Topological Graph Theory PDF
Topological Graph Theory PDF
Graph Theory
JONATHAN L. GROSS
Columbia University
New York, New York
THOMAS W. TUCKER
Colgate University
H ami/ton, New York
A Wiley-Interscience Publication
JOHN WILEY & SONS
New York • Chichester • Brisbane • Toronto • Singapore
Copyright© 1987 by Jonathan L. Gross and Thomas W. Tucker
All rights reserved. Published simultaneously in Canada.
Reproduction or translation of any part of this work
beyond that permitted by Section 107 or 108 of the
1976 United States Copyright Act without the permission
of the copyright owner is unlawful. Requests for
permission or further information should be addressed to
the Permissions Department, John Wiley & Sons, Inc.
1. INTRODUCfiON 1
References 319
Bibliography 333
In beginning the study of topological graph theory, one needs to prepare both
for the development of topological operations and for discussion of the
relationship between these operations and the other concepts of graph theory.
Most of the important topics in graph theory have some connection with the
topological viewpoint. Accordingly, this first chapter reads almost like a survey
of graph theory, encompassing trees, automorphisms, traversability, factoriza-
tion, colorings, algorithms, and more.
1 To a topologist, the word "simple" has pre-emptive meaning. In particular, a "simple" loop is
the continuous nonsingular image of a circle, that is, a loop without self-intersections or other
singularities. Our general rule is to avoid ambiguity or such conflicts of usage.
2 Introduction
1.1.1. Drawings
Any finite graph can be geometrically represented by a drawing obtained in
the following manner. First, draw a dot for each vertex. Then, for every edge
e E E with two endpoints, draw a line between the dots representing the
vertices of V( e), and for every edge with only one vertex, draw a line from
the dot representing that vertex to that dot itself. Two examples now illustrate
the representation of a graph by a drawing.
Figure 1.2. Two drawings of the graph of Example 1.1.2, one without line-crossings, the other
with crossings.
Representation of Graphs 3
e2 e3
n
1 0
0 1
1 1
0 0
2 The synonym "degree" appears to be more frequently used by graph theorists than "valence".
However, since "valence" is a self-explanatory abstraction from the terminology of chemistry, and
since "degree" is used for polynomials and as the name for a topological invariant of a continuous
function, it seems appropriate to use "valence" here for the graph-theoretic concept.
4 Introduction
Theorem 1.1.1 (Euler). The sum of the valences of the vertices of a graph equals
twice the number of edges.
n
vi v2 v3 v4 vi v2 v3
u n
vi 1 1 vi 1
v2
v3
v4
0
0
0
1
1
1
v2
v3 [i 0
2
1.1.5. Directions
A "direction" for an edge e is an onto function {BEGIN, END} ~ V(e). The
images of BEGIN and END are called the "initial point" and the "terminal
point", respectively. One says that a directed edge goes "from" its initial point
"to" its terminal point. In the incidence structure for a list format of a graph,
the direction of a proper edge may be indicated by putting the initial point
first in its endpoint list.
In topological graph theory, one frequently considers each edge e (even a
loop) to have two directions, arbitrarily distinguished as the "plus direction"
Representation of Graphs 5
Example 1.1.3. Figure 1.4 shows two graphs, G and G'. Suppose the function f
has the values
G G'
Figure 1.4. The domain G and the range G' of a graph map f: G .... G'.
6 Introduction
u§
H H'
Figure 1.5. Two isomorphic graphs.
The endpoints of the edge e1 are the vertices v 1 and v2 , which are mapped onto
v 1' and v2', respectively, the endpoints of the edge e 1', which is the image of e1 •
The endpoints v 1 and v3 of the edge e2 are mapped onto the endpoints v 1' and v2'
of the edge e 1', which is the image of e 2 • Both the endpoints v2 and v3 of the edge
e 3 are mapped onto the single endpoint v2' of its image e{ Thus the vertex map f
and the edge map f together form a graph map f: G ~ G'. However, f is not an
isomorphism.
Example 1.1.4. Figure 1.5 shows two isomorphic graphs, H and H'. One
isomorphism g: H ~ H' is given by the rule
1.1.7. Automorphisms
An isomorphism from a graph to itself is called an "automorphism". Under
the operation of composition, the family of all automorphisms of a graph
forms a group, called "the automorphism group of the graph" and denoted
Aut( G).
1.1.8. Exercises
1. Draw the graph with vertices v1 , v2 , and v 3 , with edges e 1 , e 2 , e 3 , e4 ,
and e 5 , and with endpoint sets V(e 1 ) = {v 1 , v2 }, V(e 2 ) = {v 2 , v3 },
V(e 3 ) = {v 1 , v3 }, V(e 4 ) = {v 1 , v3 }, and V(e 5 ) = {v 3 }.
2. Write the incidence stru~tures for the graphs of Figure 1.4.
3. What is the incidence matrix for the graph of Exercise 1?
4. What is the adjacency matrix for the graph of Exercise 1?
5. The "eigenvalues" of a graph are defined to be the eigenvalues of its
adjacency matrix. What are the eigenvalues of the graph in Figure 1.1?
6. The "valence sequence" of a graph is the list of all the valences of its
vertices (usually in increasing order), so that if there are n vertices, there
are n numbers in the valence sequence. Construct two graphs that have
the same valence sequence, but that are not isomorphic.
7. (a) Prove that there is no graph map from the graph G' of Figure 1.4 to
the graph G of that figure. (b) Prove that G and G' are not isomorphic.
8. Draw two nonisomorphic graphs with four vertices and four edges each
and no loops or multiple adjacencies.
9. Prove that if the graph map f: G ~ G' is an isomorphism, then the
r
inverse map 1: G' ~ G is also an isomorphism.
10. Find another isomorphism H ~ H' for the graphs of Figure 1.5 besides
the one given in Example 1.1.4.
11. In permutation notation, list all the automorphisms of the graph H of
Figure 1.5.
12. Draw a graph with two vertices and no automorphisms except for the
identity.
13. Draw a graph with six vertices, no loops, and no automorphisms except
the identity.
14. Identify the automorphism group of the graph of Figure 1.2.
15. Draw a graph whose automorphism group is isomorphic to the cyclic
group ~3 •
16. Prove that every graph has an even number of vertices of odd valence.
17. Two graphs are called "co-spectral" if they have the same eigenvalues
(see Exercise 5) with the same multiplicities. Draw two co-spectral,
connected, 6-vertex, 7-edge graphs.
whose initial vertex v0 is u, whose final vertex vn is v, and fori = 1, ... , n, the
directed edge e 1°1 runs from the vertex D;_ 1 to the vertex v,. If u-:/= v, then W
is called an "open walk". If u = v, then W is called a "closed walk".
An open walk is called a "path" if its vertices are distinct. Thus, a path is
the combinatorial analog of a homeomorphic image of a closed line segment.
The standard path with n vertices is called the "n-path" and is denoted Pn. A
closed walk is called a "cycle" if every pair of vertices except its starting and
stopping vertex are distinct. Thus, a cycle is the combinatorial counterpart to
the homeomorphic image of a circle. The standard cycle with n vertices is
called the "n-cycle" and is denoted en.
In order to relieve the cumbersomeness of the notation for a walk, path, or
cycle, one commonly omits the vertices from the sequence specified in the
definition, since there is no difficulty in inferring them from the resulting
sequence of directed edges. Moreover, when for i = 1, ... , n - 1 the edges e;
and ei+l have only one endpoint in common, one usually also omits the signs,
and writes
There are many different notations for a walk, path, or cycle, and the sensible
practice in a particular instance is to use the least cumbersome notation that
makes sense, even if it is not the same as the notation used in a previous
instance. Thus, fluency has priority over consistency.
A graph is called "connected" if for every pair of vertices u and v, there is a
path from u to v. Specifying plus directions on some of the edges does not
affect the connectedness of a graph, since the direction on an edge in a path
(or walk) is permitted to be the opposite of the plus direction. Thus, con-
nectedness in the present combinatorial sense agrees exactly with the underly-
ing topological sense.
1.2.2. Trees
A "tree" is a connected graph with no cycles, as illustrated in Figure 1.6. It is
one of the most important kinds of graphs in both applications and theory.
Theorem 1.2.1. Let u and v be any vertices of a tree T. Then there is a unique
path in T from u to v.
Some Important Gasses of Graphs 9
•
I
Figure 1.6. Four trees.
Theorem 1.2.3. Every nontrivial tree T has at least two vertices of valence 1.
Proof By Theorem 1.1.1, for any graph, the sum of the valences is twice
the number of edges. By Theorem 1.2.2, 2#Er = 2#Vr - 2, since T is a tree.
By combining these two facts, one obtains the equation
L valence{ v) = 2#Vr- 2
ve Vr
Since every vertex v of the tree has valence at least 1, there must be at least
two vertices of valence exactly 1. D
10 Introduction
0 0
4 1 4 1
--~--2
(a) (b)
Figure 1.8. Two Cayley color graphs for .2'5 : (a) with generating set {1}; (b) with generating set
{1, 2 }.
Some Important Gasses of Graphs 11
1 1
3 3 5
(a) (b)
Figure 1.9. (a) The Cayley color graph for the group .2"6 with the generators 2 and 3. (b) The
associated Cayley graph C(.2"6 , {2, 3 }) 0 .
12 Introduction
C( .J/1', X'). The following theorem shows the connection between Cayley color
graphs and group homomorphisms.
Proof Assume that the first hypothesis holds. To show that f is a group
homomorphism, it suffices to show that f(ax") = f(a)f(x)" for all a E .J/1,
x E X, and o = ± 1. Since there is an edge in the Cayley graph C(.J/1, X) from
a to ax with the color "x", it follows that in the Cayley graph C(.J/1', X')
there is an edge from /(a) to f( ax). Moreover, this edge has the same color as
the edge from 1..,, to f(x), since f is color-consistent, direction-preserving,
and identity-prese~ng. The edge from 1..,, to f(x) must be colored "f(x)", so
the edge from f(a) to f(ax) is also colored "f(x)". By the definition of
Cayley color graph, it follows that f(ax) = f(a)f(x). To show that f(ax- 1 )
= f(a)f(x)- 1, we now simply observe that f(a) = f(ax- 1x) = f(ax- 1 )/(x).
Conversely, let h: .J/1~ .J/1' be a group homomorphism such that h(X) c X'.
Define the graph map f: C(.J/1, X)~ C(.J/1', X') as follows. If a is a vertex of
C(.J/1, X), let f(a) = h(a). If e is an edge from a to ax (colored x), let f(e)
be the edge from f(a) to /(ax)= f(a)f(x) [colored /(x)]. It is easily verified
that f has the desired properties. D
of the Cayley color graph C(.J/1, X) collapse onto a single edge, denoted either
xa or xax· The "alternative Cayley graph" is denoted C(.J/1, X) 1. Both
C(.J/1, X) 0 and C(.J/1, X) 1 are called "Cayley graphs". Figure 1.10 shows a
Cayley color graph for a group with a generator of order 2 and the corre-
sponding alternative Cayley graph.
It may be observed that the Cayley color graphs in Figures 1.9 and 1.10
look quite alike, since their corresponding Cayley graphs are isomorphic, as
are their corresponding alternative Cayley graphs, even though the groups .2'6
Some Important Oasses of Graphs 13
(1 3) (1 3)
(1 2) (1 2) (2 3)
(a) (b)
(12 3) }
(12) - - - - - -
Figure 1.10. (a) The Cayley color graph for the symmetric group .9'3 with the generators (12 3)
and (1 2). (b) The alternative Cayley graph C( .9'3 , { (123), (12)} )1 .
Proof For any pair of vertices u and v, let w = vu- 1• Then the left
translation i w takes u to v. D
14 Introduction
The converse of Theorem 1.2.5 is not true; for example, the Petersen graph
(see Example 1.3.13 and Exercise 2.2.13) is not a Cayley graph, yet it is
vertex-transitive. On the other hand, a left translation ia of a Cayley graph
also has the property that it leaves no vertex fixed, unless of course the group
element a is the identity. If the automorphism group of a graph G has a
subgroup .91 that is vertex-transitive, and if every nonidentity element of .91
leaves no vertex fixed, then indeed the graph G is a Cayley graph for the group
.91 (see the corollary to Theorem 2.2.3 and see Theorem 6.2.3).
The Cayley graph associated with a given group and generating set reveals
much about the group. Every walk in the graph corresponds to a "word" in
the generators; every closed walk corresponds to a "relator", namely a product
of generators that is equal to the identity. This relationship has proved to be a
powerful tool in combinatorial group theory, which is the study of presenta-
tions of groups in terms of generators and relations. It was Dehn (1911) who
first focused on this connection. For example, one of Dehn's three famous
decision problems for a presentation of a group is to find an algorithm that
will determine whether two given words in the generators represent the same
element; this is equivalent to finding an algorithm that constructs the Cayley
graph corresponding to the given presentation of the group. [Boone (1954) and
Novikov (1955) showed this "word problem" was unsolvable for some group
presentations.]
Dehn's influence is so great that Cayley graphs are sometimes called "Dehn
Gruppenbilder", although Cayley's priority is unquestioned [Burnside's fa-
mous Theory of Groups (1897, 1911) refers to "Cayley's colour groups" and
even has a four-color illustration of a Cayley graph as the frontispiece.] More
recent examples of the importance of graphical representations of groups can
be found in the work of Stallings (1971) and Serre and Bass (1977) (see also
Tretkoff, 1975, 1980). Cannon (1984) carries the word problem one step
further by studying presentations for infinite groups that have linear-time
algorithms to draw the associated Cayley graph. General monographs for
combinatorial group theory have been compiled by Magnus, et al. (1966), by
Lyndon and Schupp (1977), and by Coxeter and Moser (1957).
every point of W, then the graph is called "complete bipartite" on the sets U
and W. The standard model for a complete bipartite graph on sets of m
vertices and n vertices is denoted Km, n. Figure 1.11 illustrates the complete
bipartite graph K 3, 4 •
1.2.7. Exercises
1. In the graph of Figure 1.10, find a path with six vertices and a cycle of
length five. Also find a closed walk with four edge-steps that is not a
cycle.
2. Prove that if .511 and .511' are groups of the same order, they have
respective generating sets X and X' such that the Cayley graphs
C(.Jil, X) 0 and C(.Jil', X') 0 are isomorphic. (Hint: Consider relatively
large generating sets.)
3. Show that a graph is connected if and only if for every pair of vertices u
and v, there is a walk from u to v.
4. Extend Theorem 1.2.4 to the case where the group .511 is infinite.
5. Draw the Cayley color graph for the group fr2 X fr2 X fr2 with genera-
tors (1, 0, 0), (0, 1, 0), and {0, 0, 1).
6. Draw the Cayley graph for the group fr8 with generating set {1, 2, 3}.
7. Draw the alternative Cayley graph for the group :Z6 with generating set
{1, 2, 3}.
8. Draw the Cayley color graph for the quaternions with generators i and j
and relations i 2 = / = {ij) 2 •
9. Draw the Cayley color graph for the alternating group .5114 with genera-
tors (1 2 3) and (2 3 4).
10. Draw a 4-regular simplicial graph on 11 vertices that is not a Cayley
graph. (Hint: Use two copies of K 5 .)
16 Introduction
1.3.1. Subgraphs
If G and G' are graphs, then G' is called a "subgraph" of G if and only if VG'•
EG'• and JG, are subsets of VG, EG, and IG, respectively. If in addition,
VG' = VG, then one says that the subgraph G' "spans" the graph G. Figure
1.13 illustrates a graph at the left and two of its subgraphs.
If a subgraph is a tree and if it spans the graph, it is called a "spanning
tree". For instance, the graph in the middle of Figure 1.13 is a spanning tree
for the graph at the left. The subgraph at the right does not span, since it does
not include the vertex v4 • Nor is it a tree, since it is not connected.
u2 u2 u2
ul
u3
e6
u4 ul u3
e6
u4 u
A 3
u5 u5 •
u5
Figure 1.13. A graph and two of its subgraphs.
New Graphs from Old 17
G Gl
Figure 1.14. A graph and two of its subgraphs.
=
For any graph G, let V' c Vc. The "induced subgraph" on the vertex
subset V' is the subgraph whose vertex set is V' and whose edge set consists of
every edge e of G whose endpoint set V(e) lies in V'. For instance, in Figure
1.14, the subgraph G 1 is an induced subgraph of G, but the subgraph G2 is
not.
There is a big difference between assertions about subgraphs and about
induced subgraphs. For example, since the complete graph Kn is, in fact, a
Cayley graph (Exercise 1.2.12), it follows that every finite graph is a subgraph
of some Cayley graph. The claim, however, that every finite graph is an
induced subgraph of some Cayley graph, seems quite surprising. This would
imply that although Cayley graphs are globally symmetric, locally they can be
arbitrarily complicated. Nevertheless, Babai (1976) has shown this claim to be
true. In fact, Godsil and Imrich (1985) prove that given any graph G on n
vertices and any group .91 of order greater than (24/3)n 3 there is a Cayley
graph for .91 containing G as an induced subgraph.
A "component" of a graph is a maximal connected subgraph. Thus, if a
graph is connected, it has only one component, itself.
u4 u4
Figure 1.15. The graph at the right is obtained from the graph at the left by subdividing the
edge e.
Example 1.3.2. For any two positive integers m and n, the m-cyc/e em and the
n-cyc/e en
are homeomorphic graphs.
Example 1.3.3. For any two integers m, n :;:: 2, them-path Pm and then-path
Pn are homeomorphic graphs.
ul
,, [ X
uJ.
eJ.
u2
e2
u§
u2
(ul, ul) (ul, u2) (ul, u 3)
(u 1, ei) (ul, e2)
(el, u2)
(u2 , ei)
Example 1.3.4. The Cayley graph C(.2'6 , {2, 3}) 0 illustrated in Figure 1.9 is
isomorphic to the cartesian product of the 2-cyc/e c2 and the 3-cyc/e c3 .
Example 1.3.5. The alternative Cayley graph C( .51'3 , { (123), (12)} ) 1 illustrated
in Figure 1.10 is isomorphic to a cartesian product of the 3-cyc/e C3 and the
complete graph K 2 •
1.3.4. Edge-Complements
The "edge-complement" of a simplicial graph G is denoted Gc and is defined
to be the graph with the same vertex set as G but such that two vertices are
adjacent if and only if they are not adjacent in G.
20 Introduction
Example 1.3.7. The edge-complement Knc of the complete graph has no edges at
all but has n vertices.
1.3.5. Suspensions
The "suspension" (elsewhere called the "join") of a graph G from a graph G'
is obtained by adjoining every vertex of G to every vertex of G', and is
denoted G + G'. Thus, its vertex set is Vc u Vc, and its edge set is Ec u Ec,
u (Vc X Vc,). The endpoints of an edge of G + G' that arises from Ec or
from Ec, are exactly as in the graphs G or G', respectively. The endpoints of
an edge ( v, v') that arises from Vc X Vc, are the vertices v and v'. Figure 1.18
shows the suspension of a 3-path from a 2-path.
1.3.6. Amalgamations
Let G and G' be graphs, and let f: H ~ H' be an isomorphism from a
subgraph H of G to a subgraph H' of G'. The "amalgamation" G • 1 G' is
obtained from the union of G and G' by identifying the subgraphs H and H'
according to the isomorphism. Figure 1.19 shows the amalgamation of a
3-cycle and a 4-cycle along the thickened edge.
f(v2)
Figure 1.19. The amalgamation of two graphs.
