ∈ V (G), eH (v1 ) = eH (v2 ) = 3 and eH (w1 ) = eH (w2 ) = 4,
so G = Cen H.
Theorem 3.14 (Bielak and Syslo [3]). A graph G is the periphery of some
connected graph iff G is complete or no vertex of G has eccentricity 1.
Theorem 3.16 (Hendry [18]). For any graphs G1 and G2 , there is a con-
nected graph H with Cen H ∼
= G1 and Med H ∼
= G2 .
Proof. We may assume that V (G1 ) and V (G2 ) are disjoint. For i = 1 or
2, set ni = n(Gi ) and δi = δ(Gi ). Let a and b be positive integers with
a > (n2 − n1 − δ2 − 3)/2 and b > 2a + n1 + n2 − δ2 . Form H by adding to
G1 ∪ G2 vertices xi for 1 ≤ i ≤ a + 2, yi for 1 ≤ i ≤ a and zi for 1 ≤ i ≤ b,
and edges
19
Since b − degG2 u2 + δ2 > 0, u2 is adjacent to z1 , and so d(u2 , zi ) ≤ 2 for
1 ≤ i ≤ b. Since b − degG2 u2 + δ2 ≤ b, the number of zi with d(u2 , zi ) = 1 is
b − degG2 u2 + δ2 . The remaining distances in H are given below; for those
involving a vertex zi , note that zi is adjacent to at least one vertex of G2 ,
namely a vertex of minimum degree in G2 . We have
Hence
e(xi ) = i + a + 2 for 1 ≤ i ≤ a + 2;
e(u1 ) = a + 2 for u1 ∈ V (G1 );
e(yi ) = i + a + 2 for 1 ≤ i ≤ a;
e(u2 ) = 2a + 3 for u2 ∈ V (G2 ); and
e(zi ) = 2a + 4 for 1 ≤ i ≤ b,
so Cen H = G1 . Further,
i−1
X a−i+2
X i+a
X
td(xi ) = j+ j + in1 + j + (i + a + 1)n2 + (i + a + 2)b
j=1 j=1 j=i+1
for 1 ≤ i ≤ a + 2;
a+2
X a
X
td(u1 ) ≥ j + n1 − 1 + j + (a + 1)n2 + (a + 2)b
j=1 j=1
20
for u1 ∈ V (G1 );
i+a+2
X i−1
X a−i
X
td(yi ) = j + in1 + j+ j + (a − i + 1)n2 + (a − i + 2)b
j=i+1 j=1 j=1
for 1 ≤ i ≤ a;
2a+3
X a
X
td(u2 ) = j + (a + 1)n1 + j + 2n2 + b − δ2 − 2
j=a+2 j=1
21
Exercises for §3
3.1. Prove that any graph G contains a path of length δ(G).
The degree matrix of G is the n×n diagonal matrix D = D(G) with diagonal
entries dii = deg vi , and the Lagrangian matrix is L = L(G) = D − A.
Finally, if m > 0, the incidence matrix of G is the m × n matrix B = B(G)
with entries (
1 if ei is incident to vj ;
bij =
0 if not.
Evidently an n × n 0-1 matrix (a matrix with all entries 0 or 1) is the
adjacency matrix of some graph iff it is symmetric and has all diagonal
entries 0, and an m × n 0-1 matrix is the incidence matrix of some graph iff
every row has exactly two non-zero entries and different rows are unequal.
Either the adjacency or the incidence matrix determines the graph up to
isomorphism.
Theorem 4.2. Let G be a graph of order n with vertices v1 , . . . , vn and
(l) (l)
adjacency matrix A = [aij ]. For any integer l ≥ 0, let Al = [aij ]. Then aij
is the number of vi -vj walks of length l.
Proof. This is by induction on l, the case l = 0 being obvious. Suppose the
result holds for some l. Then
n
(l+1) (l) (l)
X X
aij = aik akj = aik ,
k=1 k:vk vj ∈E(G)
22
Theorem 4.3. Let G be a graph of order n with vertices v1 , . . . , vn and adja-
cency matrix A. If G is d-regular, A has x1 = (1, 1, . . . , 1) as an eigenvector
of eigenvalue d, and for any eigenvalue λ, |λ| ≤ d. If also G is connected,
the multiplicity of the eigenvalue d is 1.
where vj(i,1) , vj(i,2) , . . . , vj(i,d) are the vertices adjacent to vi . Take i so that
|xi | is a maximum. Then |λxi | ≤ dk=1 |xj(i,k) | ≤ d|xi |, so |λ| ≤ d. If λ = d,
P
we must have
xj(i,1) = xj(i,2) = · · · = xj(i,d) = xi .
It follows that xj = xk if vj and vk are in the same component, completing
the proof.
Exercises for §4
23
5 Blocks
Definition 5.1. A vertex v of a graph G is a cut-vertex if G is non-trivial
and k(G − v) > k(G). A graph is non-separable if it is connected and has
no cut-vertex; otherwise it is separable.
24
Proof. Let e = uv. Suppose that e lies on a cycle C. For any two vertices
w and x in a single component of G, take a w-x path in G. If it contains
the edge e, replace e by the other u-v or v-u path in C to get a w-x path in
G − e. Hence e is not a bridge.
Now suppose e is not a bridge. Then there is a u-v path P in G − e, and
P · (v, u) is a cycle in G containing e.
(1) G has no isolated vertices and any two edges of G lie on a common
cycle.
(2) G is non-empty and any vertex and edge of G lie on a common cycle.
(4) G is non-separable.
Proof. The implications (1) ⇒ (2) ⇒ (3) are trivial. Suppose that (3) holds.
Certainly G is connected. Let u be any vertex of G, and v and w vertices
distinct from u. There is a cycle C containing v and w, and at least one
of the two v-w paths in C does not contain u, so u is not a cut-vertex by
Theorem 5.2. Thus G is non-separable.
Suppose that G is non-separable, and let e1 and e2 be edges of G. Let
k be the minimum distance from a vertex of e1 to a vertex of e2 . We prove
by induction on k that there is a cycle containing e1 and e2 . Let e1 = u1 v1
and e2 = u2 v2 , where d(u1 , u2 ) = k. We show first that there is a cycle C
containing e1 and u2 . If k = 0, this is because G has no bridges, so there
is a cycle containing e1 . If k > 0, there is a vertex w adjacent to u2 with
d(u1 , w) = k − 1, and by inductive hypothesis there is a cycle containing e1
and the edge u2 w.
25
If e2 is on C, we are done, so suppose it is not. Since u2 is not a cut-
vertex, there is a path P in G − u2 from v2 to a vertex x of C. Taking P
to have minimum length, x is the only vertex of P on C. Let Q be the x-u2
path in C containing e1 . Then P · Q · (u2 , v2 ) is the desired cycle.
(1) Two distinct blocks of G meet in at most a single vertex, which must
be a cut-vertex of G.
(3) If v and w are vertices such that no block of G contains both v and w,
there is a cut-vertex u such that v and w are in different components
of G − u.
26
cut-vertex of G different from u. Consider two distinct blocks of G. If they
lie in the same Gi , they meet in at most a single vertex which is a cut-vertex
of Gi , and if not they meet in at most u. Every cut-vertex of G other than
u is in at least two blocks of some Gi , and u is in k ≥ 2 blocks of G, one
from each Gi . Finally, suppose v and w are vertices contained in no block
of G. If they lie in a single Gi , there is a cut-vertex u′ of Gi such that v and
w lie in different components of Gi − u′ , and hence of G − u′ . Otherwise, v
and w lie in different components of G − u.
Proof. Let u and v be vertices such that no block of G contains both u and
v, and let w be the cut-vertex provided by Lemma 5.13. Let x be a vertex
with d(w, x) = e(w). At least one of u and v is in a different component of
G − w from x; say u is. Then e(u) ≥ d(u, x) = d(u, w) + d(w, x) > e(w), so
u is not a central vertex. The result follows.
The block graph of G is the disjoint union of the block graphs of its
components, so it is enough to consider Blk G for G connected. By (2) of
Lemma 5.13, the end-vertices of Blk G are precisely the end-blocks of G.
27
edge or non-cut-vertex of G is contained in a unique block, each sub-path
from uij−1 to uij is contained in a block Bj . Now
(u, B1 , ui1 , B2 , . . . , uis−1 , Bs , v)
is a u-v path in Blk G.
It remains to prove that Blk G is acyclic. If not, it contains a cycle
(B0 , u1 , B1 , . . . , uk , Bk = B0 )
where each Bi is a block and each ui is a cut-vertex. Let H = B1 ∪ · · · ∪ Bk ,
which is clearly connected. Let v be a vertex of H, and let Bi′ be Bi − v if
v is in Bi , and Bi otherwise. Each Bi′ is connected, and if v 6= ui then Bi−1 ′
′
and Bi have a common vertex, and hence lie in a single component of H − v.
Since v is equal to at most one ui , H − v = B1′ ∪ · · · ∪ Bk′ is connected, so H
is non-separable. This contradicts the maximality of the blocks Bi .
The next result is immediate from the previous one and Theorem 5.8.
Theorem 5.18. A connected graph with a cut-vertex contains at least two
end-blocks.
Theorem 5.19. Let G be a connected graph with a cut-vertex. Then there
is a cut-vertex u such that at most one block of G containing u is not an
end-block.
Proof. Let T be the result of deleting from Blk G all its end-vertices. Then T
is also a tree. If T is trivial, then G has a unique cut-vertex u and all blocks
of G are end-blocks, so u has the desired property. Otherwise, T has an
end-vertex, which must be a cut-vertex u of G since blocks are not adjacent
in Blk G. Now every block of G containing u is an end-block, except for the
unique vertex of T adjacent to u.
28
Exercises for §5
5.3. Prove that a graph with only even vertices has no bridge.
6 Connectivity
Definition 6.1. A vertex-cut of a graph G is a proper subset U of V (G)
such that G − U is disconnected.
Theorem 6.3 (Whitney [31]). For any graph G, κ(G) ≤ κ1 (G) ≤ δ(G).
29
Let the orders of G, G1 and G2 be n, n1 and n2 . Then (n1 − 1)(n2 − 1) ≥ 0,
so |X| = n1 n2 ≥ n1 + n2 − 1 = n − 1 ≥ κ(G). Otherwise, let u ∈ V (G1 ) and
v ∈ V (G2 ) be non-adjacent. For every edge e of X, pick a vertex f (e) of e
different from u and v, and set U = {f (e) | e ∈ X}, so |U | ≤ |X|. Then U
is a vertex-cut, because u and v are vertices of G − U , and every u-v path
in G must contain an edge e of X, and therefore the vertex f (e) of U . We
have proved that κ(G) ≤ κ1 (G) in all cases. The set of edges incident with
any vertex is an edge-cut, which gives κ1 (G) ≤ δ(G).
Proof. Let k = κ1 (G), and let X be an edge-cut with |X| = k. The graph
G − X has two components H1 and H2 . We show first that either every
vertex of H1 is adjacent to a vertex of H2 , or every vertex of H2 is adjacent
to a vertex of H1 . Suppose there is a vertex u of H1 adjacent to no vertex
of H2 . If v is any vertex of H2 , d(u, v) = 2 and there is a path (u, w, v).
Since w is adjacent to u, w is in H1 . Thus every vertex of H2 is adjacent to
a vertex of H1 .
Let n1 and n2 be the orders of H1 and H2 , and choose the numbering so
that n1 ≤ n2 . In the first case above k ≥ n1 and in the second k ≥ n2 , so in
any case k ≥ n1 . For a vertex u of H1 , let d1 (u) and d2 (u) be the numbers
of vertices of H1 and H2 adjacent to u. Then
so k = δ(G).
30
Proof. For independent vertices u and v of a graph G let sG (u, v) be the
minimum number of vertices that separate u and v (which exists because the
set of all other vertices separates), and let pG (u, v) be the maximum number
of internally disjoint u-v paths. If P is any such set of paths and W is a set
of vertices separating u and v, W must contain at least one internal vertex of
every path in P , so |P | ≤ |W | and hence pG (u, v) ≤ sG (u, v). If pG (u, v) = 0,
then u and v are in different components of G, and so sG (u, v) = 0. It
follows that if sG (u, v) = 1 then pG (u, v) = 1, so if pG (u, v) < sG (u, v) then
sG (u, v) ≥ 2. Suppose there exist independent vertices u and v of a graph G
with pG (u, v) < sG (u, v). Let t be the smallest integer for which there exist
such u, v and G with sG (u, v) = t, and amongst such u, v and G let G have
minimum size. We make the following claims about G.
(2) If w1 and w2 are adjacent vertices different from u and v, there is a set
W of t − 1 vertices such that W ∪ {w1 } and W ∪ {w2 } both separate
u and v.
31
By minimality of W , every vertex of W is in Ku , and since W has a vertex
not adjacent to u, Hu has at least t + 1 edges. Define Kv and Hv similarly.
Let Gu be the graph obtained from Hu by adding the vertex v and edges
vw for all w in W . Since Hv has at least t + 1 edges, Gu has smaller size
than G. Clearly W separates u from v in Gu ; let W ′ be any set of vertices
separating u from v in Gu . Every u-v path P in G gives a u-v path P ′
in Gu , namely the initial subpath in P from u to some w ∈ W , followed
by (w, v). Since P ′ contains a vertex of W ′ , so does P . It follows that
sGu (u, v) = t, and so pGu (u, v) = t. Now t internally disjoint u-v paths in
Gu give internally disjoint paths in Hu from u to each vertex of W . Similarly,
there are internally disjoint paths in Hv from each vertex of W to v. Putting
these together with the paths in Hu gives t internally disjoint u-v paths in
G, a contradiction.
At this point, (1)–(3) have been established, and we now derive a final
contradiction. Let (w0 , w1 , . . . , wk ) be a u-v geodesic in G. By (1), k ≥ 3.
By (2), there is a set W of t − 1 vertices such that W ∪ {w1 } and W ∪ {w2 }
both separate u and v in G. Since w1 is not adjacent to v, (3) implies
that every vertex of W is adjacent to v, and since w2 is not adjacent to u, it
implies that every vertex of W is adjacent to v. Since t ≥ 2, W is non-empty
and we have a contradiction to (1).
32
Proof. Suppose first that such paths exist. Take distinct vertices of G. Of
a set of k internally disjoint paths connecting them, at most one has no
internal vertices, so G has order at least k + 1. Thus it suffices to show that
any vertex-cut U has at least k elements. Let P be a set of k internally
disjoint paths joining vertices of different components of G − U . Then U
contains an internal vertex of every path in P , so indeed |U | ≥ k.
Suppose conversely that G is k-connected, and let u 6= v ∈ V (G). If
k = 1 the result is trivial, so suppose k ≥ 2. If u and v are not adjacent,
any set of vertices separating u and v has at least k elements, so Menger’s
Theorem gives at least k internally disjoint u-v paths. Otherwise, G − uv is
(k − 1)-connected by Lemma 6.7, so Menger’s Theorem gives at least k − 1
internally disjoint u-v paths in G − uv, and adding the path (u, v) gives at
least k internally disjoint u-v paths in G.
33
6.10 gives internally disjoint paths from u to every vertex of V . Replacing
these paths by their shortest initial segments ending on C, and V by the
set of terminal points of the new paths, we may assume that each path in
the set meets C only in its terminal point. Now there are vertices v1 and
v2 of V such that one of the v1 -v2 paths in C has no element of V or U as
an internal vertex. Replacing this path by a path through u gives a cycle
containing at least l + 1 vertices of U , completing the proof.
Exercises for §6
7 Trees
Recall that a spanning subgraph of a graph G is a subgraph containing all
the vertices of G.
34
Theorem 7.2. Any connected graph G has a spanning tree.
Theorem 7.3. For a graph G of order n and size m, any two of the following
conditions imply the third.
(1) G is connected.
(2) G is acyclic.
(3) m = n − 1.
Proof. (a) We show that (1) and (2) imply (3) by induction on n. If n = 1
then m = 0 and (3) holds. If n > 1 then as remarked above we may delete an
end-vertex, leaving a tree of order n − 1 and size m − 1, giving the inductive
step.
(b) Suppose that (2) and (3) hold and G has k components. Applying
part (a) to each component and adding, we have m = n − k, so k = 1, as
required.
(c) Suppose that (1) and (3) hold. Then G has a spanning tree T . By
part (a), T has size m, so T = G, giving (2).
Lemma 7.5. Let u and v be vertices of a graph G, and suppose there are
distinct u-v paths P and Q of lengths k and l. Then there is a cycle of length
at most k + l in G, and if P · Qr is not a cycle, there is one of length less
than k + l.
35
Qr , respectively. Then P1 · Qr1 is a cycle in G of length at most k + l, with
equality only if P = P1 and Q = Q1 .
Theorem 7.6. A graph G is a tree iff there is a unique u-v path for all u,
v ∈ V (G).
Proof. Suppose that any two vertices of G are connected by a unique path.
Certainly G is connected. If C is a cycle in G and u and v are any two
distinct vertices of C, there are two distinct u-v paths in C, a contradiction.
Thus G is a tree. The converse is immediate from the preceding lemma.
8 Counting trees
There is no known formula for the number of trees of order n (up to iso-
morphism). For small values of n one can enumerate the trees of order n by
considering all ways of adding a vertex and edge to a tree of order n − 1.
The unique trees of orders 1, 2 and 3 are K1 , K2 and P3 . For n = 4, there
are two trees, the path P4 and the star K1,3 . Of course, any path or star
is a tree. Other special classes of trees are the double stars, which have
36
Figure 7: The trees of order 5
exactly two (necessarily adjacent) vertices of degree greater than 1, and the
caterpillars, the trees for which deleting the end-vertices leaves a path (the
spine of the caterpillar). There are three trees of order 5, shown in Figure 7,
a path, a star, and a double star. There are 6 of order 6, whose enumeration
is left as an exercise.
For the number of trees on a given set of size n (up to equality), some-
times known as the number of labelled trees of order n, there is the following
theorem of Cayley [6]. The proof below is due to Prüfer [26].
Theorem 8.1. Let X = {1, . . . , n}. The number of trees with vertex set X
is nn−2 .
a vertex of Ti is an end-vertex
(8.1)
⇔ it is not in the set {si+1 , . . . , sn−2 }.
For i = n − 2 this is clear: Tn−2 has two vertices, which are end-vertices,
and the set in question is empty. Suppose that (8.1) holds for some i with
1 ≤ i ≤ n − 2. Ti−1 has one more vertex than Ti , namely vi , which is an
end-vertex of Ti−1 and is not in the set {si , . . . , sn−2 }. Further, a vertex
of Ti is an end-vertex of Ti−1 iff it is an end vertex of Ti and not equal to
si , which is the case iff it is not in {si , . . . , sn−2 }. This establishes (8.1) for
i − 1.
Now let (s1 , . . . , sn−2 ) be any sequence of elements of X. We construct
graphs Gi and sets Xi ⊆ X for 0 ≤ i ≤ n − 2 such that V (Gi ) = X,
|E(Gi )| = i, |Xi | = n − i, {si+1 , . . . , sn−2 } ⊆ Xi , and each component of
37
Gi has exactly one vertex in Xi . We take G0 to be the empty graph on
X and X0 = X. Suppose that 1 ≤ i ≤ n − 2 and Gi−1 and Xi−1 have
been constructed. Then at least two elements of Xi−1 are not in the set
{si , . . . , sn−2 }; let vi be the least such element. Then vi and si are in distinct
components of Gi−1 . Setting Gi = Gi−1 + vi si and Xi = Xi−1 − {vi }, Gi
and Xi have the desired properties. The set Xn−2 has two elements vn−1
and vn , and the graph Gn−2 has n − 2 edges and two components, one
containing vn−1 and the other containing vn . Hence T = Gn−2 + vn−1 vn
is a tree. For 0 ≤ i ≤ n − 2, let Ti be the subgraph of T induced by the
vertices Xi . We show that Ti is a tree whose end vertices are the elements
of Xi − {si+1 , . . . , sn−2 }. For i = n − 2, Xn−2 = {vn−1 , vn } and vn−1 vn is
an edge of T , so the result holds in this case. Suppose it holds for some i,
1 ≤ i ≤ n−2. Then Xi−1 = Xi ∪{vi }, si ∈ Xi , and vi si is an edge of Gi , and
hence of T . Thus Ti−1 = Ti + vi si , which is a tree, and its set of end-vertices
is obtained from that of Ti by deleting si (if present) and adding vi . This
proves the result for i − 1. Note also that vi is the least end vertex of Ti−1 .
It follows that the trees Ti are also the trees used in constructing the Prüfer
sequence of T , and that this sequence is equal to (s1 , . . . , sn−2 ).
Now, starting with a tree T , construct the trees T0 , . . . , Tn−2 and the
Prüfer sequence (s1 , . . . , sn−2 ) as in the first paragraph, and then the graphs
G0 , . . . , Gn−2 and the sets X0 , . . . , Xn−2 from this sequence. We show that,
for 0 ≤ i ≤ n − 2,
38
Theorem requires that any cofactor of L(Kn ) is equal to nn−2 ; this is left as
an exercise.
Proof. The sum of each row or column of L(G) is zero, from which it follows
easily that all cofactors of L(G) are equal. For any non-empty proper subset
U of V = V (G), let LU (G) be the matrix obtained from L(G) by deleting
the rows and columns corresponding to elements of U . Define a U -forest in
G to be a spanning forest in which every component contains exactly one
element of U . When U has a single element, a U -forest is just a spanning
tree, so the result will follow if we show that the number of U -forests is
equal to det LU (G). This we do by induction on |V − U |. Suppose this is
1, and let v be the element of V − U . Then det LU (G) = deg v, while the
U -forests are the spanning subgraphs with just one edge, incident to v. The
result holds in this case. Now suppose that |V − U | > 1, and the result holds
for larger subsets of V . Here we use induction on the number k of edges
with one vertex in U and one in V − U . If k = 0, there are no U -forests,
while LU (G) is just the Lagrangian of the subgraph induced by V − U , and
so has determinant zero. Suppose then that k > 0, and let e = uv be an
edge with u ∈ U and v ∈ V − U . Set H = G − e. The matrix LU (G)
is obtained from LU (H) by adding 1 to the diagonal entry corresponding
to v, so det LU (G) = det LU (H) + det LU ′ (H), where U ′ = U ∪ {v}. By
the induction on k, det LU (H) is the number of U -forests in H, and by the
induction on |V − U |, det LU ′ (H) is the number of U ′ -forests in H. The
U -forests in H are precisely the U -forests in G that do not contain e. Also,
there is a one-to-one correspondence between the U ′ -forests in H and the
U -forests in G that do contain e, given by adding e to a U ′ -forest in H. The
result follows.