Then the group !!J is said to "act on (the left of) the graph" G. If, moreover,
the additional condition
holds, then !!J is said to "act without fixed points" (or sometimes "act freely")
on G.
For any vertex v of the graph G, the "orbit" [v] (or sometimes [v]&r) is
defined to be the set {<Ph( v ): b E !!J} of all images of v under the group
action. Similarly, for any edge e, the orbit [e) (or sometimes [e]&r) is defined to
be the set {<Pb(e):bE!!J}. The sets of vertex orbits and edge orbits are
denoted Vj!!J and Ej!!J, respectively. It may be observed that the orbits
partition the vertex set and the edge set of G.
Example 1.3.13. Figure 1.20 contains a drawing of the" Petersen graph" in the
plane. ForiE fr5 , let <P, be the rotation of 27Tij5 radians. Under this action of
fr5 , there are two vertex orbits, one containing the outer five vertices and the
other containing the inner five. There are three edge orbits, one containing the
five edges of the outer pentagon, one containing the five edges of the inner star,
and one containing the five edges that run between the star and the pentagon.
The "regular quotient" G1114 is the graph with vertex set VI 114 and edge set
E 1114 such that the vertex orbit [ v 1is an endpoint of the edge orbit [e 1if any
vertex in [ v 1 is an endpoint of any edge in [e 1. It is easy to verify that each
edge orbit has one or two vertex orbits as endpoints.
Example 1.3.13 Revisited. Figure 1.21 shows the quotient of the Petersen graph
under the prescribed action of .2'5 •
Example 1.3.14 Revisited. The quotient of the n-cycle Cn under the action of the
cyclic group .2'n is the bouquet B 1 of one circle. In general, when a group acts on
a Cayley graph for itself, the quotient is a bouquet of circles, one corresponding
to each generator.
1.3.9. Exercises
1. How many different spanning trees are there in the graph at the left of
Figure 1.13? How many different isomorphism types of spanning trees
are there? How many isomorphism types of subgraphs are there?
2. Prove that every graph is homeomorphic to a bipartite graph. (Hint: Use
the operation of subdivision.)
3. How many vertices does the n-cube graph Qn have? How many edges?
4. How many vertices and edges does the n-octahedron graph On have?
5. A suspension K 1 + Cn is called a "wheel". Why?
6. Draw the suspension K/ + C4c, the suspension K/ + C6 , and the carte-
sian product K 2 X K 1, 3 .
7. Describe a free action of the cyclic group ~3 on the complete bipartite
graph K 3 , 3 and draw the quotient.
8. Describe a free action of the cycle group ~2 on the 3-cube graph Q 3
whose quotient is isomorphic to the complete graph K 4 •
9. Prove that the cyclic group ~7 cannot have a free action on the Petersen
graph.
24 Introduction
10. Prove that the cyclic group .2'6 cannot have a free action on the complete
bipartite graph K 3, 3 •
11. Prove that the even-order cyclic group .2'2 n cannot have a free action on
the complete graph K 2 n.
12. Prove that the odd-order cyclic group .2'2 n+ 1 has a free action on the
complete graph K 2 n + 1 •
13. Construct a simplicial graph whose automorphism group acts freely on it
and is a nontrivial group.
14. Prove that a local isomorphism from a connected graph onto a tree is a
global isomorphism.
15. Prove that every automorphism on a tree has either a fixed vertex or a
fixed edge. (Hint: Those familiar with the "center" of a tree may find
that concept helpful.)
16. Prove that a covering projection p: G ~ G with a simplicial base graph
G is a local isomorphism.
17. Let G and G' be graphs, and let v E Vc and v' E Vc'. Prove that the
induced subgraphs on the vertex sets Vc X { v'} and { v} x Vc' are
isomorphic toG and G', respectively, in the product graph G X G'.
e e
C> C>
e
fffiL
Figure 1.23. Adding a handle to eliminate an edge-crossing.
.______,X..__ _.
Figure 1.24. Constructing a Mobius band.
26 Introduction
1.4.3. Imbeddings
To formalize the notion of a drawing without crossings, we define an "imbed-
ding" of a graph in a surface to be a continuous one-to-one function from a
topological representation of the graph into the surface. For most purposes, it
is natural to abuse the terminology by referring to the image of the topological
representation as "the graph".
If a connected graph is imbedded in a sphere, then the complement of its
image is a family of "regions" (or "faces"), each homeomorphic to an open
disk. In more complicated surfaces, the regions need not be open disks. If it
happens that they are all open disks, then the imbedding is called a "2-cell (or
cellular) imbedding". If the boundary circuit of an open disk region has one or
more repeated vertices, then the closure of the region is not a closed disk.
Nonetheless, whether the imbedding is a 2-cell imbedding depends only on
whether all the regions are open disks, not on whether the closures of the
regions are closed disks.
A convenient notation for the set of regions of a graph imbedding i: G --+ S
is Fe, where the letter F reminds one that the regions are something like the
faces of a polyhedron. If more than one imbedding of G is under considera-
tion, then the name of the imbedding should appear somewhere in the
notation for the set of regions. An even greater convenience, if G is the only
graph in the immediate discussion, is to use the notation F instead of Fe.
The "number of sides" (or "size") of a region f is defined to be the number
of edge-sides one encounters while traversing a simple circuit just inside the
boundary of the region, and is denoted s1 . As illustrated by the exterior region
Surfaces and Imbeddings 27
in Figure 1.25, the number of sides of a region need not equal the number of
edges in its boundary. Although there are seven sides to the exterior region,
there are only five edges in its boundary.
It may be observed that the imbedding in Figure 1.25 has a one-sided
region, which is called a "monogon". It also has a two-sided region, which is
called a "digon". (Some persons use the words "unigon" and "bigon", which
are mixtures of Latin and Greek. Fortunately, these mixtures have not become
as firmly entrenched as "hexadecimal".)
If every region of a graph imbedding is three-sided, then one says the
imbedding is "triangular". In a "quadrilateral imbedding", every region is
four-sided.
#V- #E + #F = 2.
Cauchy (1813) gave basically a graph-theoretic proof of this "Euler equation",
but it was Lhuilier (1811) who sorted out the apparent exceptions (faces must
be 2-cells and the surface cannot have holes) and generalized the Euler
equation to all closed orientable surfaces. Listing (1861) later used the Euler
equation to study surfaces in detail; this work is commonly considered the
beginning of topology.
The left side of the Euler equation is called the "Euler formula". The right
side is called the "Euler characteristic" of the surface. Thus, the number 2 is
the Euler characteristic of the sphere. Before proving the Euler equation for
the sphere, it may be instructive to evaluate the Euler formula for the two
examples depicted in Figure 1.26.
The graph G in Figure 1.26 has five vertices and nine edges. Each of the
regions, including the "exterior" region, is three-sided. There are six regions in
all, since the exterior region is always counted in computations of the Euler
formula. Thus, #V - #E + #F = 5 - 9 + 6 = 2. The graph H in Figure
1.26 has four vertices and five edges. The imbedding has three regions. Thus,
#V- #E + #F = 4 - 5 + 3 = 2.
G H
Proof The sum 'Lf E F sf of the numbers of sides of the regions counts
every edge exactly twice. Thus, 2#£ = '[,fEFsf. Since there are no loops or
multiple edges in the simplicial graph G, there are no monogons or digons in
the imbedding. Thus, for every region f, sf~ 3. It follows that 2#£ ~ 3#F.
D
The complete graph K 5 and the complete bipartite graph K 3, 3 are called
"Kuratowski' s graphs" (or the "Kuratowski graphs") because Kuratowski
(1930) proved that they are a complete set of obstructions to imbedding
graphs in the sphere, in the following sense:
Kuratowski's Theorem. The graph G has an imbedding in the sphere if and only
if it contains no homeomorph of K 5 or of K 3 , 3 •
Surfaces and Imbeddings 29
Theorem 1.4.3. Let G be a connected graph that is not a tree, and let i: G ~ S
be an imbedding. Then 2#£ ~ girth(G) · #F.
1 0 1
I I
·! iI
.<(!>' ~ t
3
I
4
a
I
l! '0), it
4 3
I
I
a
I I I I
I
- - (}-------j
-®
Figure 1.28. Constructing a torus from a rectangle.
It has already been proved that neither K 5 nor K 3 , 3 imbeds in the sphere.
From Figure 1.29 it now follows that both have orientable genus 1. Although
K 5 has the same nonorientable genus as orientable genus, as does K 3• 3 , for
most graphs the orientable genus and the nonorientable genus differ.
1.4.8. Duality
Given a connected graph G, a closed surface S, and a 2-cell imbedding
i: G -+ S, there is an idea due to Poincare for constructing what are called a
"dual graph" and a "dual imbedding". First, for each region f of the
imbedding i: G -+ S, place a vertex f* in its interior. Then, for each edge e of
the graph G, draw an edge e* between the vertices just placed in the interiors
of the regions containing e. (If both sides of the edge e lie in the same region
j, then the dual edge e* is a loop based at the dual vertex f*.) The resulting
graph with vertices f* and edges e* is called the "dual graph" for the
imbedding i: G-+ S and is denoted G*', or simply G*, if i is the only
imbedding under consideration. The resulting imbedding of the graph G* in
the surface S is called the "dual imbedding". In this context, the original
graph G and the original imbedding i are called the "primal graph" and the
"primal imbedding". Figure 1.30 shows how two different imbeddings of the
same graph in the same surface can yield different dual graphs.
In Figure 1.30 the primal graph is drawn with solid dots for vertices and
solid lines for edges, whereas the dual graphs are drawn with open dots for
vertices and dashed lines for edges. Since the dual graph at the left has a
vertex of valence 7, and the maximum valence in the dual graph at the right is
-- b
- b
r-r----- - - - - - - .....71
1 I ,..... I
I I ,........... I
I I ,........... I
I I ,........... I
.....
I
I
I
--------1
_______ j
Figure 1.31. A graph in the torus and its dual.
6, the two duals are not isomorphic. It may be observed that for any region f
of the primal imbedding, the valence of the dual vertex f* equals the number
of sides of f.
Figure 1.31 shows the duality construction on a torus. The primal graph has
two vertices and three edges between them, and there is only one primal
region. The dual graph has only one vertex, at which three loops are based.
It should be noted that if one were to construct the dual of the dual
imbedding, then the primal imbedding of the primal graph would be restored.
This fact is what justifies the use of the word "dual".
1.4.9. Exercises
1. Cut out a rectangular strip of paper about 3 em wide and about 30 em
long. On both sides, draw a lengthwise center line. Next, paste the strip
into a Mobius band. Then cut the strip along the center line. How many
pieces are obtained? Next, cut each of the bands obtained in the previous
step along their center lines. What happens?
2. Draw K 5 in the sphere so that there is only one edge-crossing. Do the
same for K 3 , 3 •
3. Draw an imbedding of K 6 on the torus.
4. Draw K 6 on the sphere with only three edge-crossings.
More Graph-Theoretic Background 33
5. Prove that no simplicial graph with seven vertices and 16 edges can be
imbedded in the sphere.
6. Suppose that a simplicial graph has p vertices. What is the maximum
number of edges it can have and still have an imbedding in the sphere?
7. Suppose that a simplicial bipartite graph has p vertices. What is the
maximum number of edges it can have and still have an imbedding in the
sphere?
8. Prove that there are exactly five different simplicial graphs with six
vertices and 12 edges. Decide which ones can be imbedded in the sphere.
(Give proofs, of course.)
9. Draw three cellular imbeddings of different graphs in the torus, and then
calculate the Euler formula in each case. What number is its value? This
number is the Euler characteristic of the torus.
10. Find a subgraph of the Petersen graph (recall Figure 1.20) that is a
subdivision of K 3 3 , thereby proving that the Petersen graph is non-
spherical. '
11. Using the girth of the Petersen graph, prove it is nonspherical.
12. Draw the dual for the imbedding of Figure 1.25.
13. Draw 2-cell imbeddings for K 5 - e (i.e., the result of removing one edge
from K 5 ) in the sphere and in the torus. Then draw their duals.
14. Show that the 3-cube graph Q 3 is isomorphic to a dual of the 3-octahedron
graph 0 3 in the sphere.
15. Prove that there are exactly four isomorphism types of connected graphs
with three vertices and three edges, and that exactly two of them have
triangular imbeddings in the sphere.
16. Let G be a graph. The "line graph" L(G) has vertex set Ec, the edges of
G. Two members of Ec are adjacent in Lc if and only if they have a
common endpoint in G. Give an example of a graph G such that
y(L( G))> y( G)
Also, prove that if the maximum valence of a graph G is less than or
equal to 3, then
y(L(G)) 5; y(G)
This section gives some explicit attention to a few of the standard topics in
graph theory that arise most frequently in topological developments.
34 Introduction
1.5.1. Traversability
In the earliest known paper on graph theory, Euler (1736) proved that it was
impossible to cross each of the seven bridges of Konigsberg once and only
once on a walk through the town. Figure 1.32 shows the original Konigsberg
problem, with two land areas on the opposite sides of the Pregel River and two
islands in the river, and also its graph-theoretic abstraction, in which the four
land areas are represented by vertices and the seven bridges by edges.
From the Konigsberg bridge problem arises the definition that a walk in a
connected graph is "eulerian" if it traverses each edge exactly once. A closed
eulerian walk is called an "eulerian circuit". A graph containing an eulerian
circuit is called an "eulerian graph".
Theorem 1.5.1. A connected graph G is eulerian if and only if every vertex has
even valence.
Proof First suppose that the closed walk W is an eulerian circuit in G that
starts and stops at the vertex v. Then each occurrence in W of a vertex of G,
other than v, contributes an addend of 2 to the valence of that vertex. It
follows that the valence of any vertex of G except v equals twice the number
of times it occurs on W. The valence of v is 2 for the beginning and end of W
plus twice the number of times that v occurs elsewhere in W.
Conversely, suppose that every vertex of the graph G has even valence. By
way of induction, assume that G is the smallest such graph that has no
eulerian circuit. The graph G is not K 1 , which is eulerian. Moreover, G has no
vertices of valence 1. Thus, from Theorem 1.2.3, the graph G is not a tree; so it
must contain a cycle C. Since every vertex of the graph G - C obtained by
removing the edges of the cycle C from the graph G has even valence, and
since each component of G - C is smaller than G, it follows that each
component of G - C has an eulerian circuit.
Let G1 , •.• , Gn be the components of G- C, indexed in the order in which
they are encountered on a traversal of the cycle C. An eulerian circuit for the
graph G is constructed as follows. Start traversing the cycle C. As soon as the
More Graph-Theoretic Background 35
Theorem 1.5.1 does not quite solve the Konigsberg bridge problem, as
originally stated. Making the necessary modifications to solve that problem is
left as an exercise at the end of this section.
A more difficult problem than finding an eulerian circuit has grown out of a
puzzle invented by Sir William Hamilton (see Biggs et al., 1976). The
"dodecahedron" is the unique 12-sided regular three-dimensional polyhedron.
Each of its sides is a pentagon. The !-skeleton of this polyhedron, illustrated
in Figure 1.33, is called the "dodecahedron graph". From our perspective,
Hamilton's puzzle was to find a cycle containing every vertex of the dodeca-
hedron graph.
In general, a cycle that contains every vertex of a graph is called a
"hamiltonian cycle". A graph that has a hamiltonian cycle is called a "ham-
iltonian graph". Whereas there are very fast ways to construct an eulerian
circuit in a given graph or to determine that none exists, it is usually much
more difficult to construct a hamiltonian cycle or to determine that none exists
(see Karp, 1972).
1.5.2. Factors
The edge set of a spanning subgraph for a graph G is called an "n-factor" for
G if that spanning subgraph is regular of valence n. If the edges of a graph G
can be partitioned into n-factors for G, then G is called "n-factorable". For
instance, Figure 1.34 shows a !-factorization of the complete graph K 4 and a
2-factorization of the complete graph K 5 •
In general, the edges of a 2-factor form a family of mutually disjoint cycles
that includes every vertex. Petersen (1891) observed that every 4-regular graph
is 2-factorable, as demonstrated by the following argument. Since every vertex
36 Introduction
chromatic number of a graph, one need consider only the chromatic numbers
of simplicial graphs imbeddable in S. By the Ringel-Youngs (1968) solution
to the Heawood map-coloring problem,
chr () l
sg =
7 + f1 + 48g
2
J for every genus g ~ 1
One may observe that for g = 0, the value 3 of the right-hand side of this
equation is 4. The equation chr( S0 ) = 4 was verified separately by Appel and
Haken (1976), who thereby solved the four-color problem. Map-coloring
problems are discussed at length in Chapter 5.
1.5.6. Algorithms
There are numerous known algorithms to decide whether a graph is spherical,
and there are several to calculate the genus. A naive sphericity algorithm is
obtained from Kuratowski's theorem, with the aid of a definition that permits
modular algorithm structure.
Let u be a vertex of valence 2 in a graph G, and let e 1 and e 2 be the edges
incident on u, with V(e 1 ) = {u, v} and V(e 2 ) = {u, w}. The graph obtained
by "smoothing at u" has for its vertex set VG - { u} and for its edge set
EG- { e 1 , e 2 } plus a new edge between v and w. Subdividing this new edge
would invert the operation of smoothing at u.
Since there are 2#£ subsets of edges to consider, an execution of this naive
algorithm could be quite lengthy. In Section 1.6, we derive a polynomial-time
algorithm for sphericity. A general (exponential-time) algorithm for genus is
given in Chapter 3.
1.5.7. Connectivity
A "cutpoint" of a graph G is a vertex whose removal would increase the
number of components, where it is understood that removing a vertex from G
means also removing every edge incident on that vertex, so that the result of
the operation is a graph. A maximal subgraph for G with no cutpoints is
called a "block".
The "connectivity" (G) of a graph G is defined to be the minimum number
of points whose removal results in a disconnected graph, a trivial graph, or a
bouquet of circles. It may be observed that G has the same connectivity as
csimp.
1.5.8. Exercises
1. Prove that a connected graph G has an open eulerian walk if and only if
exactly two of the vertices have valence 1.
2. Use Theorem 1.5.1 and Exercise 1 to solve the Konigsberg bridge
problem.
3. Draw a hamiltonian circuit in the dodecahedron graph.
4. How many vertices and edges does the dodecahedron graph have?
Restriction on method: Do not count the vertices or the edges. Using the
fact that a dodecahedron is a 12-sided polyhedron, each side a pentagon,
it is possible to calculate #V and #E.
5. The "icosahedron" is the unique 20-sided regular three-dimensional
polyhedron. Each side is a triangle. The "icosahedron graph" is its
!-skeleton. It is known that the icosahedron is dual to the dodecahedron.
Draw the icosahedron graph as the dual to a spherical imbedding of the
dodecahedron graph. How many vertices and edges does the icosahedron
graph have? Restriction on method: Do not count the vertices or the
edges of the icosahedron graph.
6. Prove that the Petersen graph is not !-factorable.
7. Find a 2-factorization of the complete graph K 9 •
8. Let G and H be vertex-transitive graphs such that the neighborhood of
any vertex of G is isomorphic to the neighborhood of any vertex in H.
Are G and H necessarily isomorphic graphs? Explain.
9. Prove that any imbedding of a wheel in the sphere has for its dual graph
an isomorphic wheel.
10. What is the chromatic number of Petersen's graph?
11. Draw a map on the Klein bottle that requires six colors.
12. Draw a map on the torus that requires seven colors.
13. The Haj6s conjecture is that every n-chromatic graph contains a subdivi-
sion of Kn. Prove the Haj6s conjecture for n = 1, 2, and 3. (The Haj6s
conjecture is false for nearly all graphs. See Catlin (1979) and Erdos and
Fajtlowicz (1981).