Exercises for §8
8.1. Draw all trees of order 6.
8.2. Let A be a square matrix in which every row and column sums to zero.
Show that all cofactors of A are equal.
8.3. Show that any cofactor of L(Kn ) is equal to nn−2 .
9 Permutation groups
In what follows, we need some knowledge of groups. However, the general
theory is not really needed, and a few easy facts about permutation groups
will suffice.
39
Definition 9.1. Let X be a finite set. A permutation of X is a bijection
π : X → X. The set of all permutations of X will be denoted by S(X).
When X = {1, . . . , n}, we write Sn for S(X), and we will often identify
S(X) with Sn when X has n elements.
For any set X we have the identity permutation id = idX ∈ S(X), and
for any π ∈ S(X) the inverse π −1 is in S(X). Finally, for any π and σ in
S(X), the composite πσ is in S(X). We have:
(1) id ∈ A;
40
Example 9.5. Let G be a graph. An automorphism of G is an isomorphism
α : G → G. The set Aut(G) of all automorphisms of G is a permutation
group on V (G).
Proof. That any subgroup has this property follows from (2) and (3) in
Definition 9.4. Suppose, conversely, that A is a non-empty subset of S(X)
with the given property. Take some α ∈ A; then id = αα−1 ∈ A. Hence,
for α ∈ A, α−1 = id α−1 ∈ A. Finally, for α and β ∈ A, αβ = α(β −1 )−1 ∈
A.
41
Figure 8: Two pictures of M4
Example 10.2. Aut Kn = Aut K̄n ∼ = Sn , of order n!. For any other graph
G of order n, | Aut G| is less than n!, and divides n! by Theorem 9.8.
42
v1
1
3 4
v6
v5 5 2 v2
1
v10 v7
5 2
4 3
2 5
v9 v8
4 3
1
v4 v3
rungs, namely the rim, of length 2n ≥ 8, and any cycle containing a single
rung has length n + 1 ≥ 5. Therefore a 4-cycle must contain at least two
rungs, and hence consists of a pair of opposite edges of the rim together with
the two adjacent rungs. This shows that each edge of the rim lies on just
one 4-cycle, while each rung lies on two. It follows that any automorphism
of Mn takes rim edges to rim edges (and rungs to rungs), and therefore lies
in Aut C2n .
43
v3
v1
v2
Proof. Let a, b and c be distinct positive integers. Take the disjoint union
H of Pa , Pb and Pc , and let v1 , v2 and v3 be endpoints of the three compo-
nents. Form G by adding edges v1 v2 , v1 v3 and v2 v3 (see Figure 10). Any
automorphism α of G must take the unique cycle induced by {v1 , v2 , v3 } to
itself, and so induce an automorphism of H. Since the components of H
have different orders, each must be taken to itself, and it follows that α = id.
Thus G is an asymmetric graph of order a + b + c, which can be any integer
n ≥ 6.
It remains to show that there is no asymmetric graph of order n for
2 ≤ n ≤ 5. A reader who has done Exercise 1.1 can simply examine all
the graphs of these orders. To avoid examining all of them, we can note
the following. Since Aut(G) = Aut(Ḡ) it is enough to consider graphs of
size m ≤ 21 n2 . Any disconnected graph is either empty or has a non-trivial
component of smaller order, so if the result has already been verified for
smaller values of n it is enough to consider connected graphs. There is no
graph satisfying these conditions of order 2 or 3, and for order 4 such a
graph must be a tree, and therefore either P4 or K1,3 , neither of which is
asymmetric. For order 5, such a graph is either a tree or has size 5. The
44
three trees of order 5 are shown in Figure 7, and are not asymmetric. A
connected graph of order 5 and size 5 contains a unique cycle, of length 3, 4
or 5, so there are five such graphs; it is left to the reader to verify that they
are not asymmetric.
A reader who has done Exercise 8.1 can verify that there are no asym-
metric trees of order 6.
Exercises for §10
11 Eulerian graphs
The first paper in graph theory was Euler’s 1736 work on the problem of the
Königsberg bridges, in which he studied what are now known as Eulerian
circuits and trails. An English translation is in Biggs, Lloyd and Wilson
[4], which contains much historical detail, as well as all or part of several
important papers, and references to others.
45
Remark 11.2. We make repeated use of the following obvious facts. If C
is a circuit in a graph G, every vertex of G is incident with an even number
of edges of C, while if T is a u-v trail with u 6= v then u and v are incident
with an odd number of edges of T , and every other vertex is incident with
an even number.
Proof. The “only if” part follows from the preceding remark. Suppose that
every vertex of G is even, and let T be a trail of maximum length. Suppose
T is a u-v trail. If u 6= v, v is incident with an odd number of edges of T ,
and since deg v is even there is an edge vw not on T . But then T · (v, w) is a
longer trail, which is impossible. Thus T is a circuit. If T is not an Eulerian
circuit, there is an edge xy that is not on T but has at least the vertex x on
T . There is an x-x trail T ′ with the same edges as T , and now T ′ · (x, y) is
a longer trail than T . Thus T is indeed an Eulerian circuit.
Theorem 11.4 (Euler). A connected graph G has an Eulerian trail iff there
are exactly two odd vertices in G.
Proof. The “only if” part again follows from Remark 11.2, so suppose G has
just two odd vertices u and v. Form a graph G′ by adding to G a vertex w
and edges uw and vw. By the previous theorem, G′ is Eulerian. We may
take an Eulerian circuit in G′ of the form T · (v, w, u), and then T is an
Eulerian trail in G.
When there are more than two odd vertices, something can still be said.
46
of elements X of P with |X| odd, and suppose for a contradiction that
ν(P) > 1 for any trail partition. We associate a graph H(P) to each trail
partition. The vertices are the elements of P, and X 6= Y ∈ P are adjacent
iff they contain adjacent edges. We show that H(P) is connected. Let X
and Y ∈ P, and let u and v be vertices of edges of X and Y respectively.
Let u = u0 , u1 , . . . , ur = v be a path in G. For 1 ≤ i ≤ r, the edge ui−1 ui
of G belongs to some Zi ∈ P. In the sequence X, Z1 , . . . , Zr , Y of vertices
of H(P), any two consecutive terms are equal or adjacent, so X and Y are
connected in H(P). Hence we may define µ(P) to be the minimum distance
in H(P) between distinct elements X and Y of P with |X| and |Y | odd.
Let P be a trail partition such that for every other trail partition P′ ,
either ν(P) < ν(P′ ), or ν(P) = ν(P′ ) and µ(P) ≤ µ(P′ ). We consider two
cases, depending on whether or not µ(P) = 1. If µ(P) = 1, there are adjacent
vertices X and Y of H(P) with |X| and |Y | odd. These are the edge sets
of trails u0 , u1 , . . . , ur and v0 , v1 , . . . vs in G, where r and s are odd, and
ui = vj for some i and j. By reversing one trail if necessary, we may assume
that i 6≡ j (mod 2). We may form a new trail partition P′ by replacing X
and Y by the edge sets X ′ and Y ′ of the trails u0 , u1 , . . . , ui , vj+1 , . . . , vs and
v0 , v1 , . . . , vj , ui+1 , . . . , ur . Since |X ′ | and |Y ′ | are even, ν(P′ ) = ν(P) − 2, a
contradiction.
Now suppose that µ(P) > 1. Then there are vertices X, Y , Z and W
of H(P) with |X| and |W | odd, |Y | even, Y adjacent to both X and Z,
and dH(P) (Z, W ) = µ(P) − 2. Then X and Y are the edge sets of trails
u0 , u1 , . . . , ur and v0 , v1 , . . . vs in G, where r is odd, s is even, ui = vj for
some i and j, and some vk is a vertex of an edge of Z. Note that j 6= k, since
otherwise X and Z would be equal or adjacent. By reversing the trails if
necessary, we may assume that j < k, and then that i 6≡ j (mod 2). We may
form a new trail partition P′ by replacing X and Y by the edge sets X ′ and Y ′
of the trails u0 , u1 , . . . , ui , vj+1 , . . . , vs and v0 , v1 , . . . , vj , ui+1 , . . . , ur . Since
|X ′ | is odd and |Y ′ | is even, ν(P′ ) = ν(P). Since X ′ and Z have adjacent
edges, µ(P′ ) ≤ dH(P′ ) (X ′ , W ) ≤ µ(P) − 1. This contradiction completes the
proof.
Proof. Suppose first that G is Eulerian, so that every vertex is even, and let
e = uv be any edge. Let T be the set of all trails in G − e that start at u
47
and either do not contain v, or contain it only as the terminal vertex. For
k ≥ 0, let Tk be the set of trails in T of length at most k that either have v
as terminal vertex or are of length exactly k. For k = 0, T0 contains only the
trivial u-u path, while for sufficiently large k, Tk is the set T ′ of all trails in T
that end at v. Suppose that T ∈ Tk is a u-x trail where x 6= v. If x 6= u then
x has even degree in G − e and is incident with an odd number of edges of T ,
while if x = u then x has odd degree in G − e and is incident with an even
number of edges of T . In either case there are an odd number of edges xy of
G − e that are not on T , and for each one T · (x, y) is in Tk+1 . Every element
of Tk+1 is either of this form, or is an element of Tk ending at v. It follows
by induction that |Tk | is odd for all k, so |T ′ | is odd. Now the set P of all u-v
paths in G − e is a subset of T ′ , and is in one-to-one correspondence with
the set of all cycles in G containing e. Suppose T = (u = u0 , u1 , . . . , uk = v)
is an element of T ′ − P. Let j be the smallest integer for which there exists
i < j with ui = uj . Then
α(T ) = (u0 , u1 , . . . , ui , uj−1 , . . . , ui+1 , uj , uj+1 , . . . , uk )
is a different element of T ′ − P, and α2 (T ) = T , so |T ′ − P| is even and |P|
is odd, as required.
Now suppose that every edge of G lies on an odd number of cycles. Let
u be a vertex of G, and consider the set P of all pairs (e, C), where e is an
edge incident with u and C is a cycle containing e. Because there are an odd
number of such pairs for each edge incident with u, |P | ≡ deg u (mod 2).
Because there are two such pairs for each cycle containing u, |P | is even.
Hence deg u is even, and G is Eulerian by Theorem 11.3.
12 Hamiltonian graphs
Definition 12.1. A Hamiltonian cycle in a graph G is a cycle containing
every vertex of G, and G is Hamiltonian if it contains a Hamiltonian cycle.