14. Prove the Haj6s conjecture for n = 4.
15. Prove that a contraction of a tree is a tree.
16. Prove the Hadwiger conjecture for n = 2.
17. Give an example to show that a contraction can increase the chromatic
number of a graph.
18. Let e 1 and e 2 be proper edges with a common endpoint w in a graph G,
let u be the other endpoint of e 1 , and let v be the other endpoint of e 2 •
Let G' denote the graph obtained from G by deleting edges e 1 and e 2
and then adding a loop at the vertex w and a new edge between the
42 Introduction
vertices u and v (i.e., another new loop if u = v). Use Petersen's theorem
for 4-regular graphs to prove that if G' is 2-factorable, then so is G.
19. Using Exercise 18 and an induction on the number of proper edges,
prove Theorem 1.5.2.
1.6. PLANARITY
The study of planar imbeddings has a long and rich history that intertwines
with chromatic graph theory, algorithmic analysis, enumeration, and much
else. However, unlike higher-genus surfaces, the plane and sphere are homo-
logically trivial, which is to say, overly simple in a topological sense. Accord-
ingly, the methods used to study planar imbeddings are less topological than
the ones of primary interest to us here. Nonetheless, Kuratowski's theorem
might well be the most famous theorem in all of graph theory, and it might be
remiss to offer no proof of it.
Of the many published proofs, we have selected one whose organization is
due to Thomassen (1980). It is not the shortest, and it is not the easiest.
However, it is illuminating, in that it also yields other results about planar
imbeddings. One is Fary's theorem that the edges of a planar imbedding of a
simplicial graph can be chosen to be straight lines. Another is Tutte's theorem
that the regions of a planar, 3-connected graph imbedding can be chosen to be
convex. Still another is Whitney's theorem that a 3-connected simplicial planar
graph has only one planar imbedding. Moreover, Thomassen's proof is readily
converted into a planarity algorithm that is polynomial in the number of
vertices, a great improvement over the exponential Naive Algorithm given in
Section 1.5.
Basis step. The statement (KT) is vacuously true of all graphs with four or
fewer vertices.
Induction Hypothesis. We assume that the statement (KT) is true of all graphs
with fewer than n vertices, where n is any number greater than or equal to 5.
Let e be any edge in the graph G, and suppose that its endpoints are the
vertices u and v. Let G' be the graph obtained from G by a simplicial
contraction of edge e, and let us give the name v' to the vertex to which e is
contracted. We momentarily interrupt the proof of Kuratowski's theorem to
establish a lemma.
X y
Figure 1.37. The subgraph H obtained by reversing the contraction of edge e.
44 Introduction
Proof Suppose that the boundary of some region r is not a cycle, but
some other kind of closed walk, so that the vertex v occurs twice, as on the left
of Figure 1.40. Then there is a simple closed path in the plane that leaves from
vertex v between two edges of the boundary of r, stays the whole time within
the region r, and later comes back to vertex v between a different pair of
edges, as shown on the right-hand side of Figure 1.40. This closed path
separates the plane into two pieces, both of which contain parts of the graph
G. Since the path intersects the graph G only at vertex v, it follows that v is a
cutpoint of the graph G. Therefore, G is not 2-connected.
Conversely, suppose that G has a cutpoint. Then G may be viewed as the
amalgamation of two graphs H and K at the vertex v. In any imbedding of G,
46 Introduction
r
Figure 1.40. The region r (shaded) and vertex v.
v v'
Figure 1.41. An edge contraction of a 3-connected graph to a 2-connected graph.
Theorem 1.6.2. Let G be a 3-connected graph with five or more vertices. Then
there is some edge e of G such that the graph G/ e obtained by contracting e is
also 3-connected.
then one component of H' - y would contain both the vertices u and v, since
u is adjacent to v; and, therefore, all the other components of H' - y are
connected to the rest of the graph G through the vertices y and w. This would
imply that { y, w} disconnects G, contradicting the 3-connectivity of G. We
conclude that for some edge e, the contracted graph Gje is 3-connected. 0
lbeorem 1.6.4 (Tutte, 1960). Any planar 3-connected graph has a planar
imbedding such that every bounded region is convex.
lbeorem 1.6.5 (Fary, 1948). Any planar 3-connected graph has a planar
imbedding such that every edge is a straight line segment.
It is obvious from Theorem 1.6.1 that Tutte's theorem does not hold for
!-connected planar graphs. Moreover, it does not hold even for 2-connected
planar graphs (Exercise 1.6.2). On the other hand, Fary's theorem holds for all
planar graphs, as we shall show later in this section.
Planarity 49
There is another observation worth making about the proof for Kuratowski's
theorem in the 3-connected case. When the vertex v' is split back into vertices
u and v, there really is no choice of how the resulting imbedding looks at u
and v. Thus, there is essentially only one way to imbed G in the plane. By
"essentially one way", we mean up to a homeomorphism of pairs. Further
discussion of this concept is deferred to Chapter 3. Anyone familiar both with
Thomassen's proof and that concept of uniqueness will recognize that
Thomassen's proof of Kuratowski's theorem can be adapted to prove the
following result of Whitney, by once again using an induction based on K 4 •
Theorem 1.6.6 (Whitney, 1933). There is only one way to imbed a 3-connected
planar graph in the plane.
Proof Omitted. D
Theorem 1.6.7. Let H and K be planar graphs. Then the graph obtained by
amalgamating H and K either at a single vertex v or along a single edge e is
planar.
Theorem 1.6.7 now leads directly to a proof of Kuratowski's theorem for all
graphs, 3-connected or not, starting from Theorem 1.6.3, the version for
3-connected graphs. However, we should like to prove Fary's theorem as well,
and the proof of Theorem 1.6.7 wreaks havoc with straight line segments.
Accordingly, we take a different tack, as follows. If a graph contains no
homeomorph of K 5 or K 3 3 , and it is not 3-connected, then we add as many
edges to the graph as possible without creating K 5 or K 3 3 • If the extra edges
make the resulting graph 3-connected, then Kuratowski's theorem and Fary's
theorem will hold for the original graph, because it is a subgraph of a
3-connected graph. The following theorem is what is needed.
of the original graph G, since the amalgamation of H' and K' along the edge e
is simply G with the edge e added.
We therefore assume that G is the amalgamation of graphs Hand K along
the edge e and that Hand K share the same maximality property as G with
respect to homeomorphs of K 5 and K 3, 3 • By the induction hypothesis, the
graphs H and K are 3-connected, and hence planar. By Theorem 1.6.7, the
graph G is planar. It follows that in any planar imbedding of G, H, or K,
there can be no region with four or more sides. Otherwise a "diagonal" edge
across the region could be added, creating a planar graph with one more edge,
contradicting the maximality of the given graph (there is a slight problem if
the vertices we want to join by a diagonal are already joined by an edge
elsewhere in the imbedding; see Exercises 3 and 4). Since the graphs in
question have no loops or multiple edges, every region must have three sides.
By Euler's equation it follows that #E = 3#V- 6 for the edge and vertex
sets of any of the graphs G, H, and K (see Exercise 5). But since G is the
amalgamation of Hand K along a single edge, #V(G) = #V(H) + #V(K)
- 2 and #E(G) = #E(H) + #E(K)- 1. Therefore
Corollary. Kuratowski's theorem and Fary's theorem hold for graphs that are
not 3-connected.
1.6.6. Algorithms
The number of steps required by the Naive Sphericity (or Planarity) Algorithm
suggested in Section 1.5 is an exponential function of the number n of vertices
of a graph G. We shall see that Thomassen's design for a proof of the
Kuratowski theorem leads to an algorithm whose execution time is a poly-
nomial function of n, thereby a great improvement.
We first assume that the graph G to be tested is 3-connected. In that case,
according to Theorem 1.6.2, there exists an edge e such that contraction of G
on e yields a 3-connected graph. This step is iterated on that new 3-connected
graph, and reiterated until a 4-vertex graph is obtained.
Execution time analysis: It takes n - 4 edge contractions to convert an
n-vertex graph into a 4-vertex graph. Let us suppose that prior to each edge
contraction, we find it necessary to make an exhaustive search over all edges
52 Introduction
In other words, the number of steps needed for the edge-contracting part of
the algorithm is at most quartic in the number of vertices.
Of course, since the complete graph K4 is the only 3-connected 4-vertex
graph, it must be the graph we ultimately obtain from the sequence of
contractions. At this point, we imbed K 4 in the plane, and iteratively reverse
each of the contractions, either until some reversal yields a copy of a
Kuratowski graph, or until we have a planar imbedding of the original graph G
in the plane.
At each reverse contraction, when a vertex v' is split back into two vertices
u and v, the cycle C of vertices adjacent to v' is checked for the following
patterns:
1.6.9. Exercises
1. Give an imbedding of a 2-connected graph in the torus such that some
region has a boundary that is not a cycle.
2. Find a planar 2-connected graph that has no planar imbedding with
every bounded region convex.
3. Prove that in any planar imbedding of the complete graph K 4 , every
region is 3-sided. Conclude that if a simple graph G has a planar
imbedding with four or more vertices all lying on the same region
boundary, then an edge can be added to G (without creating multiple
edges) and still maintain planarity.
Planarity 55
4. Use Exercise 3 to prove that if G is planar and the addition of any edge
to G creates a nonplanar graph, then every region in any planar imbed-
ding of G has size 3. (Caution: The number of distinct vertices in a
region boundary can be less than the number of sides of the region.)
5. Prove, using Euler's equation, that any planar imbedding for a connected
simplicial graph with edge set E such that #E ~ 2 and vertex set V
satisfies #E ::::;; 3#V- 6, with equality if and only if every region is
3-sided.
6. Given a graph G with vertex set V = { v1 , ••. , vn} and edge set E, define
the linear transformation d: S(E) ~ S(V) by d(e) = u + v, where u
and v are the endpoints of the edge e. Show that the cycle space C(E) is
the kernel (nullspace) of d, and that if P is a path of edges from vertex u
to vertex v, then d(P) = u + v. Show that if G is connected, then
v1 + v2 , v1 + v3 , ••• , v1 + vn are all in the range of d, but v1 is not.
Conclude that if G is connected, the range of d has dimension n - 1 and
therefore the dimension of the cycle space C(E) is #E- #V + 1.
7. Prove that K 3 3 has no 2-basis. (Hint: The dimension of C(K3 3 ) is 4, by
Exercise 6. Show that at least seven edges of K 3 3 must be each' contained
in two elements of a 2-basis for K 3, 3 and the~ consider the sum of all
elements of the 2-basis.)
8. Prove that K 5 has no 2-basis (see Exercise 7).
9. Prove that if the graph G has a 2-basis, then so does G - e where e is
any edge of G.
10. Use Exercises 7-9 and Kuratowski's theorem to prove MacLane's theo-
rem.
11. Given a 2-connected graph G with edge set E, let the bond space B(E)
be the subspace of S(E) spanned by the bonds of G. Let star(v) be the
subgraph induced by collection of all edges incident to the vertex v.
Prove that { E(star(V)) 1 v E V} is a spanning set for the bond space
B(E). [If A c E and G- A has two components H and K, consider
l: E(star(V)) where the sum is taken over all vertices of H.]
12. Use Exercise 11 and MacLane's theorem to prove Whitney's theorem.
13. Find an algorithm that determines in at most kn steps, k a constant,
whether a given n-vertex graph G with at most 3n edges is connected.
(Assume you are given for each vertex a list of the adjacent vertices.)
2
Voltage Graphs and
Covering Spaces
56
Ordinary Voltages 57
Let G be a graph whose edges have all been given plus and minus directions,
and let .5!1 be a group, assumed to be finite unless explicitly stated to be
infinite. A set function a from the plus-directed edges of G into the group .5!1
is called an "ordinary voltage assignment" on G, and the pair (G, a) is called
an "ordinary voltage graph". The values of a are called "voltages", and .5!1 is
called the "voltage group". The ordinary voltage-graph construction was first
suggested by Gross (1974) and immediately improved by Gross and Tucker
(1974). Its advantage over various formalistic "covering graph" constructions,
all essentially equivalent, is largely its visual suggestiveness.
uo d2
u2
·~ (a)
el
ul
(b)
Figure 2.1. (a) A graph with voltages assigned m the cychc group ~3 ((G, a)) and (b) the
as;ociated derived graph (Gn).
58 Voltage Graphs and Covering Spaces
the directed edge ea + of the derived graph Go. runs from the vertex ua to the
vertex vab·
Remark. In the definition of the derived graph, note that whereas ea + runs
from ua to vab' the reverse edge ea- runs from vab to ua. In particular, the
subscript on the minus-directed edge e a- is the product of the subscript on its
initial vertex vab and the group element b- 1 that is inverse of the voltage on e+.
Thus, whereas the subscript of a plus-directed edge of the derived graph always
agrees with the subscript on its initial vertex, the subscript on a minus-directed
edge always agrees with the subscript on its terminal vertex.
Example 2.1.1. Since the base graph in Figure 2.1 has two vertices, u and v,
and two edges, d and e, and since the voltage group fl'3 has three elements, 0, 1,
and 2, it follows that the derived graph has six vertices, u0 , u1 , u2 , v0 , v1 , and
v2 , andsixedges, d 0 , di> d 2 , e0 , e 1 , ande 2 • Sincetheu-basedloopd+ ofthe
base graph G is assigned voltage 1, it follows that for i = 0, 1, 2, the edge d; + of
the derived graph G runs from the vertex u; to the vertex u; + 1 • Since the edge e +
is assigned the voltage 0, it follows that fori = 0, 1, 2, the edge e; + of the derived
graph runs from the vertex u; to the vertex V;.
Example 2.1.2. Figure 2.2 shows how the Petersen graph might be derived by
an assignment of voltages in the cyclic group fl'5 to a particular base graph with
two vertices and three edges, which is sometimes called a "dumbbell" graph. As
indicated, the directions on the edges of the derived graph are induced by the
directions of the corresponding edges in the base graph.
1 c
0--7-D o
d
u
e 2
(a)
c2
Figure 2.2. A derivation (b) of the Petersen graph from a voltage assignment in .2'5 to a
dumbbell graph (a).
Ordinary Voltages 59
Since the dumbbell graph in Figure 2.2 has two vertices, u and v, and since
the voltage group is f!E5 , the vertices of the derived graph are u 0 , u1 , u 2 , u 3 ,
v4 , v0 , v1 , v2 , v3 , and v4 • Since the dumbbell graph has edges c, d, and e, the
derived graph has edges c0 , c1 , c 2 , c3 , c4 , d 0 , d1 , d 2 , d 3 , d 4 , e 0 , e 1 , e 2 , e 3 ,
and e 4 • The u-based loop c+ has voltage 1 in the base graph. Thus, for
i E f!E5 , the edge c/ runs from the vertex u; to the vertex u; +1 , as shown on
the outer pentagon of the derived Petersen graph. The v-based loop e+ in the
base graph has voltage 2. Thus, for i E f!E5 , the edge e; + runs from the vertex
v; to the vertex v; + 2 , as shown on the inner star. The proper edge d+ has
initial vertex u and terminal vertex v in the dumbbell and is assigned voltage
0. Thus, for i E f!E5 , the edge d; + in the derived graph runs from the vertex u,
to the vertex v;.
If one uses a different cyclic group !!En as the voltage group and assigns
voltages 1 and k to the loops c and e, then one obtains the generalized
Petersen graph G(n, k) (Frucht et al., 1971).
Example 2.1.3. Figure 2.3 shows a 3-runged "Mobius ladder" as a base graph,
with voltages in f!E2 • Since this base has six vertices and nine edges, and since the
voltage group has order 2, the derived "circular ladder" C6 X P2 has 12 vertices
and 18 edges.
To avoid cluttering the diagram, the names of the edges have been omitted
from Figure 3.3. This does not affect the construction of the derived graph.
Quite often, the genus of the derived graph is higher than the genus of the
base graph, as illustrated by Example 2.1.2. However, the genus of the derived
graph need not be larger. Indeed, as Example 2.1.3 proves, it may be smaller
than the genus of the base graph, for we recognize that the 3-runged Mobius
ladder is isomorphic to the Kuratowski graph K 3 3 , whereas the 6-runged
circular ladder is visibly planar. '
0
Figure 2.3. A circular ladder graph derived by a voltage assignment in .2"2 to a 3-runged Mobius
ladder.
60 Voltage Graphs and Covering Spaces
Figure 2.4. The 3-octahedron graph 0 3 derived by assigning voltages in .2'6 to the bouquet 8 2 .
Example 2.1.4. The fr6-voltage 1 on the loop d of the base bouquet in Figure 2.4
has order 6, and the edge fiber over d forms one 6-cyc/e. The fr6-voltage 2 on the
loop e of the bouquet has order 3, and the edge fiber over e forms two 3-cyc/es.
Since only the edge e c + of the fiber over the loop e + starts at the vertex vc of
Ordinary Voltages 61
the fiber over the vertex v, only one of these n-cycles passes through vc, which
implies that these n-cycles are mutually disjoint.
The "natural projection" from the derived graph to the base graph is the
graph map p: Go.~ G that maps every vertex in the fiber over v to the vertex
v, for all v E Vc, and every edge in the fiber over e to the edge e, for all
e E Ec. In terms of the notation, the natural projection wipes out the
subscripts.
order and direction of that walk. For instance, in Example 2.1.2, the net
voltage on the walk e+, d-, c-, d+ is
in the derived graph Go. such that for i = 1, ... , n the edge e; is in the fiber
over the edge e; .
Example 2.1.5. In Figure 2.5, the voltage group is the cyclic group :Z3 • There
are three lifts of the walk c-, e+ from the vertex u to the vertex v of the base
graph, namely, the following walks of the derived graph:
One important observation about Example 2.1.5 is that although the walk
c-, e- of the base graph includes the same edges as the walk c-, e + and
although it has the same endpoints, its lifts are different. The lifts of the walk
c-, e are
and c0 -,e 2 -.
Continuity is the reason that the lift of the walk c-, e+ starting at the vertex
62 Voltage Graphs and Covering Spaces
Figure 2.5. Another voltage graph and its derived graph. The voltage group is .2"3 •
U; in the derived graph is the walk c;+ 1 -, e;+ 1 +, whereas the lift of c-, e-
starting at u; is c;+ 1 -, e; -. Since the voltage on the directed edge c- is 1, the
next vertex on either lifted walk is V;+ 1 • When e+ is the next edge of the base
walk, the next lifted edge must be the unique edge in the fiber over e whose
plus direction originates at v,+ 1 (i.e., e;+ 1 ). When e- is the next edge of the
base walk, the next lifted edge must be the unique edge in the fiber over e
whose minus direction originates at V;+ 1 (i.e., e,). This is one reason that in
topological graph theory, the direction in which one traverses a loop cannot be
ignored.
Theorein 2.1.1. Let W be a walk in an ordinary voltage graph (G, a) such that
the initial vertex of W is u. Then for each vertex ua in the fiber over u, there is a
unique lift of W that starts at ua.
Proof Consider the first directed edge of the walk W, say e + or e-. If it is
e+, then, since only one plus-directed edge of the fiber over e starts at the
vertex ua (i.e., the edge ea +), that edge must be the first edge in the lift of W
starting at u a. If it is e-, then, since only one minus-directed edge of the fiber
over e starts at ua (i.e., the edge eaa(e-> -),it follows that the edge must be the
first edge in the lift of W starting at ua. Similarly, there is only one possible
choice for a second edge of the lift of W, since the initial point of that second
edge must be the terminal point of the first edge, and since that second edge of
Ordinary Voltages 63
the lift must lie in the fiber over the second edge of the base walk W. This
uniqueness holds, of course, for all the remaining edges as well. D
Example 2.1.5 Revisited. The net voltage on the walk W = c-, e+ from u to v
is (-2) + 1 = 1 + 1 = 2. The walk W0 (= c1 -, e 1 +)ends at v0 + 2 = v2 , the
walk W 1 ( = c 2 -, e 2 +)ends at v 1 + 2 = v0 , and the walk W2 ( = c0 -, e 0 -) ends at
v2+2 = V1.
u0 (e5, 0)
Figure 2.6. The 3-cube graph Q3 derived by assigning .2'2-voltages to the complete graph K 4 •
the #.9.1/m left cosets of the cyclic group generated by the net voltage b, there
is a unique component of p - 1( C). 0
Example 2.1.6. Figure 2.6 shows how the 3-cube graph Q3 may be derived
using ~2-voltages on the complete graph K 4 • One observes that the sum (i.e., the
product operation for the group ~2 ) of the voltages on the 4-cycle e 1 , e 2 , e 3 , e 4
is 0 modulo 2, which has order 1, and that the preimage of that 4-cycle is the
union of the two 4-cycles ( e 1 , 0), ( e 2 , 1), ( e 3 , 0), ( e 4 , 1) and ( e 1 , 1),
(e 2 ,0),(e 3 , l),(e 4 ,0). On the other hand, one notes that the 3-cycle e 1 , e 2 , e 5 ,
in the base graph has voltage sum 1 modulo 2, which has order 2, and thus the
preimage in the derived graph of that 3-cycle is the 6-cycle ( e 1 , 0),
( e 2 , 1), ( e 5 , 0), ( e 1 , 1), ( e 2 , 0), ( e 5 , 1).
Example 2.1.5 Revisited Again. The net voltage on the 2-cycle c+, d+ in Figure
2.5 is 1 modulo 3. According to Theorem 2.1.3, the preimage of that 2-cycle
should be one 6-cycle. In fact, it is the 6-cycle c0 +, d 2 +, c 1 +, d 0 +, c 2 +, d 1 +.
Also, the net voltage on the !-cycle e- is 2 modulo 3, and its preimage should be
one 3-cycle. In fact, it is the 3-cycle e 2 -, e 1 -, e 0 -.
2.1.6. Exercises
Figure 2.7 shows four voltage graphs, whose derived graphs are to be con-
structed in Exercises 1-4.
1. Draw the derived graph corresponding to the voltage graph in Figure
2.7a.
2. Draw the derived graph corresponding to the voltage graph in Figure
2.7b.
3. Draw the derived graph corresponding to the voltage graph in Figure
2.7c.
Ordinary Voltages 65
0 u
0
1 0 0 1
u 0 w
(a)
(b)
(1,0,0)
1CD3
(c) (d)
Figure 2.7. Four voltage graphs. (a) In .2'3 , (b) in .2'2 , (c) in .2'2 X .2'2 X .2'2 , (d) in .2'5 .
(134) u
"4 (123)
ll (1 3)(2 4) ill
(a) (h)
We have seen in Section 2.1 that the Petersen graph, the octahedral graph 0 3 ,
the 3-cube graph Q 3 , and several other graphs can all be obtained as the
derived graph of an ordinary voltage-graph construction. The unifying char-
acteristic of these derivable graphs is a certain kind of global symmetry,
precisely described as the existence of a group of automorphisms that acts
freely. Immediately below, we prove that a given graph is derivable with
ordinary voltages from a proposed base graph if and only if the proposed base
graph is a regular quotient of the given graph. This facilitates proof that some
graphs, such as trees, are derivable only from themselves, whereas other
graphs, such as Cayley graphs, are derivable from small base graphs.
Theorem 2.2.1. The vertex orbits of the natural action of the voltage group on
an ordinary derived graph are the vertex fibers, and the edge orbits are the edge
fibers.
Proof For any vertex va of the derived graph Ga, the orbit of va is defined
to be the set { 1/>c( va) I c E A}, which is the same as the set { vca IcE A} =
Which Graphs Are Derivable with Ordinary Voltages? 67
E A}. That is, the orbit of va is the fiber over v. Similarly, the orbit of
{ vc 1 c
the edge e a of the derived graph is the fiber over the edge e. D
Corollary. The natural projection p: Go.~ G of the derived graph onto the base
graph is a covering projection.
Proof The composition of the quotient map qA: Go.~ Ga/A with the
graph isomorphism Go./A ~ G, which takes the fiber over a vertex or edge to
that vertex or edge, is equal to the natural projection p: co. ~ G. D
Example 2.2.1. The only nontrivial automorphism of the n-path Pn fixes the
central vertex if n is odd, or the central edge if n is even. Thus then-path has no
regular quotient except itself.
Theorem 2.2.2 (Gross and Tucker, 1974). Let .91 be a group acting freely on a
graph G, and let G be the resulting quotient graph. Then there is an assignment a
68 Voltage Graphs and Covering Spaces
of ordinary voltages in .5!1 to the quotient graph C and a labeling of the vertices of
G by the elements of VG X .5!1 such that G = ca and that the given action of .5!1
on 6 is the natural left action of .5!1 on ca.
Proof First, choose positive directions for the edges of the graphs C and
G so that the quotient map q~: G -+ C is direction-preserving and that the
action of .5!1 on(; preserves directions. Next, for each vertex v of C, label one
vertex of the orbit q,.,- 1(v) in G as v1 , and for every group element a-:/= 1,
label the vertex cf>A v1 ) as va. Now one may label edges. If the edge e of C
runs from v to w, one assigns the label ea to the edge of the oribt a_,.,- 1(e)
that originates at the vertex va. Since the group .5!1 acts freely on C, there are
#.Jil edges in the orbit q_,.,- 1( e), one originating at each of the #.Jil vertices in
the vertex orbit q_,.,- 1( v ). Thus, the choice of an edge to be labeled e a is
unique. Finally, if the terminal vertex of the edge e 1 is wb, one assigns voltage
b to the edge e of the quotient graph C; that is, a(e+) =b. To show that this
labeling of edges in the orbit q~- 1( e) and the choice of a voltage b for the
edge e really yields an isomorphism G-+ ca, one must show for all a E .5!1
that the edge ea terminates at the vertex w ab· However, since ea = cf>Ae1), the
terminal vertex of the edge ea must be the terminal vertex of the edge cf>a(e 1),
which is
As a result of Theorem 2.2.2, one knows that a given graph can be derived
from a smaller graph using ordinary voltages if and only if there is a nontrivial
free action on the given graph. To decide whether such an action exists, one
observes that the vertex orbits and the edge orbits under such a free action
would have to be of the same size as the group. Equivalently, in a derived
graph, all the vertex fibers and edge fibers have the same cardinality as the
voltage group. Thus, the order of a possible voltage group to derive a given
graph C must be a common divisor of #VG and #EG. However, this
common-divisor criterion is not always sufficient.
Example 2.2.3. A tree cannot be derived from voltages on some smaller graph,
because the number of vertices and the number of edges of a tree are relatively
prime. A more difficult proof of the fact that trees have no nontrivial regular
quotients might be obtained using Exercise 15 of Section 1.3.
Example 2.2.4. Since the 3-cube graph Q 3 has eight vertices and 12 edges, it
might have one or more regular quotients with four vertices and six edges or with
Which Graphs Are Derivable with Ordinary Voltages? 69
two vertices and three edges, but there are no other possibilities. See Exercise 5
of Section 2.1.
Example 2.2.2 Revisited. The complete graph K 4 has four vertices and six
edges. Although the number 2 is a common divisor of 4 and 6, there is no
quotient of K 4 with two vertices and three edges, because K 4 has no regular
quotients itself, as proved in the original discussion of Example 2.2.2.
Remark. One additional fact that may be used to find the quotients of a graph is
that all vertices in a fiber have the same valence, because the voltage group acts
transitively on the set of neighborhoods of all the vertices in the fiber over any
vertex of the base graph.
Proof Let v denote the vertex of the bouquet Bn and e 1 , ••• , en the edges.
Given-the group d and the generating set X, assign the generator x, as the
voltage on the loop e/, for i = 1, ... , n. Then the derived graph Bn" is
isomorphic to the Cayley color graph C(d, X) under the vertex map va-+ a
and the edge map (e;, a)-+ (x;, a). The converse is a special case of Theorem
2.2.2. D
One might recall that the alternative Cayley graph C(.s#, X) 1 is obtained
from the Cayley graph C(.J/1, X) 0 by deleting a 1-factor for each generator of
order 2.
Frucht's Theorem (1938). Let .91 be a finite group. Then there is a group G
such that Aut( G) is isomorphic to .Jil.
Which Graphs Are Derivable with Ordinary Voltages? 71
(a)
(b)
Figure 2.9. (a) A Cayley graph for the group .2"3 • (b) A graph whose automorphism group is .2"3 •
Proof Let X= { x 1 , ... , xn} be a generating set for the group .91. It is
immediately clear that every left translation <Pa on the Cayley color graph
C( .91, X) is color-consistent and direction-preserving. It is first to be proved
that every color-consistent, direction-preserving automorphism on C{s#, X) is
a left translation. For this purpose, suppose that the automorphism I is
color-consistent and direction-preserving, and suppose that the image 1(1~) of
the identity vertex is the vertex b. It is to be proved that I= q,b.
r
Since the graph automorphism q,b o 1 on C(s#, X) is color-consistent,
direction-preserving, and identity-preserving, Theorem 1.2.4 implies that its
vertex function coincides with a group homomorphism g: A ~A. Since
f/>b 0 r l is color-preserving on edges, it follows that g(X;) =X; for every
generator X; E X. Therefore, g is the identity group automorphism, for which
it follows that I= <f>b·
The next step is to modify the graph C(s#, X) 0 into a graph G so that the
features of color and direction from C(s#, X) are somehow present in G. To
do this, first insert two new vertices on each edge of C( .91, X) 0 • Then for
i = 1, ... , n and for each edge in C(s#, X) 0 corresponding to color x;, attach
a 2i-path at the new vertex near the terminal end and a (2i + I)-path at the
new vertex near the initial end. The automorphism group of the resulting
graph G is isomorphic to the group of left translations on C(s#, X), which is,
of course, isomorphic to the group .91. D
2.2.5. Exercises
1. Prove that the "wheel" K 1 + Cn has no nontrivial quotient.
2. Let G and H be graphs without cutpoints. Prove that the result of
amalgamating G and H at a single vertex has no nontrivial regular
quotients.
72 Voltage Graphs and Covering Spaces
Example 2.3.1. The main interest in Schreier graphs is for the case in which the
subgroup is not normal. Consider the subgroup ?A= {1, (2 3)} of the dihedral
group !5)3 , with generating set X= {(123),(12)}. Then S(!5)3 : ?A, X) has as
vertices the right cosets
?A= {1,(23)}
?A(12) = {(12),(123)}
?A(13) = {(13),(132)}
l I
'-""
(1 2 3) } edge dihedral group ~3
(12) - - - - colors subgroup !14= (1 ~ 3 • (2 3)}
Figure 2.10. (left) The Schreier color graph for Example 2.3.1. (right) The associated Schreier
graph.
Example 2.3.2. Let !!J be the subgroup {1Ei13, (2 3)} of the dihedral group P}3 •
Figure 2.11 shows a voltage assignment on a dumbbell graph in the group !i)3
relative to the subgroup !!J, and the resulting derived graph.
Example 2.3.3. Let !!J be a subgroup of a group .J/1, and let X = {x 1 , ... , x n }
be a generating set for .J/1. Assign directions to the edges of the bouquet Bn and
Irregular Covering Graphs 75
Z.\11(13) ZSlf(12)
U.Slf(12) Y.\11(12)
VSlf(12)
X .'.d(12) x.\lf
Y.Slt
z.\11
u.'.d V,'.d
u u
y
Figure 2.11. A relative voltage assignment on a dumbbell graph and the resulting derived graph.
assign relative voltages x 1 , ••• , xn to the respective plus-directed edges. Then the
resulting derived graph is isomorphic to the Schreier graphS(~: 114, X) 0 .
Given a relative voltage graph (G, ajl14) with voltages in the group ~.
fibers in the relative derived graph are defined as for the ordinary voltage-graph
construction. That is, for any vertex v of the base graph G, the "fiber" over v
is the vertex set { v} X ( ~: !14) in the derived graph. Similarly, for any edge e
of G, the "fiber" over e is the edge set { e} X (~: 114). The "natural
projection" of the relative derived graph onto the base graph is the graph map
that takes the fiber over every vertex onto that vertex and the fiber over every
edge onto that edge. If the subgroup 114 is not normal, then the relative derived
graph Gaf&W does not admit a natural fiber-consistent free group action, so that
the natural projection is not a regular covering. However, it may be called an
irregular covering, in a sense we can now make precise.
Example 2.3.5. The graph map f illustrated in Figure 2.12 that is defined by the
dropping of subscripts preserves valence, since all the vertices have valence 3.
However, the two edges d 1 and d 2 originating at u 1 are both mapped onto the
same edge d, originating at f(u 1 ) = u. Thus, f is not a covering projection,
because its restriction to edges originating at u 1 is not one-to-one. Moreover,
changing edge-directions on the graphs will not help at all.
u 1 ~--------...._ u2
el
Uj
u6 u2
u3
-- 0'~)
..........
u4
Figure 2.13. An irregular covering projection of the 4-regular graph (C3 U C4 )c onto the bou-
quet 8 2 •
78 Voltage Graphs and Covering Spaces
Proof Suppose that G is 2n-regular and that its vertices are v1 , ... , vr By
Theorem 1.5.2 (Petersen's theorem) the edges of G can be partitioned into n
2-factors, L 1 , ... , Ln. Assign an arbitrary orientation to each cycle of each
2-factor (thereby inducing plus directions on all the edges of G). If a 2-factor
L; is a hamiltonian cycle, then we could associate a cyclic permutation with it
exactly as in the preceding discussion of Figure 2.10. If the 2-factor L; has
several component cycles, cl' ... ' em, then with each of these edge cycles we
may associate a cyclic permutation corresponding to the order of the indices
encountered as one makes a traversal in the order of orientation, and the
permutation 'IT; to be associated with the 2-factor L; is the product of the
disjoint cycle permutations corresponding to its component cycles.
Let .91 be the group generated by the permutations 'IT1 , ••• , 'ITn correspond-
ing to the respective 2-factors L 1 , ••• , Ln, and let f1B be the subgroup of .91
consisting of all elements of .91 that fix the symbol 1. Then the graph G is
isomorphic to the Schreier graph S(.91: f!B, X) 0 , where X= { '171, ... , 'ITn}· To
construct the isomorphism, first assign each edge in the 2-factor L; the color
Irregular Covering Graphs 79
'IT,, for i = 1, ... , n. Then replace each vertex label vi by the right coset of !!4
that permutes 1 onto j, for j = 1, ... , p. The subgroup !!4 has such cosets
because G is connected. In particular, the product of the "signed colors" one
traverses on an edge path from v1 to vi is a permutation ~ such that /;(1) = j,
so that !14~ is the right coset that permutes 1 onto j. D
Corollary. Let B be a connected regular graph of odd valence such that G has a
!-factor. Then G is isomorphic to an alternative Schreier graph.
Unfortunately, not all regular graphs of odd valence have a 1-factor (see
Exercise 5). Biggs eta!. (1976) give a historical account of 1-factors, including
Tait's "theorem" that 2-connected, 3-valent graphs are !-factorable and
Petersen's counterexample, the graph that now bears his name. Akiyama and
Kano (1985) survey more recent results on graph factorization. Tutte (1947)
gives a necessary and sufficient condition for the existence of a 1-factor, from
which he obtains the following theorem.
Proof Omitted. D
2.3.5. Exercises
I. Let !!4 be the subgroup generated by the element (1 2 4) of the alternating
group .J/14 , and let X be the generating set {(1 2 3), (2 3 4)}. Draw the
Schreier graph S( .J/14 : !!4, X) 0 •
80 Voltage Graphs and Covering Spaces
(2 3)
u v
(1 2)
y
Figure 2.15. A permutation voltage assignment and the derived graph.
Proof This theorem is the analog of Theorem 2.1.1, and its proof is
exactly analogous to the proof of that theorem. D
Permutation Voltage Graphs 83
Example 2.4.1 Revisited. The walk W = y-, x-, y + in the base graph of
Figure 2.15 has three lifts, namely, the walks WI = YI -, x3 -, Y3 +; = w2
Y3 -, x2 -, Y2 +; and w3 = Y2 -, xl -, YI +.
Proof Let 'TT1 , ••• , .,, be the successive permutation voltages encountered
on a traversal of the walk W. Then the subscripts of the successive vertices of
the lift W, are
Proof Choose any vertex u of the cycle C and regard C as a walk from u
to itself. From Theorem 2.4.2, it follows that for i = 1, ... , n the lift of C
starting at u; terminates at u,<;>· Thus, the component of p- 1(C) that
84 Voltage Graphs and Covering Spaces
contains u; also contains uw(i)• uw2(i)• ••• , u.,,-'<i>• where j is the length of the
cyclic permutation containing the object i in the decomposition of 'IT. There-
fore, that component is a kj-cycle. The conclusion now follows readily. D
Theorem 2.4.5 (Gross and Tucker, 1977). Let the graph map q: G-+ G be a
covering projection. Then there is an assignment a of permutation voltages to the
base graph G such that the derived graph Ga is isomorphic to G.
Proof Choose a spanning tree Tin the base graph G, pick a root vertex u,
and assign plus directions to the edges of T so that it is possible to travel from
the root u to any other vertex of G entirely along plus-directed edges of T.
Assign plus directions to the other edges of G arbitrarily. Then assign
directions to the edges of G so that the map q is direction-preserving.
Suppose that the preimage q- 1(u) has n vertices. Label them u1 , ••• , un
arbitrarily. Then choose an edge e of T originating at u and terminating, say,
at v. By the definition of covering projection, the preimage q- 1(e) consists of
n edges, one originating at each of the vertices u1 , ••• , un. Moreover, exactly
one of these terminates at each vertex of q - 1( v ), so that there are also n
vertices in q- 1(v). Label the edge of q- 1(e) that originates at u, as e;, for
i = 1, ... , n. Also, label then vertices of q- 1(v) as v1 , ••• , vn in any way. If
we match q- 1(u) to q- 1(v), the edges of q- 1(e) define a permutation 'IT in the
symmetric group Y',. That is, if u; is matched to v1 then 'TT(i) = j. The
permutation voltage 'IT is assigned to the edge e of the base graph G. (If the
vertices v1 , ••• , vn are labeled so that for i = 1, ... , n, the edge e, terminates
at v;, then the voltage assigned to the edge e is the identity permutation.)
Permutation Voltage Graphs 85
Continue this procedure until all edges of T have been assigned permuta-
tion voltages, always selecting as the next edge of Tone whose initial point lies
on a path from the root vertex u whose every edge already has a permutation
voltage assigned. (By the parenthetic remark at the end of the preceding
paragraph, it is possible to assign the identity permutation to every edge of the
maximal tree T.) The last step is to assign permutation voltages to the edges of
G not in T. Suppose that d is such an edge, running from a vertex s to a
vertex t. Then q- 1(d) matches s 1 , ••• ,sn to t 1 , ••• ,tn, so we may assign
names dp ... , dn to the edges in q- 1(d) in agreement with the subscripts on
their initial points, and we may assign a permutation voltage to the edge d
according to the matching. D
Corollary I. Let the graph map p: G-+ G be a covering projection. Let u and v
be any two vertices of the base graph G and e any edge of G, not necessarily
incident either on u or on v. Then #p- 1(u) = #p- 1(v) = #p- 1(e). That is,
there is a number n such that all vertex fibers and all edge fibers have cardinality
n. D
(1 2)
(1 2)
1y 2
Figure 2.16. A graph with permutation voltages in 92.