A Hamiltonian graph has order at least 3 and is connected, and indeed,
by Theorem 5.9, 2-connected. Not every 2-connected graph is Hamiltonian,
however; the Petersen graph, for instance, is not Hamiltonian. The graph
of the dodecahedron, shown in Figure 11, is Hamiltonian; the association of
Hamilton’s name with this idea comes from his invention of a game based
on this fact; see [4].
Theorem 12.2. Let G be a graph of order n and u and v independent
vertices of G with deg u + deg v ≥ n. Then G is Hamiltonian iff G + uv is
Hamiltonian.
48
Figure 11: The graph of the dodecahedron
(u = u1 , u2 , . . . , un = v, u).
is a Hamiltonian cycle in G.
49
We may therefore speak of the closure of G, denoted C(G). The next
three results are immediate consequences of Theorem 12.2.
Proof. Let H = C(G). Note that G, and hence H, has no isolated vertices.
We claim that H ∼ = Kn . Suppose not, and let u and v be independent
vertices of H such that degH u + degH v is a maximum. By the definition of
closure, degH u+degH v ≤ n−1. Suppose that degH u ≥ degH v and set k =
degH v. Then 1 ≤ k ≤ (n−1)/2 < n/2. There are n−1−degH u ≥ k vertices
of H independent of u, and for each such vertex w, degG w ≤ degH w ≤
degH v = k. That is, dk ≤ k. There are n − 1 − degH v = n − 1 − k vertices
independent of v in H, and for each such vertex w, degG w ≤ degH w ≤
degH u ≤ n − 1 − k. Since also degG v ≤ k ≤ n − 1 − k, there are at least
n − k vertices of degree at most n − k − 1 in G, and dn−k ≤ n − k − 1.
50
of Qi that is on C and Pi the u-vi subpath of Qi . If C contains an edge
vi vj , we may replace it by Pir · Pj to get a longer cycle, which is impossible.
Let wi be the vertex following vi in one cyclic ordering of C. If some wi is
adjacent to u, we may replace the edge vi wi by Pir · (u, wi ) to get a longer
cycle, which is again impossible. Now |{u, w1 , . . . , wk }| = k + 1 > β(G), so
this set is not independent and G contains an edge wi wj (which is of course
not in C). Replacing the edges vi wi and vj wj by Pir · Pj and (wi , wj ) gives
a longer cycle, and this contradiction completes the proof.
12.2. Let G be a graph of order n ≥ 3 such that deg u + deg v ≥ n for any
independent vertices u and v of G. Let (d1 , . . . , dn ) be the degree
sequence of G with d1 ≤ · · · ≤ dn . Prove that there is no k with
1 ≤ k < n/2 such that dk ≤ k and dn−k ≤ n − k − 1.
12.3. Show that the complete k-partite graph Kr1 ,...,rk , where r1 ≤ · · · ≤ rk ,
Pk−1
is Hamiltonian iff rk ≤ i=1 ri .
51
(a) (b)
14 Planar graphs
Definition 14.1. A plane embedding of a graph G consists of the following.
First, a bijection α from V (G) to a subset of R2 . Second, a bijection β
from E(G) to a set of arcs in R2 . These are required to satisfy the following
conditions. First, if e is the edge uv, then the endpoints of β(e) are α(u)
and α(v). Second, if e and f are distinct edges, then β(e) and β(f ) intersect
only if e and f are adjacent, and then only in α(u) where u is the common
vertex of e and f .
Less formally, a plane embedding of G is a picture of G in which no
edge crosses itself or any other edge. A graph is planar if it has a plane
embedding. A plane graph is a graph G together with a plane embedding
of G. Given a plane graph we usually identify each vertex v with the point
α(v) in R2 , each edge e with the arc β(e), and the graph G with the union
of these subsets of R2 . Figure 12 shows two pictures of K2,3 . That in (a)
is not a plane embedding, while the one in (b) is, and shows that K2,3 is
planar.
52
If G is a plane graph, the path-components of the open set R2 − G
are called the regions of G. There is one unbounded region, the exterior
region. The boundary of a region R is a subgraph of G denoted by ∂R;
we say that a region is adjacent to the vertices and edges in its boundary.
A plane graph G may be re-embedded so that any given vertex or edge is
adjacent to the exterior region as follows. Regard R2 as the plane of the
first two coordinates in R3 , let S 2 be the unit sphere centered at the origin
in R3 , and let N = (0, 0, 1), the “north pole” of S 2 . There is a bijection
φ : S 2 − {N } → R2 , called stereographic projection. For x ∈ S 2 − {N }, φ(x)
is the point in which the line through N and x meets R2 . Let x be a point
of some region adjacent to the given vertex or edge, and take the image of
G under the composite of φ−1 , followed by a rotation of S 2 taking φ−1 (x)
to N , followed by φ.
Clearly a graph is planar iff each component is planar. Let G be a non-
trivial, connected plane graph. A arrow of a graph is an ordered pair (u, v)
of adjacent vertices (i.e., it is a path of length 1). For each arrow (u, v) of G,
there is a region R+ to its left and a region R− to its right. If uv is a bridge
then R+ = R− ; otherwise R+ 6= R− . A side of a region R is an arrow
having R to its left. The number of sides of R is thus twice the number
of bridges plus the number of other edges of ∂R. We may assume that at
each vertex v there is a circle centered at v meeting each edge incident to
v just once. Traversing this circle either clockwise or counterclockwise gives
a cyclic ordering of the vertices adjacent to v. For any region R of G we
construct a closed walk in its boundary as follows. Start with a side (u0 , u1 )
of R. Suppose we have constructed (u0 , u1 , . . . , uk ) so that (ui , ui+1 ) is a
side of R for 0 ≤ i < k. Let v be the vertex adjacent to uk that follows
uk−1 in the clockwise ordering, or uk−1 if uk has degree 1. Then (uk , v) is
a side of R. If uk = u0 and v = u1 , stop; otherwise extend the walk by
taking uk+1 = v. The result is a walk containing each side of R just once.
If there are no bridges in ∂R the walk is a circuit, and if there are also no
cut-vertices it is a cycle. In particular, if G is 2-connected, the boundary of
any region is a cycle.
Theorem 14.2 (Euler’s formula). Let G be a connected plane graph of order
n and size m having r regions. Then n − m + r = 2.
Proof. The proof is by induction on m. If G is a tree then m = n − 1, and
since every edge is a bridge, r = 1, so the result holds. Otherwise, let e be
an edge that lies on a cycle. Then G − e is a connected plane graph of order
n and size m − 1 having r − 1 regions (the two regions of G adjacent to e
lying in a single region of G − e). The result follows.
53
Corollary 14.3. Let G be a plane graph of order n and size m having r
regions and k components. Then n − m + r = k + 1.
Proof. The second statement implies the first, because it gives 2m = 3r,
where r is the number of regions, and together with Euler’s formula this
gives m = 3n − 6. Clearly G is connected; in fact, it is 2-connected, because
if u were a cut vertex, there would be vertices v and w adjacent to u, with
w following v in the counterclockwise order of the neighbors of u, and in
different components of G − u. But then we could embed G + vw by putting
the edge vw in the region to the left of (u, v). Thus the boundary of every
region is a cycle. Suppose that the boundary of some region R has length
greater than 3, and let u1 , u2 , u3 and u4 be consecutive vertices around ∂R.
Not both of u1 u3 and u2 u4 can be edges of G, and for one that is not we can
embed a corresponding arc in R to contradict maximal planarity of G.
54
Proof. One direction is trivial: every subgraph of a planar graph is planar.
Suppose that G is a graph all of whose blocks are planar. We may assume
that G is connected. The proof is by induction on the number k of blocks,
the case k = 1 being obvious. Suppose that k > 1. By Theorem 5.18, G has
an end-block B. Let v be the cut-vertex in B, and H the union of all other
blocks of G, so that G = B ∪ H and B and H have only v in common. By
induction, H is planar. We may embed both B and H in R2 so that v is
adjacent to the exterior region. Further, we may choose these embeddings so
that v = (0, 0) in both cases, every other point (x1 , x2 ) of B has x1 < 0, and
every other point (x1 , x2 ) of H has x1 > 0. This gives a plane embedding of
G.
Definition 14.9. The girth of a graph that contains a cycle is the minimum
length of a cycle.
Theorem 14.10. Let G be a 2-connected planar graph of order n, size m
g
and girth g. Then m ≤ g−2 (n − 2).
Proof. Embed G in the plane and let the number of regions be r. Since G is
2-connected, the boundary of any region is a cycle, so it has at least g sides.
Thus 2m ≥ gr, so 2g = gn − gm + gr ≤ gn − (g − 2)m.
g
Example 14.11. For K3,3 , m = 9, n = 6, g = 4 and g−2 (n − 2) = 8, so
K3,3 is non-planar.
Example 14.12. For the Petersen graph, m = 15, n = 10, g = 5 and
g 40
g−2 (n − 2) = 3 , so it is non-planar.
Theorem 14.13. Let G be a connected plane graph with δ(G) ≥ 3. Then
there is a region with at most 5 sides.
Proof. Let the order, size and number of regions be n, m and r. We have
2m ≥ 3n and if every region has at least 6 sides 2m ≥ 6r, or m ≥ 3r. Hence
6 = 3n − 3m + 3r ≤ 0, a contradiction.
Theorem 14.14. Let G be a connected plane graph with δ(G) ≥ 3. Let nk
be the number of vertices of degree k and rk the number of regions with k
sides. Then n3 + r3 ≥ 8.
Proof. Let the order, size and number of regions
P be n, m and
P r. Note that
nk = rk = 0 for k <
P P3. We have n = k≥3 nk , r = k≥3 rk , 2m =
k≥3 kn k , and 2m = k≥3 krk . Hence
X
8 = (4n − 2m) + (4r − 2m) = (4 − k)(nk + rk ) ≤ n3 + r3 .
k≥3
55
The following theorem is due independently to Fáry [13] and Wagner
[30]. The proof below is based on Wood [32].
Theorem 14.15. Any planar graph has a plane embedding in which each
edge is a straight line segment.
Proof. It is enough to prove this for a maximal planar graph G, which we
do by induction on the order n. In fact we prove that for any embedding of
G there is a line-segment embedding in which the counterclockwise order on
the neighbors of any vertex is unchanged. When n = 3, G is isomorphic to
K3 and the result is clear, so suppose n ≥ 4. Recall that all regions of the
embedding are bounded by triangles. If u and v are adjacent vertices, the
edge uv is in the boundary of two regions, whose other vertices are common
neighbors of u and v. We show that u and v may be chosen so that these are
their only common neighbors. If there are three common neighbors, uv is
in three triangles of G, at least one of which has a vertex in its interior and
one in its exterior. Amongst all triangles of G which have a vertex in both
their interior and their exterior, choose an innermost one, choose a vertex
u in its interior, and let v be any neighbor of u. Then indeed u and v have
just two common neighbors p and q.
With a suitable choice of notation, the neighbors of u in counterclockwise
order are v, p, x1 , . . . , xr , q, and those of v are u, q, y1 , . . . , ys , p. Form a graph
G′ by deleting the vertex u and adding edges vxi for 1 ≤ i ≤ r. (This oper-
ation is called contracting the edge uv, and will appear in the next section.)