2.4.5. Exercises
1. Construct the derived graph corresponding to the permutation voltage
graph in Figure 2.16.
2. Prove that the composition of two covering projections is a covering
projection.
3. Prove that the composition of two regular covering projections can
sometimes be an irregular covering projection. (Hint: Use Exercise 1.)
4. Let G be a graph obtained from the Petersen graph by doubling the
1-factor that extends from the star to the pentagon. (That is, for each
edge e in the 1-factor, attach another edge e' with the same endpoints.)
Assign permutation voltages in .51'10 to the bouquet B2 so that the
derived graph is isomorphic to G.
5. Use Exercise 4 to prove that the Petersen graph is an alternative Schreier
graph, by explicitly exhibiting the group, the subgroup, and the gener-
ating set.
6. Prove that any two graphs that cover the same graph have a common
covering space.
7. Let C be a hamiltonian cycle in the base space of a permutation voltage
graph. Under what circumstance is the preimage of C (under the natural
projection) a hamiltonian cycle in the derived graph?
8. Show that there exists both a regular covering projection and an irregular
covering projection of the complete graph K 7 onto the bouquet B3 •
9. Let F be an n-factor in the base space of a covering projection
p: G-+ G. Prove that the preimage p- 1(F) is ann-factor in the covering
space.
10. Prove or disprove: If the base graph of a (finite) covering projection is
n-connected, then so is each component of the covering graph.
11. Prove that any 2-sheeted covering projection is a regular covering projec-
tion.
To decide whether two vertices in an ordinary derived graph are in the same
component or to count the number of components, one considers a form of
Subgroups of the Voltage Group 87
isotropy group called a local voltage group, which is a subgroup of the voltage
group itself. The considerations involved here also lead to the proof of Babai's
theorem, that given a subgroup !14 of a group .511 and a Cayley graph G for .511,
there is a Cayley graph H for !14 such that G contracts onto H.
Example 2.5.1. Since every u-based closed walk in the base graph of Figure
2.17 traverses the edge between u and v an even number of times, it follows that
the local group at u is the subgroup {0, 2, 4} of the voltage group ~6 • Similarly,
the local group at v is also {0, 2, 4 }.
Theorem 2.5.1 (Alpert and Gross, 1976). Let u be a vertex in the base space of
an ordinary voltage graph (G, a) with voltages in the group .511. Then the vertices
ua and ub are in the same component of the derived graph Ga. if and only if the
voltage group element a- 1b lies in the local group .Jil(u).
Proof If the vertices ua and ub are in the same component of the derived
graph Ga., then there is a path from ua to ub. The image of that path under the
Figure 2.17. A connected ordinary voltage graph whose derived graph is not connected. The
voltages are in .2'6 •
88 Voltage Graphs and Covering Spaces
natural projection is a u-based closed walk in the base graph G, whose net
voltage lies in the local group .Jil(u). The value of that net voltage must be
a- 1b, by Theorem 2.1.2. Reversing these steps yields the converse conclusion.
0
Corollary 1. Let u and v be the vertices in the base graph G and W a walk from
u to v with net ordinary voltage c. Then the vertices ua and vh of the derived
graph lie in the same component if and only if the voltage group element a-lb lies
in the right coset .Jil(u)c of the local group at u. 0
Since the voltage group ..# acts transitively on an ordinary derived graph
ca, it follows that the components of Ga are mutually isomorphic. Example
2.5.2 illustrates that the components of a permutation derived graph need not
be isomorphic. Example 2.5.2 also illustrates another fact. If c is the ordinary
or permutation net voltage on a walk from a vertex u to a vertex v in the base
graph, then the local group at u is conjugate to the local group at v. In
particular, ..#( u) = c- 1..#( u)c.
Example 2.5.2. In the permutation voltage graph of Figure 2.18, the local group
at u is the direct product of the subgroup of all permutations on {1, 2} with the
subgroup of all permutations on {3, 4, 5}. The local group at v is the direct
product of the subgroup of all permutations on {1, 3} with the subgroup of all
permutations on {2, 4, 5}.
For the most part, there are permutation voltage analogs to the theorems
about ordinary voltage graphs. For instance, Theorem 2.5.2 and its corollaries
are analogs to Theorem 2.5.1 and its corollaries.
Theorem 2.5.2 (Ezell, 1979). Let u be a vertex in the base space of the
permutation voltage graph (G, a)n· Then the vertices U; and u1 are in the same
Figure 2.18. A connected permutation voltage graph whose derived graph is not connected. The
voltages are in 5"5 .
Subgroups of the Voltage Group 89
component of the derived graph G if and only if the symbols i and j are in the
same orbit of the action of the local group at u on {1, ... , n }.
Proof The vertices u; and u1 are in the same component of the derived
graph if and only if there is a path from u; to u1 . The image of such a path
under the natural projection is a u-based closed walk in the base graph, whose
net voltage lies in the local group at u. The value of that net voltage must be a
permutation taking i onto j, by Theorem 2.4.2. D
Corollary 1. Let u and v be vertices in the base graph G and W a walk from u
to v with net permutation voltage '1T in Y,. Then the vertices u; and v1 of the
derived graph lie in the same component if the right coset Y,(u)'TT contains a
permutation that maps i onto j. D
Corollary 2. Let (G, a)n be a connected permutation voltage graph, and let u
be a vertex of the base graph. The number of components of the derived graph
equals the number of orbits induced by the local group at u on the set {1, ... , n }.
D
Example 2.5.3. Figure 2.19a shows a voltage graph (G, a) with voltage group
.2"12 • Figure 2.19b shows the choice of a root u and a spanning tree T.
For every vertex v in the base graph G, there is a unique path in the tree T
from the root u to v. Define the T-potential a( v, T) of the vertex v to be the
net voltage along that path. The second step is to compute the T-potential of
every vertex of G. Figure 2.20 shows the result of this computation for the
voltage graph of Example 2.5.3.
90 Voltage Graphs and Covering Spaces
(a)
u 8 w u w
Figure 2.19. (a) A graph with voltages in the cyclic group .2"12 • (b) The choice of a root and a
spanning tree.
If the plus direction of the edge e in the base graph G has initial vertex v
and terminal vertex w, then the "T-voltage" aT( e) is defined to be the product
If there is an edge between two vertices whose voltage is the identity, then
those two vertices have the same local group, by Corollary 1 to Theorem 2.5.1
(or by Corollary 1 to Theorem 2.5.2, for permutation voltages). If both of
those vertices are identified with the new vertex obtained as a result of
contracting the base graph along that edge, then it follows that the local group
at every vertex is preserved in the resulting voltage graph. Thus, the local
group at the root vertex u with respect to the T-voltages could be obtained
from the bouquet that results from contracting the base graph G along the
entire spanning tree T. That is, the voltages on that bouquet generate the local
group at u with respect to the T-voltage assignment. However, these genera-
tors are simply the set of T-voltages on those edges of the base graph G that
are not in the spanning tree T, so one need not actually carry out the
contraction.
For Example 2.5.3, the local group at u with respect to the T-voltages is
now calculated to be the subgroup {0, 3, 6, 9}. The purpose of this calculation
is to show that the local group at u with respect to the original voltage
assignment a is the same subgroup, as proved by the following theorem.
lbeorem 2.5.3 (Gross and Tucker, 1979a). Let G be a graph, let a be either an
ordinary or a permutation voltage assignment for G, and let u be any vertex of G.
Then the local group at u with respect to the T-voltages, for any choice of a
spanning tree T, is identical to the local group at u with respect to a.
Proof The net voltage on any u-based closed walk is the same with
respect to T-voltages as with respect to a. D
Corollary. The number of components of the derived graph Ga equals the index
in the voltage group of the subgroup generated by the T-voltages. D
92 Voltage Graphs and Covering Spaces
such that for k = 1, ... ' r, the walk w(k) differs from its predecessor w(k-l)
by either the insertion or the removal of a directed edge followed by its
reverse. By induction, one readily verifies that every u-based closed walk is
equivalent to a unique reduced walk.
The equivalence classes formed in the fundamental semigroup of u-based
closed walks form a group, in which the inverse of the equivalence class of a
u-based closed walk is the equivalence class of the reverse walk. This group is
called the "fundamental group of the graph G based at u ", and is denoted
'1T 1(G, u). For a connected graph G, the isomorphism type of its fundamental
group is independent of the choice of the base vertex u.
There is an important correspondence between voltage assignments for a
graph G and representations of its fundamental group. Since the net voltage
on any two equivalent u-based closed walks is the same, any voltage assign-
ment for G in a group .J/1 induces a homomorphism '1T1(G, u) -+ .J/1. That is,
every element of '1T1(G, u) is mapped onto the net voltage assigned to any
u-based closed walk representing that element.
Conversely, every homomorphism '1T 1(G, u) -+ S/1 is induced by some volt-
age assignment in S/1 on the graph G. Given such a homomorphism, first
choose a spanning tree T and assign to every edge of T the identity voltage.
For any edge of G not in T, suppose that its initial point is v and that its
terminal vertex is w. Let W be the u-based closed walk that begins with the
unique path in T from u to v, next traverses the edge e, and then concludes
with the unique path in T that runs from w to u. Assign as a voltage to the
edge e the image in the group S/1 of the equivalence class in '1T 1(G, u) of the
walk W.
From the correspondence just described, one may obtain for graphs all of
the standard topological theorems on the relationship between fundamental
groups and covering spaces.
Proof Choose a spanning tree T for the base graph G. By Theorem 2.5.4,
the derived graph GaT corresponding to the T-voltages is isomorphic to Ga, so
it is sufficient to prove that GaT contracts onto a Cayley graph for d. Since
ar(T) = 0, by Theorem 2.1.3, the preimage p- 1(T) of T under the natural
projection has #d components, each mapped isomorphically by p onto the
spanning tree T. In particular, for each element d of the voltage group d,
one component of p -I(T) contains all of the vertices of GaT with the
subscript a. If each component of p- 1(T) is contracted onto a single vertex,
which is identified with the common subscript of all the vertices in that
component, then the resulting graph is a Cayley graph for d. One might
observe that one can also contract the base graph G along the spanning tree T,
thereby obtaining a bouquet of circles. The voltages on these circles generate
d, because GaT is connected. Then the contracted derived graph is precisely
the derived graph for the contracted voltage graph. 0
Proof Since the group d acts freely on the Cayley graph G, so does its
subgroup !!4. It follows that G may be derived by assigning voltages in !!4 to
the quotient of G under the action of !!4. By the theorem, it follows that G
contracts onto a Cayley graph for !!4. o
The "genus of a group" is defined to be the least genus occurring among all
its Cayley graphs. White (1973, p. 80) asked whether a finite group d can have
a subgroup !!4 whose genus exceeds the genus of d. It is an immediate
consequence of Corollary 1 that it cannot.
Corollary 2 (Babai, 1977). The genus of a finite group is greater than or equal
to the genus of each of its subgroups. 0
2.5.5. Exercises
Exercises 1-4 concern the voltage graphs in Figure 2.22.
1. Calculate the local group at the vertex u of the voltage graph in Figure
2.22a. How many components are there in the derived graph?
2. Calculate the local group at the vertex u of the voltage graph in Figure
2.22b. How many components are there in the derived graph?
3. Calculate the local groups at vertices u and v of the voltage graph in
Figure 2.22c. How many components are there in the derived graph?
94 Voltage Graphs and Covering Spaces
9 u
7 (2,3)~(4,2)
(b)
w
2
(a)
(12)
(d)
(c)
Figure 2.22. Four voltage graphs. (a) Ordinary voltages in .;2"18 • (b) ordinary voltages in
(c) ordinary voltages in 94. (d) permutation voltages in .5/'7.
.;2"6 X .;2"6 •
4. Calculate a generating set for the local group at the vertex u of the voltage
graph in Figure 2.22d. How many components does the derived graph
have?
5. Prove that any graph map f: G -+ H induces a homomorphism
J.: '11'1(G, u)-+ (H, f(u)), defined as follows: If [w] is the equivalence
class of any closed u-based walkinG, then J.([w]) = [f(w)].
6. Prove that if p: G-+ H is a covering projection, then p.: '11'1(G, u)-+
'11'1(H, p(u)) is one-to-one.
7. Let !!4 be a subgroup of finite index in '11'1(G, u). Prove that there is a
covering p: G-+ G and a vertex u E p- 1(u) such that p.('11'1(G, u)) = !!4.
Prove that G is unique up to graph isomorphism. (Thus, G can be called
the covering space of G corresponding to the subgroup !14.)
3
Surfaces and Graph lmbeddings
It is a classical theorem of Rado (1925) that any compact surface can be
dissected into a finite number of triangular regions. The incidence structure of
these regions determines a combinatorial object called a "complex", which is a
higher-dimensional analog of a graph. Whereas a graph has only zero-dimen-
sional "cells" called vertices and one-dimensional "cells" called edges. a
complex may include pieces of any dimension. In particular. a 2-complex also
contains some two-dimensional cells. Often these cells correspond to triangu-
lar regions, but more generally they may be arbitrary polygons, with any
number of sides. Knowing that any closed surface can be represented by such
a finite combinatorial object enables us to classify the closed surfaces, that is.
to construct a countable list that includes them all and contains no repetitions.
The combinatorial viewpoint also enables us to decide which graphs can be
imbedded in which surfaces. Even more significantly, it enables us to construct
all of the essentially different possibilities for imbedding an arbitrary graph in
an arbitrary surface.
95
96 Surface~ and Graph lmbeddinll"
is called the "carrier" of the simplicial complex K. It is possible that the same
point set in R 11 can be the carrier of more than one simplicial complex, as
illustrated in Example 3.1.1. If m is the largest integer such that K contains an
m-simplex, then K is an "m-complex". The collection of all k-simplexes of K
for k ~ r is called the "r-skeleton" of K and is denoted Kul.
Example 3.1.1. Figure 3.1 shows three simplicial 2-complexes in R ·1• The
complex on the left has one 2-simplex: (v. w, x): nine 1-simplexes: (u. t•).
(u. w). (u. x). (v. w). (v. x). (v. y). (w. x). (w. y). (x. y): and .ftt•e
0-simplexes: (u), (v), (w), (x). (y). The center 2-complex has two 2-sim-
plexes: (u. v, w), (u. w, x); nine }-simplexes: (u. v). (u. w). (u. x). (u. y).
(v. w). (v. z). (w. x), (w, y), (x, z); and six 0-simplexes: (u). (v). (w).
(x). (y). (z). Note that the right-hand 2-complex has the same carrier as the
one on the left. but this time the 2-simplex ( v, w. x) is replaced by the three
2-simplexes (v. w, z). (w. x, z). (x. v, z) and their faces. In each instance. the
1-skeleton is the graph that remains when the 2-simplexes are discarded. (One
discards only the 2-simplexes, not their various faces.)
Surf:tl'l'' mul Simplid:tl ( 'ompll'U'' 97
II II
II'
z
\' y
3. 1.3. Triangulations
A "triangulation" of a topological space X is a homeomorphism h from the
carrier of some simplicial complex K to the space X. The image of a simplex
of K under h is called a simplex of the triangulation. The aforementioned
theorem of Rado states that every compact surface has a triangulation.
Example 3. 1.2. On the left of Figure 3.2 is a triangulation of a torus, where the
sides of the rectangle are to he identified in pairs, as in Chapter 1, so that vertices
nwtch up. Enen though all its faces are three-sided, the right side of Figure 3.2 is
not a triangulation (~(a torus since no two different simplexes of a triangulation
ha11e the exact same vertices. A /so an n-simplex of a triangulation must have
n + 1 distinct vertices, and this "triangulation" has only one vertex.
The vertices and edges of a graph in Ill 3 are not the 0-simplexes and
]-simplexes of a triangulation if the graph has any multiple edges or loops.
This can always be rectified, of course, by a subdivision of the graph inserting
one vertex on each multiple edge and two vertices on each loop.
The process of subdividing self-loops and multiple edges to make a nonsim-
plicial graph simplicial extends to higher dimensions. We might think of a
loop as an "inadequately subdivided" edge, and we might also think of a
nontriangular polygon (or a polygon with repeated vertices or edges) as
u 11 w u
11 u
X X
y y
11 11
u II w u
Figure 3.2. A triangulation and a non-triangulation or a torus.
Surfaces and Simplicial Complexes 99
1\
w w
u e u .
u
,.
e v
Figure 3.3. The first barycentric subdivision of a 2-simplex.
Example 3.1.3. The imbedding of a bouquet of two circles in a torus on the left
of Figure 3.4 is not cellular, since one region is homeomorphic to an open
annulus. The imbedding on the right is cellular. Observe that the closure in the
surface of the one and only face is not a closed disk, but instead, the whole torus.
be cellular. Thus, when an imbedding in the sphere is drawn in the plane, the
reader is expected to remember the necessary point at infinity.
Given a 2-cell imbedding G ~ S, the collection of all edges and vertices
lying in the closure of a region can be organized into a closed walk in the
graph G by traversing a simple closed curve just inside the region. This closed
walk is unique up to the choice of initial vertex and direction, and is called the
"boundary walk" of the region. A face of a 2-cell imbedding can usually be
specified unambiguously simply by giving its boundary walk (see Exercise 17).
For this reason, we shall frequently abuse the language with phrases such as
"the face abc d " where abc d is in fact the boundary walk of the face. An
occurrence of an edge in the boundary walk of a region is called a "side" of
the region. The "size" of a region is the length of its boundary walk. Since
individual edges may occur twice in a boundary walk, the size of a region can
be greater than the number of edges that are sides of the region. An n-sided
face in a 2-cell imbedding is called an "n-gon", although 3-gons and 4-gons
are also called triangles and quadrilaterals, respectively.
Example 3.1.4. The imbedding in the sphere given in Figure 3.5 has three
regions. There is a quadrilateral abc d, a triangle e e f, and a 9-gon
abggcdhfh. The 9-gon has only seven different edges as its sides.
Example 3.1.5. Four different imbeddings of the same tree in the sphere
(plane) are given in Figure 3.6. A rotation by 90° gives an equivalence between
the first and second imbeddings. A reflection in a horizontal line through the
vertex y yields an equivalence of the second and third imbeddings. However, the
fourth imbedding is not equivalent to any of the previous three.
·xx xx
X W W
Figure 3.6.
U X U
the fourth imbedding in Figure 3.6 is weakly equivalent to the other three. It
will become apparent in Section 3.2 that equivalence is easier to analyze than
weak equivalence, although it may seem at first less natural. In any case, it
must be remembered that any discussion of equivalence here presumes that the
vertices and edges of the imbedded graph have been labeled. "Congruent" is a
possible synonym for "weakly equivalent".
Given a 2-cell imbedding G ~ S, the "dual imbedding" G* ~Sis defined
as in Chapter 1: the vertices of G* correspond to regions of the imbedding
G ~ S and vice versa. The dual graph gives an efficient description of the
incidence relationship among the regions of an imbedding, for which purpose
it is used extensively in Section 3.3.
Example 3.1.6. In Chapter 1 the projective plane was defined to be the surface
obtained by identifying the boundary of a disk to the boundary of a Mobius band.