We may extend the embedding of G − u to an embedding of G′ , in which
the neighbbors of v in counterclockwise order are p, x1 , . . . , xr , q, y1 , . . . , ys .
By inductive hypothesis we may re-embed G′ so the edges are line seg-
ments, preserving orderings of neighbors. Suppose the angle θ between
vp and vq, measured counterclockwise, is at most π. Consider the cycle
C = (v, p, x1 , . . . , xr , q, v). Its interior contains the edges vxi , except for
their endpoints. Delete these edges, take u to be in the interior of C, and
add the line segments from u to v, p, q and each xi . If u is taken sufficiently
close to v, these segments are in the interior of C, except for their endpoints
on C, and we have the desired embedding of G.
If θ > π, rename v as u, and similarly situate v in the interior of
(u, q, y1 , . . . , ys , p, u).
15 Kuratowski’s Theorem
Definition 15.1. An elementary subdivision of a graph G is the operation
of deleting an edge uv and adding a new vertex w and edges uw and vw.
56
A subdivision of G is a graph isomorphic to one obtained by zero or more
elementary subdivisions.
Example 15.2. The Petersen graph has a subgraph that is a subdivision
of K3,3 , obtained by deleting two edges with the same label in Figure 9.
Clearly a sudivision of G is planar iff G is planar, and as already re-
marked, a subgraph of a planar graph is planar. Together with Examples
14.5 and 14.11, these observations prove the “only if” part of the following
theorem.
Theorem 15.3 (Kuratowski [23]). A graph G is planar iff no subgraph of
G is a subdivision of K5 or K3,3 .
Lemma 15.4. Let G be a 2-connected graph of order n ≥ 4. Then G has
either a vertex v such that G − v is 2-connected, or a vertex of degree 2.
Proof. Suppose that G has no vertex v such that G − v is 2-connected.
Then, for evey vertex x of G there is a vertex y 6= x such that G − {x, y}
is disconnected. Let u and v be such a pair of vertices for which some
component of G − {u, v}, say G1 , has minimum order. Note that each of u
and v is adjacent to some vertex in every component of G − {u, v}. Let G2
be the union of the components of G − {u, v} other than G1 , and for i = 1
or 2, let Hi = hV (Gi ) ∪ {u, v}i. Let w1 be a vertex of G1 . Then there is a
vertex w2 6= w1 such that G − {w1 , w2 } is disconnected. If w2 is in G1 , then
because H2 is connected, some component of G − {w1 , w2 } is contained in
G1 − {w1 , w2 }, a contradiction. If w2 = u, then because hV (G2 ) ∪ {v}i is
connected, some component of G − {w1 , w2 } is contained in G1 − w1 , again
a contradiction. A similar contradiction arises if w2 = v, so w2 is in G2 .
For i = 1 or 2, every vertex of Hi − wi is connected to either u or v in
Hi − wi , because it is connected to, say, u in G − wi and a path can only
leave Hi at u or v. It follows that Hi − wi has exactly two components, one
containing u and the other containing v. Let the components of H1 − w1 be
Ku and Kv . If Ku is non-trivial, then Ku − u is a union of components of
G − {w1 , u}, contradicting the minimality of the order of G1 . Thus Ku is
trivial, and similarly Kv is trivial. This means that G1 has only the single
vertex w1 , which is therefore adjacent to only u and v, and of degree 2.
57
2-connected, we are done. Otherwise there is a vertex u such that G − e − u
is disconnected; u is not a vertex of e since otherwise G − e − u = G − u.
Now e is a bridge of G − u but its vertex v is not a cut-vertex of G − u.
Hence v has degree 1 in G − u, and therefore degree 2 in G.
Proof. The sets (u, v) and (v, u) are non-empty; let w ∈ (u, v) and w′ ∈ (v, u)
be adjacent to u. We first define a w-w′ walk with all edges (and hence
vertices) on or interior to Γ. Set w0 = w. Let R1 be the interior region
58
of G adjacent to uw0 , and take the path going counterclockwise around
the boundary of R1 from w0 to its other vertex w1 adjacent to u. If uw1
is interior, let R2 be the other (necessarily interior) region of G adjacent
to uw1 , and take the path going counterclockwise around the boundary of
R2 from w1 to its other vertex w2 adjacent to u. Continue in this fashion,
stopping when uwk is on Γ, and therefore wk = w′ . Stringing these paths
together gives the desired walk W . Note that neither u nor v is on W .
Let W be w = x0 , x1 , . . . , xr = w′ . There exist i and j with 0 ≤ i < j ≤ r
such that xi is on (u, v), xj is on (v, u), and xk is interior to Γ for i < k < j.
An xi -xj path contained in the walk xi , xi+1 , . . . , xj is the required chord.
Lemma 15.7. Suppose that G is a non-planar graph that has an edge e such
that G − e is 2-connected and planar. Then G has a forbidden subgraph.
Proof. Let e = u0 u2 . (The reason for the peculiar indexing will emerge
later.) Set H = G − e. Since H is 2-connected, it has a cycle Γ containing
u0 and u2 . The pieces of G relative to Γ are the pieces of H together with
the chord C0 induced by e. For any plane embedding of H, there must be
an interior piece of H crossing C0 , since otherwise Lemma 15.6 would give
an interior region adjacent to u0 and u2 , and hence a plane embedding of
G.
Now fix some embedding of H, and choose Γ to have the maximum
number of interior regions amongst all cycles of H containing u0 and u2 .
Let P be an exterior piece of H. Since H is 2-connected, P has at least two
vertices on Γ. Suppose it has two such v and w for which (v, w) contains
neither u0 nor u2 . Then the union of the path [w, v] and a v-w chord in P is
a cycle containing u0 and u2 with more interior regions than Γ, contrary to
hypothesis. It follows that P has exactly two vertices on Γ, with one, vP , on
(u0 , u2 ), and the other, wP , on (u2 , u0 ). Amongst the regions of the subgraph
Γ ∪ P of H, there are three whose boundaries contain edges of Γ. One is the
interior of Γ. Of the others, one, RP− , has boundary meeting Γ in [wP , vP ],
and the otherRP+ , has boundary meeting Γ in [vP , wP ]. The boundary of
any other region has at most the vertices vP and wP in common with Γ, and
cannot have both. (Suppose it does. Then its boundary is the union of two
vP -wP chords of Γ, both of which must contain exterior vertices. However,
any path in Γ ∪ P from an exterior vertex of one chord to an exterior vertex
of the other must pass through vP or wP , contradicting the fact that the
59
w1 = w2 R1
R0 u0 u2 R2
v1 v2
60
u0 u3 u0 u3
x3
x1 x2 x1
x4
x2
u1 u2 u1 u2
(a) (b)
u0 u3 u0 u3
x1
x1
x2
u1 u2 u1 u2
(c) (d)
61
u
v w
*
v w G
j
H
u
v w
62
Definition 15.8. A minor of a graph G is a graph isomorphic to one ob-
tained from G by zero or more applications of the operations of deleting
or contracting an edge, or deleting a vertex. A minor obtained using only
contractions is a contraction of G.
16 Crossing number
In this section we need to be more precise about what we mean by a drawing
of a graph.
63
conditions. First, if e is the edge uv, then the endpoints of β(e) are α(u)
and α(v). Second, if e and f are adjacent edges, β(e) and β(f ) intersect
only in α(u) where u is the common vertex of e and f . Thirdly, if e and f
are distinct non-adjacent edges then β(e) and β(f ) intersect in at most one
point, at which β(e) and β(f ) cross, and which lies on the image of no other
edge. The crossing number of G is the minimum number ν(G) of crossings
in any drawing of G.
64
for 0 ≤ i < k, and the segments to k 2 arcs which are the edges of a drawing
of Kk,k . To obtain a drawing of Kn we first add edges between distinct ui
and uj that are straight line segments. These do not cross any edges of the
first type, and for any four of the ui there is just one crossing on the edges
they induce. Thus the number of crossings on edges of this type is k4 . We
may similarly add edges between distinct vi and vj lying in the set of points
at distance at least 2 from the origin, having a further k4 crossings. We
now count, for a vertex w, the number of crossings on edges of the first type
incident to w. This is independent of w, so we take w = u0 . For given
a and b with 0 ≤ a, b < k, we determine the number of crossings of u0 va
with edges ui vi+b . If a = b, there are none. If a 6= b the corresponding line
segments cross if either i < 0 and i + b > a, or i > 0 and i + b < a; that is,
if a − b < i < 0 or 0 < i < a − b. At most one of these ranges is non-empty,
and different values of i in the range give different edges, so the number of
crossing edges is |a − b| − 1. Hence the number of crossings on edges of the
first type incident with w is
k X
X a−2 k
X
a−1 k
2 (a − b − 1) = 2 2 =2 3 .
a=3 b=1 a=3
There are 2k vertices, and each crossing is on edges incident with four
vertices, so the total number of crossings on edges of the first type is
2k · 2 k3 /4 = k k3 , and the number of all crossings in the diagram is
k
k
k(k−1)2 (k−2) 1
n n−1 n−2 n−3
k 3 +2 4 = 4 = 4 2 2 2 2 .
crossings.
65
Figure 16: C4 × C5 with 10 crossings
It is conjectured that both bounds in this Theorem are exact, and this
is known for r ≤ 6 and any s, and r = 7 and s ≤ 10.
Figure 16 shows a straight-line drawing of C4 × C5 with (4 − 2) · 5 = 10
crossings, from which the general case of Cr × Cs given in the next result
should be clear.
Theorem 16.5. ν(Cr × Cs ) ≤ ν̄(Cr × Cs ) ≤ (r − 2)s.
It is conjectured that, for r ≤ s, both these bounds are exact, and this
is known for r = 3 and 4. We shall prove it for r = 3.
Lemma 16.6. ν(C3 × C3 ) = 3.
66
Proof. What remains to be proved is that any drawing of G = C3 × C3 has
at least three crossings. Consider a drawing with c crossings. Certainly G
is not planar (see below), so c > 0. Pick an edge containing a crossing, and
then a 4-cycle containing this edge. Deleting the edges of this cycle leaves
a subdivision of K5 , so there is at least one crossing among the remaining
edges. If there are two or more then c ≥ 3, so suppose there is just one.
This means that our original drawing can be obtained from a one-crossing
drawing of K5 by picking points subdividing the edges of a 4-cycle in K5 ,
and adding an edge between the subdivision points on each adjacent pair of
edges. There is an essentially unique one-crossing diagram of K5 ; replacing
the crossing by a vertex gives the plane embedding of the graph of the
octahedron. We must consider all ways of choosing a 4-cycle in the drawing
of K5 , and for any edge on the cycle through the crossing, choosing which
side of the crossing to put the subdivision point.
We consider the drawing as lying on the unit sphere in R3 , with vertices
u± = (±1, 0, 0), v± = (0, ±1, 0) and w = (0, 0, 1), and with edges lying in
the coordinate planes. The edges u+ u− and v+ v− are semicircles crossing
at (0, 0, −1), and the other edges are quarter-circles. Rotation through a
multiple of π2 about the z-axis induces an automorphism of the K5 , as does
reflection in any of the vertical planes x = 0, y = 0, x = y and x = −y.