In Figure 3.7, polygon 2 becomes a Mobius band when the h-edges are
identified. Thus the hexagon formed by polygons 1, 2, and 3 represents the
projective plane. If the consecutive sides a, b, care considered to be just one edge,
we get the simpler representation on the right of Figure 3.7. In particular, the
projective plane can be viewed as the closed unit disk with antipodal points on the
unit circle identified in pairs. Thus, the definition given here for a projective plane
Surfaces and Simplicial Complexes 103
b 2 b e
corresponds to that given in geometry as the space of lines through the origin in
R 3 (represented by points on the northern hemisphere of the unit sphere with
antipodal points on the equator identified).
Example 3.1.7. The Klein bottle was defined in Chapter 1 as the surface
obtained by identifying the boundaries of two Mobius bands to the boundary of a
sphere with two holes. Since a sphere with two holes is homeomorphic to an
annulus and identifying one boundary component of an annulus to a Mobius band
still/eaves a Mobius band (with a "collar" on its boundary), the Klein bottle can
also be viewed as the surface obtained by identifying the boundaries of two Mobius
bands alone. In Figure 3.8, polygon 2 becomes a Mobius band when the h-edges
are identified. Polygons 1 and 3 also form a single Mobius band when their edges
are identified (first identify the d-edges to get a rectangle like polygon 2). Thus
the polygons I, 2, and 3 together represent the Klein bottle. On the right in
Figure 3.8 is a simpler representation.
If the direction of one copy of edge e in Figure 3.7 is reversed, the resulting
surface is the sphere. Similarly, if the direction of one copy of edge e in Figure
3.8 is reversed, the resulting surface is the torus.
Example 3.1.8. If all the identifications are performed on the two polygons in
Figure 3.9 except on the sides labeled e, one obtains two tori each with a
puncture. When thee-sides are then identified, the result is a surface of genus 2.
:
a c d
1
c
2
3
d
b
a ·I d
I·
Figure3.8. Two representations of the Klein bottle.
104 Surfaces and Graph Imbeddings
(l·D -e e - 88 Figure 3.9. How to obtain a genus two surface from polygons.
a b c d a b c d
g 3 4 g
f
f f f
1 2
-
e
e e
d a
b c d a b c
Figure 3.10. Three views of the same toroidal imbedding of K 4 , 4 minus a 1-factor.
viewed as a surface in which finite sets of points are identified. Figure 3.11
shows a typical pseudosurface.
Alpert (1975) showed that covering space techniques could be useful in
constructing triangulated pseudosurfaces that represent "twofold triple sys-
tems", thereby reviving interest in the connection between block design and
topological graph theory explored earlier by Heffler (1897) and Emch (1929).
Subsequent advances in exploiting this connection were achieved by White
(1978) and Garman (1979). The basic ideas are easy enough to explain.
A "block design" consists of a set V of objects and a set B of subsets of V,
called "blocks", such that (i) each object appears in the same number of
blocks, called the "replication number", and (ii) each block has the same
number of objects, called the "blocksize". If every pair of objects appears in
the same number of blocks, called the "balancing index", then the design is
called "balanced". A trivial kind of balanced block design has every object in
every block, and is called a "complete" block design. If the blocksize is less
than # V then the design is called "incomplete".
Balanced incomplete block designs, abbreviated BIBDs, are important to
statisticians because they facilitate the design of experiments on samples of
limited size. They are also of considerable interest to algebraists and combina-
torial analysts.
A twofold triple system is a BIBD with blocksize equal to 3 and balancing
index equal to 2. Alpert proved that there is a bijective correspondence
between the twofold triple systems and the (equivalence classes of) triangula-
tion of closed pseudosurfaces by complete graphs. This topological insight
facilitates an easy proof that some classical twofold triple systems of
Bhattacharya (1943) and Hanani (1961) are distinct and that they are also
distinct from new twofold triple systems that Alpert developed from the
Ringel-Youngs constructions in cases 1, 6, 9, and 10 of the Heawood
map-coloring problem.
White and Garman both explored the extension to "partially" balanced
incomplete block designs with two association classes of objects. These associ-
ation classes partition the objects so that pairs of objects from one class have
one balancing index while pairs from the other class have another balancing
index. To obtain his results, Garman formally extended the imbedded voltage
graph construction described in Chapter 4 to voltage graphs imbedded in
pseudosurfaces.
106 Surfaces and Graph Imbeddings
Figure 3.12. An orientation for an annulus and an attempt at an orientation for a Mobius band.
3.1.7. Orientations
Each face of an imbedding G ~ S has two possible directions for its boundary
walk. A face is assigned an "orientation" by choosing one of these two
directions. An "orientation for the imbedding" G ~ S is an assignment of
orientations to all the faces so that adjacent regions induce opposite directions
on every common edge. If the graph G is the 1-skeleton of a triangulation of
the surface S, then an orientation for the imbedding G ~ S is called an
"orientation of the triangulation". A surface is "orientable" if it has an
imbedding with an orientation.
It is a simple matter to test whether a given connected imbedding G ~ S
has an orientation. First, assign an arbitrary orientation to a particular face.
This forces an orientation of each face that shares an edge with the original
face. Since the surface is connected, the process can be continued until the
orientation of each face has been forced. Either the result is an orientation for
the imbedding or else the given imbedding has no orientation.
incident on v and the endpoints of those edges. Such a graph does actually
look something like a heavenly star, which is the inspiration for the terminol-
ogy. A topological space X is a closed surface if and only if there is a
triangulation of X in which the carrier of the star of each vertex is homeomor-
phic to a closed disk.
The "link" of a vertex v is defined to be the maximal subcomplex of star( v)
that does not contain v itself. If K is a graph, then link( v) is the totally
disconnected graph comprising the neighbors of v and no edges. If the carrier
of K is a surface, then the link of every vertex is a cycle through its neighbors.
If the carrier of K is a pseudosurface, then the link of a singular vertex is the
disjoint union of two or more cycles.
The standard definition of a "local property" of a complex K is to say it is
a property that is true of the star or the link of every vertex. In the case of
graphs, we sometimes associate it with the subgraph induced by the link of a
vertex.
For instance, we might say that a graph is "locally hamiltonian" if for every
vertex v, the subgraph induced by link(v) is hamiltonian. The principal of this
concept is that a graph cannot triangulate a surface unless it is locally
hamiltonian.
3.1.9. Exercises
1. Which points in the given set X fail to have a neighborhood homeomor-
phic to the unit disk?
(a) X= {(x, y, z) I xyz = 0}.
(b) X= {(x, y, z) lx 2 + y 2 = z 2 }.
2. Prove that the r-skeleton of a simplicial complex is a simplicial complex.
3. Figure 3.3 shows the first barycentric subdivision of the 2-complex
consisting of a 2-simplex and its faces. Prove that the first barycentric
subdivision is itself a simplicial complex.
4. Draw imbeddings in the plane of the 1-skeleton of the first two 2-com-
plexes in Figure 3.1. Show that the 1-skeleton of the first barycentric
subdivisions of these 2-complexes do not have planar imbeddings. (Hint:
Show that the first contains K 3, 3 and that the second contains K 5 .)
5. Find a 2-complex that cannot be imbedded in the sphere (or for that
matter in any surface) but whose first barycentric subdivision has a
1-skeleton that is imbeddable in the plane.
6. Let vn = (cos2'1Tjn, sin2'1T/n) E 111 2 for n = 1,2, .... Consider the col-
lection K of all simplexes of the form (vn, vn+l) or (vn) for n = 1, 2, ....
If the restrictions of finiteness were deleted from the definition of
simplicial complex, would K be a simplicial complex? Could K triangu-
late the unit circle?
108 Surfaces and Graph Imbeddings
Example 3.2.1. Figure 3.13 shows a graph imbedding and the associated
reduced band decomposition. The 0-bands assume the names of their respective
vertices and the 1-bands assume the names of their respective edges.
- a
r----- -----...,
I I
bl I I t•
- a
Figure 3.13. (left) An imbedding of a graph with two vertices and three edges in the Klein bottle.
(right) The corresponding reduced band decomposition.
3.2.2. Orientability
A band decomposition is called "locally oriented" if each 0-band is assigned
an orientation. Then a 1-band is called "orientation-preserving" if the direc-
tions induced on its ends by adjoining 0-bands are the same as those induced
by one of the two possible orientations of the 1-band; otherwise the 1-band is
called "orientation-reversing". These two possibilities are illustrated in Figure
3.14.
An edge e in the graph imbedding associated with a locally oriented band
decomposition is said to have "(orientation) type 0" if its corresponding
1-band is orientation preserving, and "(orientation) type 1" otherwise. A walk
in the associated graph has type 1 if it has an odd number of type-1 edges and
has type 0 otherwise.
Figure 3.14. Two orientation preserving bands (left) and two orientation reversing bands (right).
Band Decompositions and Graph Imbeddings Ill
and only if there exists a type-1 cycle in the associated graph, which implies
immediately that the surface contains a Mobius band. For instance, the cycle
e 2 , e 3 in Figure 3.13 has type 1, no matter what the choice of local orientations.
Example 3.2.2. Consider the reduced band decomposition shown on the left in
Figure 3.15. First, reverse the orientation of the 0-band labeled u. Then reverse
the orientation of the 0-band v. The resulting decomposition has no orientation
reversing 1-bands, so the underlying surface is orientable, despite all the twisted
!-bands in the leftmost drawing.
Lemma 3.2.1. Let S and T be surfaces with the same number of boundary
components. Let S' and T' be closed surfaces obtained by identifying the
boundaries of disks to the boundary components of S and T, respectively. Then
the surfaces S and T are homeomorphic if and only if the surfaces S' and T' are
homeomorphic. 0
Example 3.2.3. For instance, if we orient the 0-bands of Figure 3.13 both to be
clockwise on the page, we have the following combinatorial description, where the
superscript "1" designates a type-1 edge and the absence of a superscript a type-0
edge:
v1• e1 e2 e/
v2 • e/ e 1 e 2
To make this method of combinatorial description precise, we define a
"rotation" at a vertex v of a graph to be an ordered list, unique up to a cyclic
permutation, of the edges incident on that vertex. Naturally, every v-based
Band Decompositions and Graph lmbeddings 113
Theorem 3.2.3. Every pure rotation system for a graph G induces (up to
orientation-preserving equivalence of imbeddings) a unique imbedding of G into
an oriented surface. Conversely, every imbedding of a graph G into an oriented
surface induces a unique pure rotation system for G. 0
w
Figure 3.17. A rotation projection for K 4 and the corresponding reduced band decomposition.
If a given graph is simplicial, then the list format of a rotation system may
give adjacent vertices instead of edges. This is called the "vertex form of a
rotation system". In this context, it is also natural to list only the sequence of
vertices in a boundary walk.
Example 3.2.4. The rotation system for K 4 whose projection is illustrated on the
left in Figure 3.17 can be given a list format in either the edge form or the vertex
form.
u. c1 ba u. x 1 wv
v. fad v. xuw
w. dbe w. DUX
X. efcl X. wvu 1
Face Tracing Algorithm. Assume that the given graph G has no 2-valent
vertices. Choose an initial vertex v0 of G and a first edge e1 incident on v0 • Let v 1
be the other endpoint of e 1 . The second edge e 2 in the boundary walk is the edge
after (resp., before) e 1 at v 1 if e 1 is type 0 (resp., type I). If the edge e 1 is a
loop, then e 2 is the edge after (resp., before) the other occurrence of e1 at v 1 • In
general, if the walk traced so far ends with edge e; at vertex D;, then the next
edge ei+l is the edge after (resp., before) e; at D; if the walk so far is type 0
(resp., type I). The boundary walk is finished at edge en if the next two edges in
the walk would be e 1 and e 2 again. To start a different boundary walk, begin at
the second edge of any corner that does not appear in any previously traced faces.
If there are no unused corners, then all faces have been traced.
Observe that the walk does not necessarily stop when the first edge e 1 is
encountered a second time; we might not be on the same side of e 1 as at the
beginning. The followup by the edge e 2 is what confirms that we are on the
original side of e 1, assuming of course that the vertex v 1 does not have valence
2 (see Exercise I4).
u. a 1 fbd 1 a 1 e 1 be
v. cfg
w. e1 d 1 g
Begin the first face at vertex u with the edge a at the corner ( c, a). The next edge
in the boundary walk is d, the edge before (since the walk a is type I) the other
occurrence of the edge a at the vertex u. The next edge is g, the edge after d
(since the walk ad is type 0) at the vertex w. The next edge is c. Since the
following two edges would be a and d again, the walk terminates with c, yielding
the face adgc. Since the corner (a, f) has not yet appeared, we can begin a
second boundary walk with edge f at vertex u. This time the face f g e a is
obtained. The third and fourth faces b c f and deb are obtained in a similar
fashion by starting with edge b from corner (/, b) and edge d from corner ( b, d).
Figure 3.I8 shows these four faces, first separately, then assembled together.
116 Surfaces and Graph Imbeddings
u a u
di<' , ~,
b
g a
u u
w ' g I' v w e u
a a
fA g'
u u
u u
I b
u c v w e u
Figure 3.18. The faces from Example 3.2.5 assembled into a Klein bottle, so that the local
orientations at vertices u, v, and w are correctly completed.
t. d 1 be 1
u. fd 1 a 1
v. c1 ba 1
w. cl elf
has as its faces the polygons dfe, bad, ecb, and fca. When these faces are
assembled, they yield a 2-sphere. Thus, as we have mentioned, the existence of
orientation reversing edges need not mean that the surface is nonorientable,
provided that no cycle contains an odd number of orientation reversing edges.
3.2.7. Duality
Let B be a band decomposition of the surface S. Suppose each 2-band has
been given a specific orientation. The dual band decomposition of B, denoted
B *, is defined by letting the i-bands of B be the (2 - i)-bands of B*, for
i = 0, 1, 2; the ends of each 1-band of B become the sides of the corre-
sponding 1-band of B* and vice versa. If G ~ S is the graph imbedding
associated with the decomposition B, then the imbedding associated with B*
is naturally the dual imbedding G* ~ S*.
Let e be an edge of the imbedding G ~ S associated with the decomposi-
tion B. The orientation type of the dual edge e* depends on the choice of
orientations for the 2-bands of the primal imbedding or, equivalently, on the
direction given to the closed walk around the boundary of each primal face.
The edge e appears twice in the course of listing all directed face boundaries.
If the two appearances have opposite directions, then the dual edge e* is type
0; otherwise it is type 1. Thus, in Example 3.2.4, if the directed boundaries are
u vx wux wvx and uvw, then (ux)*, (v w)*, and (wx)* are type 0, whereas
Band Decompositions and Graph Imbeddings 117
Historical Note. The earliest form of Theorem 3.2.3 was implicitly given in a
dualized vertex form by Heffter (1891) and used extensively by Ringel in the
1950s. Evidently unaware of the existence of this work, Edmonds (1960)
invented the primalized vertex form, and its details were subsequently made
accessible by Youngs (1963). The full generalization to all graphs, simplicial or
not, requires the switch from the vertex form to the directed-edge form, and it
was developed by Gross and Alpert (1974).
118 Surfaces and Graph Imbeddings
3.2.9. Exercises
1. Consider the pure rotation system
u. cbd x. abg
v. ajh y. idf
w. ecfe z. hijg
Figure 3.20. A rotation system for a cubic graph. Hollow dots have counterclockwise rotation.
10. Let G --+ S be an imbedding. Suppose that / 1 and / 2 are two distinct
faces of the imbedding whose boundaries share the edge e. Let G' --+ S'
be the imbedding obtained from the rotation system for G --+ S by
"giving the edge e an extra twist" (that is, by changing the orientation
type of e). Show that the imbedding G --+ S' has one less face than the
imbedding G--+ S.
ll. Use Exercise 10 to show that every connected graph has an imbedding
with only one face (Edmonds, 1965).
12. Let G be a regular graph of valence 3. Each vertex has exactly two
possible rotations. Thus a rotation system for G can be given from a
drawing of G in the plane by, say, a hollow dot for a vertex whose
rotation is counterclockwise in the drawing and a solid dot if clockwise.
How many faces are there in the rotation system given by this method in
Figure 3.20?
13. Two rotation projections are given in Figure 3.21. Choose a spanning tree
in each graph and reverse some vertex orientations so that each edge in
the spanning tree is type 0. Then determine which of the two surfaces is
orientable.
14. Show by an example that the procedure given in the Face Tracing
Algorithm for terminating a boundary walk may stop the walk too soon
if the vertex v1 is 2-valent. Suggest a way of handling 2-valent vertices.
types of closed, connected, orientable surfaces, and moreover, that they are
distinguishable by a single integer-valued invariant, the Euler characteristic.
Similarly, the surfaces N1, N2 , N3 , ••• form a complete set of representatives of
the homeomorphism types of closed, connected, nonorientable surfaces, and
they too are distinguishable by Euler characteristic.
The proof of the classification theorem is facilitated by a designation of
convenient models of the surfaces. As a model of Sg we take the surface of an
unknotted g-holed solid doughnut in 3-space.
A "meridian" in this model is a topological loop on Sg that bounds a disk
in the g-holed solid that does not separate the solid. The topological closure of
the complement of the g-holed solid is also a g-holed solid doughnut. A
"longitude" on Sg is a topological loop that bounds a disk in the complemen-
tary solid that does not separate the complementary solid. Figure 3.22 shows a
three-holed solid doughnut and standard sets of meridians and longitudes.
As a model of Nk we take the surface obtained from a sphere by first
deleting the interiors of k disjoint disks and then closing each resulting
boundary component with a Mobius band. No closed nonorientable surface is
imbeddable in 3-space.
The classifications are derived in three steps. The first step is to calculate an
Euler characteristic-relative to a particular cellular imbedding of a particular
graph-for each of the standard surfaces. Second, we show that the Euler
characteristic is an invariant of every standard surface, that is, it is indepen-
dent of the choice of a graph and of the choice of an imbedding (provided that
the imbedding is cellular).
Since the standard orientable surfaces have different Euler characteristics, it
follows from the second step that no two of them are homeomorphic.
Similarly, the nonorientable standard surfaces are pairwise distinct. In the
third step, we introduce the concept of "surgery" to prove that every closed
surface is homeomorphic to one of the standard surfaces.
longitudes
meridians
Figure 3.22. An unknotted 3-holed solid doughnut in 3-space, with standard meridians and
longitudes.
The Classification of Surfaces 121
Proof Theorem 1.4.1 implies this result for g = 0. For g > 0, let the
vertices of the graph G be the g intersection points of a standard set of g
meridian-longitude pairs on the surface Sg. The edges of G include the g
meridians, the g longitudes, and a set of g - 1 "bridges" that run directly
between consecutive intersection points. Thus,'the graph G has g vertices and
3g - 1 edges. One verifies by an induction on g that there is only one face
and that it is a 2-cell. It follows that x(G--+ Sg) = g- (3g- 1) + 1 = 2 - 2g.
D
Proof For k = 0, this is true by Theorem 1.4.1. For k > 0, choose a point
on the center loop of each of the k crosscaps as a vertex of G. The edges of G
are the k center loops plus k - 1 bridges to link the vertices to each other.
Thus, there are k vertices and 2k - 1 edges. An induction argument on the
number k proves that there is only one face and that it is a 2-cell. Therefore,
x( G ~ Nk) = k- (2k- 1) + 1 = 2- k. D
the Euler characteristic, and thereby, the distinctness of the standard surfaces,
is the second step in the classification of surfaces.