This means we may assume that our 4-cycle is one of the four below. In
each case, we denote the subdivision points on the edges by a1 , a2 , a3 and
a4 in the order in which the edges appear in the cycle as written.
(a) (u+ , v+ , u− , v− , u+ ). In this case every added edge must have a cross-
ing, so c ≥ 5.
(b) (u+ , v+ , v− , u− , u+ ). In this case we may further assume that a4 has
positive x-coordinate and that a2 has positive y-coordinate. Then each edge
added at a3 must have a crossing, so c ≥ 3.
(c) (u+ , v+ , u− , w, u+ ). Here each of a1 a2 and a3 a4 must have a crossing,
so c ≥ 3.
(d) (u+ , u− , v+ , w, u+ ). The edge a1 a4 must have one crossing if a1 has
positive x-coordinate, and two if it has negative x-coordinate, and if a1
has positive x-coordinate the edge a1 a2 must also have a crossing, so again
c ≥ 3.
67
hui , vi , wi i and the s-cycles are induced by the sets U = {ui | i ∈ Zs },
V = {vi | i ∈ Zs } and W = {wi | i ∈ Zs }, with edges ui ui+1 , vi vi+1 and
wi wi+1 . We must prove that, given a drawing of G with c crossings, c ≥ s.
If any edge of some Ti contains a crossing, deleting the edges of Ti gives a
drawing of a subdivision of C3 × Cs−1 with at most c − 1 crossings, so by
inductive hypothesis c − 1 ≥ s − 1.
Suppose then that there are no crossings on the edges P of any Ti . Let ce
be the number of crossings on the edge e and note that e∈E(G) ce = 2e.
P i ∈ Zs , let Xi be the set of edges {ui ui+1 , vi vi+1 , wi wi+1 }. We show that
For
e∈Xi ce ≥ 2, which will complete the proof. If two edges of Xi cross each
other this is clear, so suppose they do not. Then our drawing contains a
plane embedding of the subgraph Hi induced by the vertices of the elements
of Xi , which is isomorphic to C3 × K2 . It has five regions, two bounded by
triangles and three by 4-cycles. The triangle Ti+2 must lie in one of these
regions, say R. If R were bounded by Ti , the cycle induced by U would meet
Ti in a point other than ui , which would be a crossing on an edge of Ti . Thus
R is not bounded by Ti , and similarly not by Ti+1 . Without loss of generality,
we may assume that R is bounded by the 4-cycle (ui , vi , vi+1 , ui+1 , ui ). Now
the cycle induced by W must meet this 4-cycle in two points, which are
crossings on edges of Xi .
17 Vertex colorings
A vertex coloring, or just a coloring, of a graph G from a set C of colors is a
function γ : V (G) → C such that for adjacent vertices u and v, γ(u) 6= γ(v).
It is a k-coloring if the image of γ has k elements, and the chromatic index
χ(G) is the minimum k for which G has a k-coloring. G is k-colorable
if χ(G) ≤ G. It is clear that χ(G) is the maximum of χ(H) over the
components H of G.
Lemma 17.1. For any graph G, χ(G) is the maximum of χ(B) over all
blocks B of G.
Proof. We may assume that G is connected. Let k be the maximum over
blocks; clearly k ≤ χ(G). We prove the reverse inequality by induction on
the number of blocks, the case of a single block being trivial. Suppose that
G has more than one block. Let B be an end-block, and let G′ be the union
of the other blocks. Then B and G′ meet in a single cut-vertex u of G.
By induction, B and G′ are k-colorable. We may color them both from the
same set of k colors, so that u receives the same color from B and G′ , and
this gives a k-coloring of G.
68
A graph has chromatic index 1 iff it is empty, and chromatic index 2 iff
it is non-empty and bipartite. More generally, a k-coloring of G partitions
V (G) into k color classes, the sets of vertices of each color, and G is k-
partite with respect to these sets. Conversely, if G is k-partite with partite
sets U1 , . . . , Uk , we may color G from {1, . . . , k} by setting γ(u) = i for
u in Ui . Some classes of graphs have easily determined chromatic indices:
χ(Kn ) = n, χ(Pn ) = 2 for n > 1, and χ(Cn ) = 2 if n is even and 3 if n is
odd. In general, though, determining χ(G) is hard.
A graph G is k-chromatic if χ(G) = k, and, for k ≥ 2, critically k-
chromatic if in addition χ(G −u) < k (and hence χ(G −u) = k −1) for every
vertex u. By Lemma 17.1, a critically k-chromatic graph is non-separable.
The only critically 2-chromatic graph is K2 . An odd cycle is critically 3-
chromatic. Conversely, if G is critically 3-chromatic, it is not bipartite, and
so contains an odd cycle. Let C be a shortest odd cycle in G. If vertices
of C are not adjacent in C, they are not adjacent in G, since otherwise we
could find two shorter cycles, one of them odd. If G had a vertex not on C,
then G − u, containing an odd cycle, would be 3-chromatic, a contradiction.
Thus G is an odd cycle.
69
From the inequality κ1 (G) ≤ δ(G) of Theorem 6.3 we have the following
corollary.
Theorem 17.5. For any graph G, χ(G) ≤ max δ(G′ ) + 1, where the maxi-
mum is over all subgraphs G′ of G.
70
Figure 17: The Grötzsch graph
Theorem 17.6. For any graph G, let l(G) be the maximum length of a path
in G. Then χ(G) ≤ l(G) + 1.
Proof. By the previous theorem there is a subgraph G′ of G with χ(G) ≤
δ(G′ ) + 1, and by Exercise 3.1 or Theorem 7.8, l(G) ≥ l(G′ ) ≥ δ(G′ ).
71
Proposition 17.8. Let G be a triangle-free graph of order n ≤ 11 with
χ(G) ≥ 4. Then G is isomorphic to the Grötzsch graph.
Proof. We assume that n is minimal, and that amongst triangle free graphs
of order n with chromatic index at least 4, G is maximal. If we prove that
G is isomorphic to the Grötzsch graph, the result will follow since deleting
any edge of the Grötzsch graph leaves a 3-chromatic graph. Since adding
edges does not decrease the chromatic index, the maximality requirement
means that G has diameter 2.
By Theorem 17.4, ∆(G) ≥ 4. Let w be a vertex of maximum degree,
V the set of its neighbors, and U the set of all other vertices of G. Then
|U | ≤ 6. The subgraph induced by U is not bipartite, since otherwise we
could 2-color it, give the vertices of V a third color, and w one of the original
2 colors to obtain a 3-coloring of G. Hence there is a 5-cycle C with vertices
in U , and at most one vertex x of U not on C. Any vertex v ∈ V is a
adjacent to at most two vertices of C. Any vertex u of C not adjacent to v
has a common neighbor with v, which must be one of its neighbors in C, or
x. It follows that v is adjacent to two vertices of C. Moreover, distinct v1
and v2 in V cannot be adjacent to the same two vertices of C, since then the
set of neigbors of one of them, say v1 , would be a subset of the neighbors of
the other. A 3-coloring of G − v1 could then be extended to a 3-coloring of
G by giving v1 the same color as v2 .
If ∆(G) = 5, it follows that G is isomorphic to the Grötzsch graph.
Otherwise, let G′ be G − x if x exists, and G if not. Then G′ is isomorphic
to the Grötzsch graph with some vi (in the notation of the proof of Theorem
17.7) deleted. Hence there is a 3-coloring of G′ in which some color appears
only at two vertices of degree 4 in G′ . Since x, if it exists, cannot be
adjacent to these vertices, it can be given the same color to 3-color G, a
contradiction.
72
region chromatic index χ∗ (G) is the minimum k for which G has a region
k-coloring. G is region k-colorable if χ∗ (G) ≤ k.
The basic idea here is that a coloring of the vertices of a plane graph
corresponds to a coloring of the regions of its dual. The dual of a plane graph
G is obtained by taking a vertex in every region and taking, for each edge
e of G, an edge in the dual adjacent to the vertices in the adjoining regions
and crossing e at one point. However, the dual is generally a multigraph,
and we are avoiding multigraphs in our formal treatment. (Loops in the
dual correspond to bridges in G, and parallel edges arise when regions of G
have more than one edge in common.) Therefore our proof will avoid taking
the dual of a general plane graph.
Proof. Suppose first that every bridgeless cubic planar graph is region k-
colorable. It is enough to prove that every maximal planar graph G is
k-colorable. Take a plane embedding of G. Then the dual G∗ of G is a
graph, is bridgeless and cubic, and has dual G, so a region coloring of G∗
gives a vertex coloring of G using the same colors.
Suppose now that every planar graph is k-colorable, and let G be a
bridgeless plane graph, which we may assume is connected. Form a plane
graph H as follows. Take a vertex in every region of G. For any regions
of G with an edge in common, pick one such, e, and take an edge in H
whose vertices are those in the regions and which crosses e at one point.
A coloring of the vertices of H gives a region-coloring of G using the same
colors. (Here the dual of G has no loops since G is bridgeless, and H is
obtained by dropping all but one of each set of parallel edges.)
For the early history of the four-color theorem, see [4]. A fallacious
proof was published by Kempe in 1879. The error was only pointed out by
Heawood in 1890. The four-color theorem was then reduced to the status
of a conjecture, only becoming a theorem again in 1976 (Appel, Haken and
Koch [1]). Even the simplified proof of Robertson, Sanders, Seymour and
Thomas [27] relies heavily on the use of a computer, and no more will be
said about it here. We give Kempe’s incorrect proof, Heawood’s observation
73
y b
b r r y
g g
r y b g y b r g
r y
g y g b
b r
74
It remains to consider the case where u has degree 5, with neighbors
v1 , . . . , v5 in counterclockwise order. By maximality, (v1 , . . . , v5 , v1 ) is a
cycle, and so without loss of generality the colors of v1 , . . . , v5 are b, r, y, g,
r respectively. If the by-chain containining v1 does not contain v3 , we are
done as before, so we assume it does. Hence the rg-chain containing v2 does
not contain v4 . Similarly we may assume that the bg-chain containing v1
contains v4 , and so the ry-chain containing v5 does not contain v3 . Kempe
thought that these two chains could be used to get a coloring of G − u
in which red does not appear at a neighbor of u, completing the proof.
Heawood pointed out that switching colors on one chain can alter the other,
so that the new chain contains two neighbors of u, and gave the example
shown in Figure 18. Here G − u and its coloring are shown; u is in the
unbounded region. It will be seen that the relevant by- and bg-chains have
the properties stated, and changing colors on the rg- and ry-chains, in either
order, leaves a coloring with all four colors at neighbors of u.
19 Edge colorings
An edge coloring of a graph G from a set C of colors is a function γ : E(G) →
C such that for adjacent edges e and f , γ(e) 6= γ(f ). It is an edge k-coloring
if the image of γ has k elements, and the edge chromatic index χ1 (G) is
the minimum k for which G has an edge k-coloring. G is edge k-colorable if
χ1 (G) ≤ G. Clearly ∆(G) ≤ χ1 (G).