Proof Theorem 1.4.1 implies that the conclusion holds for the sphere S0 •
As an inductive hypothesis, assume that it holds for the surface Sg, and
suppose that the graph G is cellularly imbedded in Sg+ 1 •
First, draw a meridian on Sg+ 1 so that it meets the graph G in finitely many
points, each in the interior of some edge of G and each a proper intersection,
not a tangential intersection. If necessary, we subdivide edges of G so that no
edge of G crosses the meridian more than once, which does not change the
relative Euler characteristic of the imbedding into Sg+ 1 . Next thicken the
meridian to an annular region R, as illustrated in Figure 3.23.
A homeomorphic copy of Sg can be obtained by excising from Sg+ 1 the
interior of region R and by then capping off the two holes with disks. We now
construct a cellular imbedding G' -+ Sg. The vertex set of the graph G' is the
union of the vertex set VG and the intersection set of G with the boundary of
R. If the graph G intersects the meridian in p places and if the region R is
selected to be adequately narrow, then #VG' = #VG + 2p.
The edge set EG' contains all the #EG- p edges of G that do not cross the
meridian. Furthermore, it contains the arc segments on the boundary compo-
nents of the region R that connect adjacent points of intersection of
boundary( R) and G. Since G intersects the meridian in p places, there are p
such arcs on each boundary component, for an additional subtotal of 2p. The
edge set EG' also contains the segments of edges of G that run from vertices of
G to intersections of G and boundary(R). There are 2p such segments in all.
Thus,
Each face of the imbedding G -+ Sg+l is cellular. It follows from the Jordan
curve theorem that each of the p subarcs of the meridian subtended by G,
considered in succession, cuts a region into two parts, thereby yielding p
additional faces in the imbedding G' -+ Sg, it follows that
Corollary. Let i andj be distinct nonnegative integers. Then the surfaces S; and
sj are not homeomorphic.
x( G ~ Nk+l) = 1 - k = 2- (k + 1) D
Corollary. Let i andj be distinct nonnegative integers. Then the surfaces N; and
~are not homeomorphic. D
Remark 3.3.1. Suppose that the imbedding G ~ S induces the rotation system
R and that the imbedding G' ~ S' results from edge-deletion surgery at edge e.
Then the rotation system R' induced by G' ~ S' can be obtained simply by
deleting both occurrences of edge e from R. For a rotation projection, this means
erasing the edge e. For a reduced band decomposition, it means deleting the
e-band.
The Classification of Surfaces 125
There are three different possible occurrences when the edge e is deleted,
which are called cases i, ii, and iii. The effect of the surgery depends on the
case.
i. The two sides of the edge e lie in different faces, / 1 and / 2 . Then
deleting the fcband, the / 2-band, and the e-band leaves one hole,
which is closed off with one new 2-band.
ii. Face f is pasted to itself along edge e without a twist, so that deleting
the /-band and the e-band leaves two distinct holes (and possibly
disconnects the surface). These holes are closed off with two new
2-bands.
iii. Face f is pasted to itself with a twist along edge e, so that the union of
the /-band and the e-band is a Mobius band. Then deleting the /-band
and the e-band leaves only one hole, which is closed off with one new
2-band.
-
Figure 3.24. Sliding edged along edge e.
126 Surfaces and Graph lmbeddings
Proof Suppose that a; and b; are the arcs in boundary(T) to which the
band for 1'; is attached, for i = 1, 2. Then there is a homeomorphism of T to
itself taking a 1 to a 2 and b1 to b2 • This homeomorphism can be extended to
the attached bands by the product structure of I X I, since neither band is
"twisted" (the surfaces T, T1 , and T2 are all orientable by assumption). The
result is a homeomorphism from T1 to T2 • D
Example 3.3.1. It is not visually obvious that the two reduced band decomposi-
tion surfaces in Figure 3.25 are homeomorphic. However, one can easily verify
by tracing along the boundary walks that both surfaces have one boundary
component. Moreover, deleting band c from both yields two surfaces that have
two boundary components each and are obviously homeomorphic. It follows from
The Oassification of Surfaces 127
a b c d a b c d
~ Figure 3.25.
~
Homeomorphic reduced band decomposition surfaces.
Lemma 3.3.6 that the two surfaces illustrated are also homeomorphic. The
corollary implies that the imbedding surfaces induced by the rotation systems
are homeomorphic.
Proof As the basis for an induction, suppose that #Ec + #EH = 0. Then
the graphs G and H have only one vertex each, and the imbeddings G ~ S
and H ~ T have one face each. Hence, the surfaces S and T are both
2-spheres, and x(G ~ S) = x(H ~ T) = 2.
Assume that the inclusion holds when #Ec + #EH = n, and suppose that
G ~ S and H ~ Tare imbeddings such that #Ec + #EH = n + 1. If the
imbedding G ~ S had more than one face, then the graph G would have an
edge e common to two faces. By case i of Theorem 3.3.5, the imbedding
( G - e) ~ S obtained by surgery at e would have Euler characteristic equal
to x(G ~ S). Thus, the conclusion would follow by induction. Therefore, one
may assume that the imbedding G ~ S has only one face. Likewise, one may
assume that the imbedding H ~ T also has only one face. Furthermore, the
same argument applied to the dual imbeddings permits one to assume that the
graphs G and H have but one vertex each.
Now let d and e be arbitrary edges of the graphs G and H, respectively,
and let G' ~ S' and H' ~ T' be the imbeddings that result from surgery at d
and e, respectively. Case ii of Theorem 3.3.5 implies that x(G' ~ S') =
x(G ~ S) + 2 and that x(H' ~ T') = x(H ~ T) + 2. Thus, x(G' ~ S') =
x(H' ~ T'). Since G and H are both one-vertex graphs, it follows that G',
H', S', and T' are each connected. Therefore, by induction, x(G' ~ S') and
x(H' ~ T') are even numbers less than 2, from which it follows that x(G ~ S)
and x(H ~ T) are even numbers less than 2. By induction, it follows that the
surfaces S' and T' are homeomorphic. By the corollary to Lemma 3.3.6, we
infer that the surfaces S and T are homeomorphic. D
128 Surfaces and Graph Imbeddings
Lemma 3.3.8. Let T be a surface with one boundary component, and let T1 be a
surface also with one boundary component, obtained from T by identifying the
ends of a 1-band to disjoint arcs in the boundary. Let T2 be another surface with
one boundary component obtained from Tin the same way. Then T1 and T2 are
homeomorphic.
Proof The proof is essentially the same as that for Lemma 3.3.6, but this
time one may observe that both attached bands must have a "twist" so that
the surfaces T1 and T2 still have only one boundary component. D
Proof Since the surfaces S' and T' are homeomorphic, and since the
imbeddings (G- d)~ S' and (H- e) ~ T' are both one-faced, the reduced
band decomposition surfaces for (G- d) ~ S' and for (H- e)~ T' are
homeomorphic surfaces with one boundary component. By Remark 3.3.1 and
Lemma 3.3.8, it follows that the reduced band decomposition surfaces for
G --+ S and H ~ T are homeomorphic surfaces with one boundary compo-
nent each. Lemma 3.2.1 now implies that Sand Tare homeomorphic. D
Example 3.3.2. Both the surfaces in Figure 3.26 have one boundary component.
Moreover, when the b-band is deleted from both, the resulting surfaces are
obviously homeomorphic. It follows from Lemma 3.3.8 that the two surfaces
illustrated are homeomorphic, even though this might not be immediately obvious.
The Classification of Surfaces 129
a
a b
The Corollary implies that the imbedding surfaces induced by the rotation systems
are homeomorphic.
by Theorem 3.3.5, that the resulting imbeddings G' ~ S' and H' ~ T' have
one face each.
The surfaces S' and T' are both nonorientable, since the original im-
beddings G ~ S and H ~ T had at least one orientation-reversing loop each
besides the deleted loops d and e, respectively. Since x(G' ~ S') =
x(G ~ S) + 1 and x(H' ~ T') = x(G ~ T) + 1, it follows that x(G' ~ S')
= x(H' ~ T'). Moreover,
#Ec, + #Ew = ( #Ec- 1) + ( #EH- 1) = #Ec + #EH- 2 = n- 1 < n
By the induction hypothesis, the surfaces S' and T' are homeomorphic, and
x(G' ~ S') = x(H' ~ T') ~ 1
The corollary to Lemma 3.3.8 now implies that the surfaces S and T are
homeomorphic. Moreover,
X ( G ~ S) = X ( H ~ T) = X ( G' ~ S') - 1 ~ 0 ~ 1. D
Theorems 3.3.3 and 3.3.4 and the corollaries to Theorems 3.3.7 and 3.3.9
justify the definition of the "Euler characteristic" x(S) of a surface as the
value of the formula #V- #E + #F for any cellular imbedding. In particu-
lar, x(Sg) = 2 - 2g and x(Nk) = 2- k.
3.3.6. Exercises
1. For each of the three surfaces in Figure 3.27, decide whether it is
orientable, calculate its Euler characteristic, and state to which of the
standard surfaces S0 , S 1, ••. or N1, N2 , ... it is homeomorphic.
2. List the faces of the imbedding obtained by edge-deletion surgery on the
edge uv in the imbedding of Figure 3.17. List the faces obtained when
the surgery is on edge vw instead.
3. Let e be an edge of the imbedding G ~ S whose dual edge e* is not a
loop. Let the imbedding H ~ T be the result of surgery on e. Show that
the dual imbedding H* ~ T* can be obtained from the imbedding
G* ~ S* by contracting the edge e*.
The Oassification of Surfaces 131
- b
--
a
~
a
·tOt· ct
tJ ~c o~
- b ';;'.
b
~
/a
Figure 3.27. Three surfaces.
~~
/c
a
u. . .. de .. .
v. . .. e .. .
Give the rotation at vertices u and v after sliding the edge d along the
edge e, first when e is type 0, then when e is type 1.
10. Use your answer to Exercise 9 to give a sequence of edge slides that turns
the one-vertex imbedding with rotation ababc 1c1 into the one-vertex
imbedding with rotation a 1a 1b1b1c1c1• (For example, sliding b along c
gives ab 1ac1b1c1 and then sliding c along a gives ab1ac 1c1b1.)
11. The "disk sum" of two connected, closed surfaces S and T, denoted
So T, is the surface obtained by removing the interior of a disk from
132 Surfaces and Graph Imbeddings
both S and T and then identifying the two resulting boundary compo-
nents (see Figure 3.28). Prove that x(S o T) = x(S) + x(T) - 2.
12. Prove that the disk sum defined in Exercise 11 is an associative operation
on closed, connected surfaces. What is its unit element? Is the operation
commutative?
13. Define the connected surface S to be prime if S is not the sphere and if
S = RoT implies that either R or T is a sphere. What are the prime
surfaces? Give two different factorizations of N3 into primes. Do orienta-
hie surfaces have unique factorizations?
14. Prove that if the surface S contains k disjoint Mobius bands, then
x(S) ~ 2- k. (Hint: Use the fact that if the surface S contains a
Mobius band, it can be factored into S' o N1.) Conclude that for any
nonorientable surface S, x(S) = 2 - k if and only if S contains k
disjoint Mobius bands but no more.
15. Explain how to imbed a graph in the surface N2 k+l so that there exists
an edge e on which surgery yields the orientable surface Sk. (Hint: See
Exercise 13.) What is the minimum number of edge-deletion surgery
steps required to make the surface N2k orientable, for k ~ 1?
according to the interpolation theorem, the genus range GR(G) is the interval
[y(G), YM(G)].
Analogously, the "crosscap range" CR(G) is defined to be the set of
numbers k such that the graph G can be cellularly imbedded in the surface
Nk. The crosscap number y(G) is the minimum value in this range. The
maximum number in CR(G) is called the "maximum crosscap number" of G,
and is denoted YM(G). There is also an interpolation theorem for nonorienta-
ble surfaces, from which we infer that CR(G) = [y(G), YM(G)].
Because of the interpolation theorems, computations of the limiting values
of the imbedding ranges have received great emphasis. In general, the genus
and the crosscap number of a graph are difficult to compute, although they
have been determined for various special classes of graphs. By way of contrast,
the maximum genus is thoroughly tractable, mainly because of its character-
ization by Xuong (1979). The time needed for computation of the maximum
genus is bounded by a polynomial function of the number of vertices, as
proved by Furst, Gross, and McGeoch (1985a).
In this section we are concerned primarily with establishing the interpola-
tion theorems, with some relationships between genus and crosscap number,
and with computing the maximum crosscap number and maximum genus.
Some general statistical questions about the set of all imbeddings will also be
raised.
Example 3.4.1. The three pure rotation projections illustrated in Figure 3.29
correspond to mutually adjacent imbeddings, since edge-deletion surgery on the
edges in the "nine o'clock" position would make them all equivalent. The
left and middle imbedding surfaces visibly have genus 1. The third surface has
genus 2.
134 Surfaces and Graph Imbeddings
Proof Let L 1 and L 2 be list formats for any two rotation systems of G.
Of course, L 2 may be obtained from L 1 by permuting the entries in the
various rows. Any such permutation may be accomplished by a sequence of
steps that moves one edge symbol at a time. The consecutive list formats in
such a sequence represent adjacent imbeddings, by Remark 3.3.1. Thus, there
exists a sequence of adjacent cellular imbeddings of the graph G starting with
one in the surface of genus y(G) and ending with one in the surface of
maximum genus YM(G). Since adjacent imbedding surfaces differ by at most
one in genus, the conclusion follows. D
i. The two sides of e lie in separate faces f 1 and f 2 • Then deleting the
fcband and the f 2-band leaves two holes, which are merged into a
single hole by twisting the e-band. Since the union of the e-band and
The Imbedding Distribution of a Graph 135
Figure 3.30. Case i of edge-twisting surgery goes left to right, whereas case ii goes right to left.
The 2-band boundaries are thickened.
the new 2-band used to close the hole is a Mobius band, the resulting
surface is nonorientable, regardless of whether the starting surface is
orientable.
ii. Both sides of e lie on the same face f. Then deleting the /-band leaves
one hole, which is changed to two holes by twisting the e-band.
Theorem 3.4.2. Let G--+ S be a cellular graph imbedding, and let e be an edge
of the graph. Let F be the set of faces of G --+ S, and let F' be the set of faces of
the imbedding G --+ S' obtained by edge-twisting surgery at e. Then in case i,
#F' = #F- 1, and x(S') = x(S) - 1; in case ii, #F' = #F + 1, and
x(S') = x(S) + 1. D
Edge-twisting surgery yields not only a proof that there are no gaps in the
crosscap range but also an easy formula for the maximum crosscap number.
The "Betti number" (also called the "cycle rank") of a connected graph G,
denoted {J(G), is defined by the equation {J(G) = 1 - #V + #E, and it
equals the number of edges in the complement of a spanning tree for G.
Proof If G is a tree, then CR( G) is-the degenerate interval [0, 0]. Otherwise,
starting with an imbedding G -+ N into a surface that realizes the crosscap
number, one performs a succession of edge-twisting surgeries. At each stage,
one selects an edge that lies in two faces of the imbedding for that stage. Thus,
in YM(G)- y(G) steps, one has obtained a cellular imbedding into every
surface N, such that r lies in the interval [y(G), YM(G)]. D
The Betti number of a graph also gives an upper bound for the maximum
genus of a graph, namely, YM(G).::;; l/3(G)/2J with equality occurring only if
the graph G has an orientable imbedding having one or two faces. However,
unlike the maximum crosscap number, this upper bound for the maximum
genus is not always achieved. One might hope that the Betti number also
provides a useful upper bound on the minimum genus. Indeed, Duke (1971)
once conjectured that y( G) .::;; l f3{ G) 14J. Many counterexamples were found
to Duke's conjecture. Milgram and Ungar (1977) found perhaps the most
discouraging result: given any £ > 0, there are infinitely many graphs G such
that y(G)//3(G) > ! -£.In other words, we cannot obtain an upper bound
for y(G) of the form cf3(G) that is any better than {3(G)/2.
#CR{G) ~ 2#GR(G)- 2
The Imbedding Distribution of a Graph 137
# CR( G) = yM (G) - Y( G) + I
~ 2yM(G)- {2y{G) +I)+ 1
= 2(yM(G)- y(G) + 1)- 2
= 2#GR(G)- 2 o
Theorem 3.4.6. The genus of the complete graph K 1 is I, and the crosscap
number of K 7 is 3.
1 2
Figure 3.31. An imbedding of the complete graph K 7 in the torus.
138 Surfaces and Graph lmbeddings
It follows from Theorem 3.4.5 that Y(K 7 ) ~ 3. Thus, to complete the proof
that Y( K 7 ) = 3, it is sufficient to show that K 7 has no imbedding in N2 • By
the invariance of Euler characteristic, if K 7 did have an imbedding in N2 ,
there would have to be 14 faces. Since K 1 has no 1-cycles or 2-cycles, each of
the 14 faces would have to be three-sided. Therefore, suppose that K 7 ~ S is
an imbedding such that every face is three-sided. It is now to be proved that
the surface S is orientable, thereby establishing that K 7 is not imbeddable in
the Klein bottle N2 •
Regarding K 7 as a Cayley graph for the cyclic group f!E7 , we label its
vertices 0, 1, ... , 6. It may be assumed, without loss of generality, that the
rotation at vertex 0 is
0. 123456
The edges incident on vertex 0 form a spanning tree, and we locally orient the
vertices so that all six of these edges are type 0. The left side of Figure 3.32
shows a clockwise partial rotation projection "centered" at vertex 0.
Since each of the six faces incident on vertex 0 is three-sided, we may infer
from the given rotation at vertex 0 and the Face Tracing Algorithm the
following partial rotation system (in vertex format):
0. 123456
1. 206 00 0
2. 301 0 00
3. 402 000
4. 503 000
5. 604 000
6. 105 0 0 0
If the edge 12 had type 1, then the sequence 10, 02,21 in an execution of the
Face Tracing Algorithm would be followed by some edge 1x, x -:/= 0, thereby
creating a face that is not three-sided. Thus, the edge 12 must have type 0.
3 or4
6 or 1
Figure 3.32. Partial rotation projections for putative triangular imbeddings of K 1 .
The Imbedding Distribution of a Graph 139
Similarly, each of the edges 23, 34, 45, 56, and 61 must have type 0, as
indicated. We shall see that there are but two ways to extend this partial
system to a rotation system that has only three-sided faces, and that all of the
other edges in the complete system also have type 0, thereby proving that the
surface is orientable.
The edge 16 must lie in a second triangle besides 016. The third vertex of
this triangle must be 3 or 4, since the edges 56 and 12 are already in use. By
similar reasoning applied to the five edges 12, 23, 34, 45, and 56, we arrive at
the partial rotation projection shown on the right of Figure 3.32, where the
vertex 0 and its incident edges are now omitted, to simplify the picture.
Suppose that 3 is the third vertex of the second triangle containing the edge
16. Then the third vertex of the second triangle containing the edge 56 must be
2. Continuing in this manner around the figure, we obtain the partial rotation
projection shown at the left of Figure 3.33. The remaining free edge at each
vertex is now determined, thereby yielding the following complete rotation
system, whose projection is illustrated at the right of Figure 3.33:
0. 123456
1. 206354
2. 301465
3. 402516
4. 503621
5. 604132
6. 105243
If the edge 13 had type 1, then the successive edges 61 and 13 would be
followed by the edge 35 in the Face Tracing Algorithm execution, which
blocks the completion of a three-sided face. Thus, the edge 13 must have type
0. Similarly, the edges 24, 35, 46, 51, and 62 must have type 0. The only
3 3
6 6
Figure 3.33. Extensions of the partial rotation proJection of Figure 3.32.