75
Proof. This is by induction on the size, the case of an empty graph being
obvious. Suppose that G is non-empty and the result holds for any graph
with fewer edges. Let C be some fixed set of ∆(G) + 1 colors. In this proof,
an edge coloring from C will be called just a coloring. Let e = uv0 be an
edge of G, and set H = G − e. Since ∆(H) ≤ ∆(G), H has a coloring. For
any coloring γ of H and vertex x of G, we say that a color c ∈ C appears at x
if there is an edge f of H incident with x such that γ(f ) = c. At any vertex,
there is at least one color that does not appear. Suppose that, for some
coloring γ of H and integer k ≥ 0, there are distinct vertices v0 , v1 , . . . , vk
adjacent to u such that
(1) for 1 ≤ i ≤ k, the color ai = γ(uvi ) does not appear at vi−1 ;
(2) there is a color b that appears at neither u nor vk .
Then we obtain a coloring of G by giving uvi−1 the color ai for 1 ≤ i ≤ k,
uvk the color b, and every other edge its color under γ.
Now let γ be an arbitrary coloring of H, and suppose we cannot find
vertices as above. Let k be the maximum integer for which there exist
distinct vertices v0 , v1 , . . . , vk adjacent to u satisfying (1). (There is such
an integer, because (1) is vacuously true if k = 0.) There is some color not
appearing at vk , which must appear at u, and by maximality of k it must
be aj for some j with 0 < j < k. Let b be a color not appearing at u, which
must appear at every vi for 0 ≤ i ≤ k. Consider the subgraph K of H
induced by the edges colored aj or b. Every component is either a cycle or
a path, and the endpoints of the paths are the vertices at which just one of
aj and b appears. One such vertex is vk ; let the component of K containing
it be P , and let w be the other endpoint of P . Define a new coloring γ ′ of
H by switching aj and b on the edges of P . Note that at every vertex other
than vk and w the same colors appear in γ and γ ′ . We claim that there is
some k ′ , 0 ≤ k ′ ≤ k, such that, with respect to γ ′ ,
(1′ ) for 1 ≤ i ≤ k ′ , the color a′i = γ ′ (uvi ) does not appear at vi−1 ;
(2′ ) there is a color b′ that appears at neither u nor vk′ ,
which will complete the proof. Note that a′i = ai except perhaps for i = j.
Suppose first that w = u, so the vertex adjacent to w in P is vj . Then we
may take k ′ = j − 1 and b′ = aj . If w 6= u then we also have a′j = aj .
Suppose that w = vl , where 0 ≤ l < k and l 6= j. Since b appears at vl in
γ, it does not in γ ′ and we may take k ′ = l and b′ = b. The only remaining
possibility is that w is different from u and all vi , 0 ≤ i ≤ k, and then we
may take k ′ = k and b′ = b.
76
A graph G is of class 1 if κ1 (G) = ∆(G), and of class 2 if κ1 (G) =
∆(G) + 1.
Theorem 19.2. A regular graph of class 1 has even order.
Proof. Suppose G is d-regular of class 1. In any edge d-coloring of G, every
color appears at every vertex. Thus the edges with any given color span G.
Since they are pairwise non-adjacent, the result follows.
The converse is false; see Exercise 19.1. The existence of edge 3-colorings
of cubic graphs is related to the four-color theorem by the following result.
Theorem 19.3 (Tait). Let G be a connected, bridgeless, cubic plane graph.
Then G is edge 3-colorable iff it is region 4-colorable.
Proof. We color regions using Z2 ×Z2 and edges using the non-zero elements.
Given a region coloring, we color each edge with the sum of the colors of the
adjacent regions, which is non-zero because adjacent regions have different
colors. Since the three regions around a vertex have different colors, so do
the three edges.
Conversely, given an edge coloring, give one region the color (0, 0). Then
the color of every other region is determined by the rule that the sum of
the colors of adjacent regions is the color of their common edge. This is
well-defined because at every vertex the colors of the edges are the three
non-zero elements of Z2 × Z2 , whose sum is zero.
19.2. Let n be odd. Show that in any edge n-coloring of Kn , each color
appears (n − 1)/2 times.
77
Figure 19: The flower snark
The condition that d < n − 1 just rules out complete graphs, and the
requirement that G be connected rules out graphs whose components are all
complete of the same order. We have 0 ≤ λ < d, 0 < µ ≤ d and diam G = 2.
Theorem 20.2. Let G be strongly regular with parameters (n, d, λ, µ). Then
d(d − λ − 1) = (n − d − 1)µ.
78
Theorem 20.4. Let G be strongly regular with parameters (n, d, λ, µ). Then
!
1 (n − 1)(µ − λ) − 2d
2 n−1±
p
(µ − λ)2 + 4(d − µ)
are positive integers.
Proof. By Lemma 20.3, every eigenvector of A is an eigenvector of J. Now
J has a 1-dimensional eigenspace with eigenvalue n, generated by x1 =
(1, 1, . . . , 1), and an (n − 1)-dimensional eigenspace with eigenvalue 0. The
vector x1 is an eigenvector of A, with eigenvalue d. The eigenvalue of any
independent eigenvector of A must be a root of x2 + (µ − λ)x − (d − µ). The
roots of this quadratic are
p
r± = 21 −µ + λ ± (µ − λ)2 + 4(d − µ) .
79
Exercises for §20
21 Moore graphs
Theorem 21.1. Let G be a connected graph of order n, diameter k and
Pk−1
maximum degree ∆. Then n ≤ 1 + ∆ i=0 (∆ − 1)i .
Definition 21.2. A Moore graph is a graph for which equality holds in the
inequality of Theorem 21.1.
This is the original definition of a Moore graph. The term has since
been extended to a larger class of graphs, as will be discussed later. If G
is a Moore graph, every inequality in the above proof must be an equality.
This implies that G is regular, rad G = diam G, and that for every vertex
v0 the subgraph induced by the edges of all geodesics starting at v0 is a tree
T , and the vertices of an edge not in T are end-vertices of T . The graphs of
diameter less than 2 are the complete graphs, and these are Moore graphs.
The 2-regular Moore graphs are the odd cycles. Recall that the girth of a
graph that is not a forest is the length of a shortest cycle. Clearly the girth
of a Moore graph of diameter k is 2k + 1.
80
Proof. As before, fix a vertex v0 and let Sr be the set of vertices v with
d(v0 , v) = r. Then S0 = {v0 } and |S1 | ≥ δ. For 0 < r ≤ k, every vertex of
Sr is adjacent to only one vertex of Sr−1 , since otherwise there would be a
cycle of length at most 2r (by Lemma 7.5). If there are adjacent vertices of
Sr then there is a cycle of length at most 2r + 1, so r ≥ k. It follows that
|Sr | ≥ δ(δ − 1)r−1 for 0 < r ≤ k.
81
k from v, and P , a v-w path of length k and the edge wu give a (2k + 1)-
cycle. Hence the number of (2k + 1)-cycles through u and v is deg u − 1, and
by symmetry in u and v, deg u = deg v. Suppose C = (v0 , v1 , . . . , v2k , v0 )
is a (2k + 1)-cycle, and take the subscripts modulo 2k + 1. The sequence
(v0 , vk , v2k , . . . , v(2k)k ) contains all vertices of C, and consecutive terms are
antipodes, so all vertices of C have the same degree.
Now fix a (2k + 1)-cycle C. For any vertex v, we may construct a path
of length k starting at v by taking a shortest path from v to a vertex of C,
followed by a path in C. Hence v has an antipode. It follows that no vertex
has degree 1, since its neighbor would not have an antipode. Hence any
path of length l < 2k can be extended to a path of length l + 1. This implies
that, for any vertices u and v there is some path of length k containing u
and v, and this is contained in a (2k + 1)-cycle. Thus deg u = deg v.
82
Figure 20: The Hoffman-Singleton graph
83
vertices of A and 2 of B.
Now consider a 5-cycle (v0 , v1 , v2 , v3 , v4 , v0 ). There are 25 · 5 · 4 · 4 = 2000
such cycles with v0 , v2 ∈ A and v1 , v3 ∈ B, and for each one there is
just one other cycle of the same form giving the same pentagon, namely
(v4 , v3 , v2 , v1 , v0 , v4 ) if v4 ∈ A and (v2 , v1 , v0 , v4 , v3 , v2 ) if v4 ∈ B, so these
cycles account for 1000 of the pentagons. There are 25 · 2 · 5 · 2 = 500 cycles
with v0 , v1 ∈ A and v2 , v3 ∈ B, and again for each one there is one other cycle
of the same form giving the same pentagon, namely (v0 , v4 , v3 , v2 , v1 , v0 ) if
v4 ∈ A and (v1 , v0 , v4 , v3 , v3 , v1 ) if v4 ∈ B. These cycles give a further 250
pentagons, so there are 10 pentagons not yet accounted for. These are given
by cycles in which v0 , v1 , v2 and v3 belong to the same one of A and B, and
there are 50 · 2 · 1 · 1 = 100 of these. If v4 is not in the same set, there is only
one other cycle of this form giving the same pentagon, (v3 , v2 , v1 , v0 , v4 , v3 ),
which gives too many pentagons. Hence A is the disjoint union of pentagons
P0 , . . . , P4 , and B of pentagons Q0 , . . . , Q4 .
Each vertex of Pi is adjacent to one vertex of every Qj , and vice-versa.
We shall label the vertices of Pi as xi,j and those of Qi as yi,j (i, j ∈ Z5 ),
in such a way that the edges are xi,j xi,j+1 and yi,j yi,j+2 . However this is
done, there are, for i, j ∈ Z5 , elements a(i, j) and b(i, j) = ±1 of Z5 such
that xi,k is adjacent to yj,a(i,j)+b(i,j)k . We may choose the labelling of the
x0,k arbitrarily, then that of the yj,l so that a(0, j) = 0 and b(0, j) = 1,
then that of the remaining xi,k so that a(i, 0) = 0 and b(i, 0) = 1. Suppose
that there exist i and j with b(i, j) = −1. Then for any k ∈ Z5 there is a
trail (xi,k , yj,a(i,j)−k , x0,a(i,j)−k , y0,a(i,j)−k , xi,a(i,j)−k) ) of length 4. However,
if k = 3a(i, j) this is a cycle, which is impossible. Thus xi,k is adjacent to
yj,a(i,j)+k . Since there are no 4-cycles, for i1 6= i2 and j1 6= j2 we have
In particular,
84
Following [19], we shall show that the only Moore graph of diameter 3
is the heptagon C7 . (It was subsequently proved by Bannai and Ito [2] and
Damerell [8] that all Moore graphs of diameter greater than 2 are odd cycles,
so we have expended a lot of effort determining the properties of the graphs
in a small, and, with the exceptions of the Petersen and Hoffman-Singleton
graphs, boring class. Such is life.)
For integers d ≥ 2 and r ≥ 0, define polynomials Fd,r ∈ Z[x] by
1
if r = 0;
Fd,r = x+1 if r = 1;
xFd,r−1 − (d − 1)Fd,r−2 if r ≥ 2.
Proof. The discussion after Definition 21.2 shows that if u and v are vertices
with d(u, v) = r > 0 then u has one neighbor at distance r − 1 from v, and if
r < k its other neighbors are at distance r + 1 from v. We show by induction
that, for 0 ≤ r ≤ k, the entry of Fd,r (A) corresponding to vertices u and v
is 1 if d(u, v) ≤ r and 0 otherwise, the cases r = 0 and 1 being obvious, and
the case r = k giving the lemma. Suppose that r ≥ 2 and the result is true
for smaller values. The entry of AFd,r−1 (A) corresponding to u and v is the
number of neighbors of u whose distance from v is at most r − 1, which is d
if d(u, v) ≤ r − 2; 1 if d(u, v) = r − 1 or r; and 0 if d(u, v) > r. The result
follows.
85
Corollary 21.11. There is no d-regular Moore graph of diameter 3 for
d ≥ 3.
Proof. We have Fd,3 = x3 + x2 − 2(d − 1)x − (d − 1). If this polynomial
2 (r+1)
is reducible, it has an integral root r. But then d − 1 = r 2r+1 , and since
2r + 1 is coprime to r and r + 1, this implies 2r + 1 = ±1 and d = 1, a
contradiction.
22 Kneser graphs
Let n and r be integers with 0 < r < n and let X be the set {1, . . . , n}.
We denote the collection of all r-element subsets of X by Xr . The Kneser
graph Kn(n, r) has vertex set Xr , and two vertices A and B are adjacent
iff they are disjoint. In the proof of Theorem 10.6 we showed that the
Petersen graph is isomorphic to Kn(5, 2). If s is an integer with 0 ≤ s < r,
the generalized Kneser graph Kn(n, r, s) also has vertex set Xr , and two
vertices A and B are adjacent iff |A ∩ B| ≤ s. Thus Kn(n, r) = Kn(n, r, 0).
If n < 2r − s then Kn(n, r, s) is empty, while if s = 0 and n = 2r then every
component is a K2 , so for the rest of this section we exclude these cases.
We shall always use t to denote n − 2(r − s); writing t as (n − (2r − s))+ s
makes it clear that our restrictions imply that t > 0. Sending A ∈ Xr to
X − A gives an isomorphism from Kn(n, r, s) to Kn(n, n − r, n − 2r + s), so
we could restrict attention to the cases n ≥ 2r, but this does not seem to
simplify any arguments significantly.
Lemma 22.1. Let A, B ∈ Xr . Then A and B have a common neighbor
in Kn(n, r, s) iff |A ∩ B| ≥ r − t.
Proof. Let k = |A ∩ B|. Suppose first that there is a common neighbor C
of A and B. Then |(A ∪ B) ∩ C| ≤ 2s, so
n − 2r + k = |X − (A ∪ B)|
≥ |C − (A ∪ B)|
≥ r − 2s,
86
s ≤ r −k, we may take sets A′ ⊆ A−B and B ′ ⊆ B −A with |A′ | = |B ′ | = s,
and set Y = A′ ∪ B ′ ∪ (X − (A ∪ B)). We have |A ∩ Y | = |B ∩ Y | = s and
|Y | = 2s + n − 2r + k ≥ r. If s > r − k, take Z ⊆ A ∩ B with |Z| = r − s and
set Y = X − Z. Then |A ∩ Y | = |B ∩ Y | = s and |Y | = n − (r − s) ≥ r.
|A ∩ B| ≥ |A ∩ B ∩ C|
= |A ∩ C| + |B ∩ C| − |(A ∩ C) ∪ (B ∩ C)|
≥ (r − pt) + (r − t) − r
= r − (p + 1)t.
Proof. By
Lemma 22.2, there is an A-B walk of length 2p + 1 iff there exists
C ∈ Xr with |A ∩ C| ≥ r − pt and |B ∩ C| ≤ s. Suppose such a C exists.
Then |A ∩ B| ≤ |A − C| + |B ∩ C| ≤ pt + s. Suppose conversely that
|A ∩ B| ≤ pt + s. If |A ∩ B| ≤ s we may take C = A. Otherwise, we may
take C1 ⊆ A ∩ B with |C1 | = s and, since |A − B| < r − s ≤ n − r = |X − B|,
C2 with A − B ⊆ C2 ⊆ X − B and |C2 | = r − s. Setting C = C1 ∪ C2 , we
have |C| = r, |A ∩ C| = |A − B| + s ≥ r − pt, and |B ∩ C| = s.
87
The odd girth of Kn(n, r) was determined by Poljak and Tuza [25], and
this was extended to Kn(n, r, s) under mild restrictions on n, r and s by
Denley [9].
Proof. For any graph, the odd girth is the minimal length of an odd closed
walk. By Lemma 22.3, there is a closed walk in Kn(n, r, s) of length 2p + 1
iff r ≤ pt + s, and the result follows.
88
have diam Kn(n, r, s) ≥ a, completing the proof. Certainly there exist A
and B with |A ∩ B| = k provided as ≤ k ≤ r, so it is enough to prove
that s ≤ r − t a−12 − 1 and s + t 2 − t + 1≤ r. The first inequality is
a−1 a a−1
equivalent to t 2 ≤ r − s − 1 and (since 2 ≤ 2 + 1) implies the
second. Now t a−1 2 ≤ 21 (r − s + t − 2), which is at most r− s − 1 provided
r−s−1
t ≤ r − s. On the other hand, if t > r − s then a = t + 1 ≤ 2, and so
a−1
t 2 = 0 ≤ r − s − 1.
Lemma 22.8. For 0 < k < r, f (k) ≥ f (k − 1), and f (r − 1) > f (r − 2).
Remark. We need not have f (k) > f (k −1), even if f (k) > 0. For example,
with n = 8, r = 4 and s = 2 we have f (1) = f (0) = 36.
89
We define a new graph with vertex set Xr , the maximal overlap graph
90
X
Proof. Let A, B and C be elements of r with A adjacent to both B and
C in M (n, r). We claim that:
A′ = A1 − {a1 } ∪ {a2 };
B ′ = B1 − {b1 } ∪ {a2 }.
91
A More topology
Here I fill in a few details from section 13 that my inner pedant insists must
be covered. I have tried, perhaps quixotically, to limit the prerequisites to
things that should have at least been stated in 1550 – 2057. The first lemma
is left as an exercise.
Lemma A.1. Let A and B be subsets of R2 . If A and B are open (resp.
closed), then A ∪ B and A ∩ B are open (resp. closed).
Lemma A.2. Let f : [0, 1] → [0, 1] be a continuous bijection. Then f ({0, 1}) =
{0, 1}.
Proof. Suppose, for a contradiction, that 0 < f (0) < 1. Since f is surjective,
there exist a, b ∈ [0, 1] with f (a) = 0 and f (b) = 1. By the Intermediate
Value Theorem, there is some c between a and b with f (c) = f (0), contra-
dicting the injectivity of f . A similar contradiction arises if 0 < f (1) < 1,
completing the proof.
Lemma A.3. Let f : [a, b] → R2 be continuous. The image of f is a closed
and bounded subset of R2 .
Proof. Let A = f ([a, b]). We show first that A is closed. Let x ∈ R2 −A, and
define g : [a, b] → R by g(s) = kf (s) − xk. By the Extreme Value Theorem,
g has a minimum value ǫ, and ǫ > 0. If y ∈ B(x, ǫ) then y ∈ / A, so R2 − A
is open and A is closed.
Now define h : [a, b] → R by h(s) = kf (s) − 0k. If M is the maximum
value of h then A ⊆ B(0, M ).
Lemma A.4. Let f and g be injective paths in R2 with the same image.
Then the composite h = g −1 ◦ f : [0, 1] → [0, 1] is continuous.
Proof. Let x0 ∈ [0, 1], y0 = h(x0 ). and ǫ > 0. We must show that there
exists δ > 0 such that x ∈ [0, 1] and |x − x0 | < δ imply that |h(x) − y0 | < ǫ.
Let Y = {y ∈ [0, 1] | |y − y0 | ≥ ǫ}. Replacing ǫ by a smaller number, if
necessary, we may assume that Y is a closed interval or a disjoint union of
two closed intervals. By Lemmas A.1 and A.3, g(Y ) is a closed subset of
R2 . Hence there exists γ > 0 such that B(f (x0 ), γ) ⊆ R2 − g(Y ), and then
δ > 0 such that x ∈ [0, 1] and |x − x0 | < δ imply f (x) ∈ B(f (x0 ), γ), and
therefore h(x) ∈
/ Y and |h(x) − y0 | < ǫ, as required.
The next result follows immediately from Lemmas A.2 and A.4.
Theorem A.5. Let f and g be injective paths in R2 with the same image.
Then f ({0, 1}) = g({0, 1}).
92
References
[1] K. Appel, W. Haken and J. Koch, Every planar map is four-colorable,
Illinois J. Math. 21 (1977) 429–567.
[2] Elichi Bannai and Tatsuro Ito, On Moore graphs, J. Fac. Sci. Univ.
Tokyo Sect IA 20 (1973) 191–208.
[3] H. Bielak and M.M. Syslo, Peripheral vertices in graphs, Studia Sci.
Math. Hungar. 18 (1983) 269–75.
[4] Norman L. Biggs, E. Keith Lloyd and Robin J. Wilson, Graph Theory
1736–1936, Clarendon Press, Oxford 1976.
[7] G. Chartrand and L. Lesniak, Graphs and Digraphs (4th edition), Chap-
man and Hall/CRC 2004.
[9] Tristan Denley, The odd girth of the generalised Kneser graph, Euro-
pean J. Combin. 18 (1997) 607–11.
[11] P. Erdös and T. Gallai, Graphs with prescribed degrees of vertices (Hun-
garian), Mat. Lapok 11 (1960) 264–74.
[14] Martin Gardner, Mathematical Games, Sci. Amer. 234 (1976) 126–30.
93
[15] S.L. Hakimi, On the realizability of a set of integers as degrees of the
vertices of a graph, J. SIAM Appl. Math. 10 (1962) 496–506.
[21] G. Kirchoff, Über die Auflösung der Gleichungen, auf welche man bei
der Untersuchung der linearen Verteilung galvanischer Ströme geführt
wird, Ann. Phys. Chem. 72 (1847) 497–508; Gesammelte Abhandlun-
gen, Leipzig (1882) 22–3.
[25] S. Poljak and Zs. Tuza, Minimum bipartite subgraphs of Kneser graphs,
Graphs Combin. 3 (1987) 191–9.
[26] H. Prüfer, Neuer Beweis eines Satzes über Permutationen, Arch. Math.
Phys. 27 (1918) 142–4.
[27] N. Robertson, D.P. Sanders, P.D. Seymour and R. Thomas, The four-
color theorem, J. Combin. Theory Ser. B 70 (1997) 2–44.
94
[29] Mario Valencia-Pabon and Juan-Carlos Vera, On the diameter of
Kneser graphs, Discrete Math. 305 (2005) 383–5.
95
List of errata
Problem 6.1 in the notes cannot be answered as stated. See the corrected
version in the separate collection of exercises.
Page Line reads should read
15 -9 0f of
16 13 (u0 , u1 , . . . , ul ) W1 = (u0 , u1 , . . . , ul )
16 13 (v0 , v1 , . . . , vk ) W2 = (v0 , v1 , . . . , vk )
16 14 0≤j≤k 0 ≤ j ≤ k and every edge
of W2 is an edge of W1
29 7 – 8 Theorem 5.14 Corollary 5.15
77 3 regular graph non-empty regular graph