140 Surfaces and Graph lmbeddings
remaining edges of undetermined type are now 36, 25, and 14. However, in
order for the face tracing sequence 61, 13, 36 to continue with 61, rather than
64, it is necessary that the edge 36 have type 0. Similarly, the edges 25 and 14
have type 0.
If the third vertex of the second triangle containing the edge 16 were 4,
instead of 3, then the same arguments would again yield a pure rotation
system. It follows that the complete graph K 7 has no imbedding in the Klein
bottle, and that Y( K 7 ) = 3. D
in G. Then the bands for the vertices and edges of C form a topological
neighborhood N( C) that is homeomorphic to an annulus, if C is orientation
preserving, or to a Mobius band, if C is orientation reversing. In either case,
the cycle C is a topological loop that runs around the middle of the
neighborhood.
Suppose that we cut the surface S along the cycle C, that we regard the
0-band for each vertex v; as having been split into a v;-band and v;' -band, and
that we regard the 1-band for each edge e1 as having been split into an e)-band
and an ej'-band. This is illustrated in Figure 3.34.
Suppose further that we cap off the hole or holes created by cutting on C
with one or two new 2-bands, as required. The graph imbedding G' -+ S'
consistent with the resulting band decomposition is said to have been obtained
by "splitting" the imbedding G -+ S "along the cycle C ". An easy counting
e; ~
u \. .-...;_~(
--(0 \-s--- Vf+ I
to 5
Example 3.4.2 (Auslander et al. 1963). For each positive integer n, the graph
A. is formed from n + 1 concentric cycles, each of length 4n, plus some
additional edges, as shown in Figure 3.35. There are 4n 2 "inner" edges that
connect the n + 1 cycles to each other and 2 n "outer" edges that adjoin
antipodal vertices on the outermost cycle.
Suppose that the n + 1 concentric cycles of the graph A. and the inner
edges are imbedded on a disk so that the vertices of the outermost cycle
appear equally spaced on the boundary of the disk. Suppose also that a
rectangular sheet with 2n lines from top to bottom has its left and right sides
identified so that it becomes a Mobius band. If the boundary of the Mobius
band is identified to the boundary of the disk so that the 2n lines become
images of the outer edges of A., then the result is an imbedding A.~ N1 •
Figure 3.36 shows that the graph A 1 contains a subdivision of K 3 3 • Since A.
contains a subdivision of A1 , for all n ~ 1, it follows that y(A.) ~ 1.
It is clear from Figure 3.36 that y(A 1) = 1. Suppose, by way of induction,
that y(A._ 1 ) ~ n - 1. Then let C be the innermost cycle of the graph A., as
it appears in Figure 3.35, and let A. ~ S be an orientable cellular imbedding.
First, assume that the cycle does not separate the surface S. Then let the
imbedding A~~ S' be the result of splitting the imbedding A.~ S along the
cycle C. By Theorem 3.4.7, y(S) = y(S') + 1. Since the graph A._ 1 is a
subdivision of a subgraph of the graph A., it follows from the induction
hypothesis that y(S') ~ n - 1. Thus y(S) ~ n.
On the other hand, one might assume that the cycle C does separate the
surface S. Since the graph A.- C is connected, the cycle C must bound a
142 Surfaces and Graph lmbeddings
to 5
to 7
face of the cellular imbedding An~ S. Now suppose that images of all the
edges on the n other concentric cycles of the graph An are deleted, and that
every resulting vertex of valence 2 is smoothed over. The result is a possibly
noncellular imbedding Yn ~ S. If every noncellular face of Yn ~ Sis replaced
by a 2-cell, then one obtains the imbedding Yn ~ T whose rotation projection
is shown in Figure 3.37. Obviously, y(T) 5: y(S).
An elementary face-tracing argument shows that the imbedding Yn ~ T
has only two faces, so that x(T) = 4n - 6n + 2 = 2- 2n, which implies that
y(T) = n. It follows that y(S) ~ n.
We conclude from this two-pronged inductive argument that y(An) ~ n.
Since there is only one outer face in the imbedding of An given by the rotation
projection of Figure 3.35, it follows from a calculation of Euler characteristic
that the imbedding surface has genus n. Thus y(An) = n.
Lemma 3.4.8. Let d and e be adjacent edges in a connected graph G such that
G - d - e is a connected graph having an orientable one-face imbedding. Then
the graph G has a one-face orientable imbedding.
Proof Let V(d) = {u, v}, and let V(e) = {v, w}. First, extend the one-
face imbedding (G- d- e)-+ S to a two-face imbedding (G- e) -+ S by
placing the image of d across the single face. Of course, the vertex v lies on
both faces. Thus, if one attaches a handle from one face of ( G - e) -+ S to
the other, one may then place the image of edge e so that it runs across the
handle, thereby creating a one-face imbedding G-+ S'. 0
Example 3.4.3. Consider the complete graph K 5 as a Cayley graph with vertices
0, 1, 2, 3, 4. Then the spanning tree T with edges 01, 12, 23, and 34 has a
one-face imbedding in the sphere. By Lemma 3.4.8, the graph T' = T + 04 + 42
also has a one-face imbedding, as does the graph T" = T' + 20 + 03, as does
K 5 = T" + 31 + 14. Thus YM(K 5 ) = 3.
Lemma 3.4.9. Let G be a connected graph such that every vertex has valence at
least 3, and let G have a one-face orientable imbedding G -+ S. Then there exist
adjacent edges d and e in G such that G - d - e has a one-face orientable
imbedding.
Example 3.4.4. On the left in Figure 3.38 is a spanning tree T for a graph G
such that ~(G, T) = 3. On the right is a spanning tree T' for the same graph G
such that ~( G, T') = 1. Since the edge complement of any spanning tree for this
graph has 1 - 7 + 11 edges, an odd number, the deficiency of any spanning tree
must be at least one. Therefore ~(G)= ~(G, T') = 1.
144 Surfaces and Graph Imbeddings
Figure 3.38. Two spanning trees for the same graph, one of deficiency 3, the other of de-
ficiency 1.
Lemma 3.4.10. Let T be a spanning tree for a graph G, and let d and e be a
pair of adjacent edges in G- T. If ~(G- d- e, T) = 0, then ~(G, T) = 0.
Lemma 3.4.11. Let G be a graph other than a tree, and letT be a spanning tree
such that ~(G, T) = 0. Then there are adjacent edges d and e in G- T such that
~(G- d-e, T) = 0.
Proof As the basis for an induction, one observes that if #EG = 0, then
both clauses of the conclusion are trivially true. Next, assume that the
conclusion holds for any graph with n or fewer edges, and let G be a graph
with n + 1 edges.
As a preliminary, suppose that G has a vertex v of valence 1 or 2, and let
G' be the graph obtained by contracting an edge incident on vertex v. Then
obviously, the graph G has a one-face orientable imbedding if and only if the
graph G' does. Also, ~(G) = 0 if and only if ~(G') = 0. Since tfie graph G' has
one edge less than the graph G, it follows from the induction hypothesis that
G' has a one-face orientable imbedding if and only if ~(G') = 0. The conclu-
sion follows immediately.
In the main case, every vertex of G has valence 3 or more. Suppose first
that G has a one-face orientable imbedding. By Lemma 3.4.9, there exist
adjacent edges d and e in G such that G- d-e has a one-face orientable
imbedding. It follows from the induction hypothesis that ~( G - d - e) = 0,
The Imbedding Distribution of a Graph 145
o--o
Figure 3.39. Two graphs that have deficiency 2.
Example 3.4.5. The dumbbell graph G on the left of Figure 3.39 has only one
spanning tree, from which it follows that ~(G)= 2. It is easily verified that the
removal of any pair of adjacent edges from the graph G' on the right of Figure
3.39, followed by a sequence of edge contractions to eliminate vertices of valence
1 or 2, yields the dumbbell graph. Therefore by Lemma 3.4.9, it follows that
{3(G') = 2 also.
lbeorem 3.4.13 (Xuong, 1979). Let G be a connected graph. Then the mini-
mum number of faces in any orientable imbedding of G is exactly ~(G) + 1.
Example 3.4.6. Let T be the tree in the complete graph K n consisting of all
edges incident on a particular vertex. Then
or equivalently,
if n =1 or 2 modulo4
if n =0 or 3 modulo 4
Gross and Furst (1985) raise the problem of determining exactly how many of
these imbeddings lie in each surface of the genus range. From such a
determination, one would immediately know the minimum genus and the
maximum genus. Furst, Gross, and Statman (1985b) combined topological
methods and enumerative techniques to obtain a genus distribution formula
for two infinite classes of graphs with expanding genus ranges. Gross, Robbins,
and Tucker (1986) calculated the genus distributions of bouquets of circles,
with the aid of a formula of Jackson (1986).
Another problem raised by Gross and Furst (1985) is to compute the
distribution of face sizes over all imbeddings of a graph, which appears to be
harder than computing genus distributions. They introduce an extensive
hierarchy of invariants of the imbedding distribution of a graph, in which the
genus distribution and the face-size distribution lie near the bottom. Higher
invariants keep track of the distribution of face sizes within each imbedding
and the relationships between the faces.
The set of imbeddings of a graph has various statistical properties of
interest. For instance, all the genus distributions mentioned above are strongly
unimodal. Also, Gross and Klein (1986) classified the graphs of average genus
less than or equal to 1.
Little else is known about the number of imbeddings of a given graph in a
given surface, except when the imbedding surface has low genus. The proof of
Theorem 3.4.6 shows that the complete graph K 7 has 2 · 5! different im-
beddings in the torus (once a rotation at vertex number 1 is fixed, there are
only two choices for the imbedding). By Whitney's theorem (Theorem 1.6.6),
3-connected planar graphs have "only one" equivalence class of imbeddings in
the plane. Negami (1983, 1985) has investigated the uniqueness of imbeddings
in the torus and the projective plane.
3.4.8. Exercises
1. Show that the sequence of rotation projections for the complete graph K 5
given in Figure 3.40 are adjacent in pairs from left to right and that the
associated surfaces have genus 1, 2, and 3 (again from left to right).
lbeorem 3.5.1. Let G-+ S be an imbedding such that every boundary walk is a
cycle, and such that the only boundary walks passing through both endpoints of
an edge are the two boundary walks containing that edge. Then every aqjacent
imbedding has larger genus.
Proof Let e be any edge of the graph G, and let G' -+ S' be the
imbedding obtained by deletion surgery on edge e. This imbedding has one
less face than the original imbedding, since the two faces containing edge e
have been joined into one. Moreover, no face of the imbedding G' -+ S' passes
through both endpoints of edge e, except the joined face. Therefore, when the
edge e is reattached in a different way, it must form a bridge joining two more
faces. Hence, any adjacent imbedding has fewer faces and larger genus. D
Example 3.5.1 (Gross and Tucker, 1979b). It is easy to verify that the toroidal
imbedding on the left of Figure 3.41 satisfies the hypothesis of Theorem 3.5.1.
Thus, every adjacent imbedding has larger genus. However, the given graph has
the planar imbedding given on the right. It follows that the imbedding at the left
is a local minimum that is not a global minimum.
Figure 3.42. Any amalgamation of two graphs is imbeddable into a disk sum of any two of their
respective imbedding surfaces.
the amalgamation vertex. Next, one chooses an analogous disk on the surface
Sk. Then, one deletes the interiors of both disks and identifies the resulting
surface boundaries so that the graphs H and K are amalgamated, as il-
lustrated in Figure 3.42. As we have mentioned in Exercise 3.3.11, this
composition of surfaces is called a "disk sum".
For this simplified type of graph amalgamation, we often use the notation
H *v K, where v is the vertex of amalgamation, to represent the resulting
graph. From the above discussion, it is obvious that
The so-called BHKY theorem (for Battle, Harary, Kodama, and Youngs)
asserts the nonobvious fact that
such that the edge a is inserted between edges c and d, so that vertex v has the
resulting rotation
v. b ... cad
and all other rotations remain the same. Then y(S') 5: y(S).
Theorem 3.5.3 )The BHKY Theorem (Battle, Harary, Kodama, and Youngs,
1963)). y(H•vK) = y(H) + y(K).
We claim that the edges d and d' must lie in the same H-segment. If not, then
suppose that the lengths of the H-segments containing d and d' are n and n',
respectively. Without loss of generality, we shall assume that n 5: n'. Apply
Lemma 3.5.2, with edges d, e, e', and d' filling the roles of a, b, ~' and d,
respectively. Evidently, the adjacent imbedding with the modified rotation
v. e ... e'dd'
(and no other changes) would have the same imbedding surface-lower genus
is impossible, since S is a minimum-genus imbedding surface-but a larger
Algorithms and Fonnulas for Minimum Imbeddings 153
because n ~ n'. Thus, the edges d and d' are in the same H-segment.
Similarly, the edges e and e' are in the same K-segment. Therefore, we have
established an assertion that there is only one H-segment and only one
K-segment.
We must still prove that a closed arc C through vertex v from corner de to
corner e'd' in the interior of their common face must separate the imbedding
surface. However, if not, then closed arc C would lie on a handle of surfaceS.
If the surface S were cut open along C and reclosed with two disks, the result
would be an imbedding of the disjoint union of the graphs H and K on a
surfaceS' of genus y(S) - 1. Such an imbedding would be noncellular, since
any region that contained both a closed H-walk and a closed K-walk in its
boundary would fail to be simply connected. By the construction of S', no
such region could arise. It follows that the closed arc C separates the surface
S. Since graph H lies on one side of the separation, that side must have genus
at least y(H). Similarly, the other side must have genus at least y(K). It
follows that
lbeorem 3.5.4. If the blocks of graph G are G1, ... ,Gn, then
Proof One uses a simple induction on the number of vertices of the graph
G, while applying the fact that a subgraph His a block graph G if and only if
G = H *v K for some vertex v and some subgraph K (of course, K itself may
have several blocks). 0
More general types of amalgamations have also been studied. The situation
here is much more complicated, as the following example illustrates.
Example 3.5.2. The rotation projection shown in Figure 3.44 gives an im-
bedding in a torus of an amalgamation of K 5 with K 5 along an edge. Thus the
genus of this graph is I, not the seemingly obvious value of 2.
The results known for other amalgamations are all of the form y( H • 1 K)
= y(H) + y(K) + B, where the parameter B is given a possible range of
values, and each of these values does occur for some pair of graphs Hand K.
Alpert (1973) shows that when the pasting function f identifies an edge of H
with an edge of K, then B = 0 or -1. Stahl (1980) and Decker et al. (1981)
prove that B = 1, 0, or -1 when f identifies copies of K/. Stahl also shows in
the same paper that B = 2, 1, 0, -1, or - 2, when the pasting function f
amalgamates two copies of K 3c.
In general amalgamations, the range of B values is much greater and
determining when B takes on a particular value is quite complicated [see, for
example, Decker et al. (1985)]. The best general theorem would be a bound on
B that depends only on the number k of vertices being identified by the
amalgamation. An upper bound for B if k = 1 is easily obtained and is
achieved for some amalgamations when k = 2. Archdeacon (1986) has shown,
however, that there is no lower bound for B even when k = 3; he constructs a
sequence of n-vertex graphs Gn and 3-vertex amalgamations Hn of Gn with
itself such that y(Hn) = 2y(Gn) + n.
Proof Let g denote the value of the right-hand side of the equality and let
G = H•vK. To see that y(G) ~ g, use disk sums and the general fact that
y(H) ~ 2y(H) + 1 (and similarly forK).
Conversely, to show that y(G) ~ g, it suffices to prove, as in Theorem 3.5.3,
that there is an imbedding of the graph G in the nonorientable surface of
minimum crosscap number such that the rotation at vertex v has one H-seg-
ment and one K-segment. The details of the proof of Theorem 3.5.3 can be
used again. The only difficulty lies in the nonorientable version of Lemma
3.5.2. The type of the edge d might have to be changed when it is reattached
at vertex v, in order not to decrease the number of faces. (Certainly one of the
two possible types for edge d will be wrong: we want the sides of edge d in
the new imbedding to belong to separate faces, yet an extra twist in an edge
whose sides belong to separate faces joins the faces.) One might fear that
changing the type of edge d could make the new imbedding surface orientable.
However, in the present context, the sides of the edge d belongs to distinct
faces in the original imbedding. As we observed in Section 3.4 in the
discussion of extra twists, changing the type of edge under such a circumstance
always yields a nonorientable surface. D
1 2 1 2 1 2
4 3 4 3 4 3
f 1 2 1 2 1 2 t
3 4 3 4 3 4
Figure 3.45. -
Some disjoint patchworks in a toroidal imbedding of C4 X C6 .
"residual". A patchwork is called "even" if all of its faces have even size, and
it is called "quadrilateral" if all of its faces have size 4.
Example 3.5.5. On the left of Figure 3.46 is an edge-coloring of the cycle graph
C6 using two colors, both of which are saturated. On the right is an edge-coloring
of the graph C5 using three colors, none of which is saturated.
2
Figure 3.46. Edge-colorings of the graphs C6 and C5 .
Proof Let 1, 2, ... , r denote the colors of the edge-coloring of the graph H
with the colors that are not saturated listed first. Let P 1, P2 , ••• , P, be the
disjoint even patchworks of the imbedding G ~ S with the quadrilateral
patchworks listed first. We construct the desired imbedding for the graph
G X H by giving its rotation at each vertex ( u, v ). Recall that edges incident
to vertex (u, v) are of the form (d, v) where d is an edge of the graph G
incident on vertex u or of the form ( u, e) where e is an edge of the graph H
incident on vertex v. For convenience, we call the former G-edges and the
latter H-edges. We call the edges d and e their projected edges. The G-edges
are to appear in the rotation at vertex ( u, v) in the same order as their
projected edges in the rotation at vertex u for the imbedding G ~ S. The
H-edges are then interspersed as follows: if the projection of an H-edge is
colored i in the edge-coloring of the graph H, then that edge is placed between
the two G-edges whose projections form the corner of the face from patchwork
P; at vertex u. The type of every G-edge is the same as its projected edge while
every H-edge is given type 1. It is best to view this imbedding as #V(H)
copies of the imbedding G ~ S placed at different levels with "tubes" joining
corresponding faces at different levels where needed for H-edges. Figure 3.47
shows how a pair of H-edges having the same projection fit into four corners
of corresponding faces from patchwork P;. A pair of corresponding faces of
Figure 3.47. Part of a rotation projection of the imbedding for G X H; on the left, before
H-edges are added and, on the right, after.
158 Surfaces and Graph Imbeddings
The important part of the conclusion of Theorem 3.5.6 is that the con-
structed imbedding has all faces quadrilateral. The number of disjoint patch-
works in the imbedding is useful for repeated cartesian products.
Proof The graph en, X en 2 has an imbedding in the torus that has all
faces quadrilateral and four disjoint patchworks (see Example 3.5.4), whenever
n 1 and n 2 are even. Repeated application of Theorem 3.5.6 gives an all-
quadrilateral imbedding of the graph G, orientable in case a and nonorientable
in case b. Since the girth of the graph G is greater than three, the equations
follow from Euler's equation. 0
Proof The proof is like that for Theorem 3.5.7; use the quadrilateral
imbedding of Q2 = e4 to get started. 0
Theorem 3.5.11 (Vizing). Any graph with maximum valence r can be edge-col-
ored with r + 1 colors or less.
Proof Omitted. D
3.5.6. Exercises
1. Prove that given any integer m, there is a graph having an edge whose
contraction reduces the genus of the graph by m. (Hint: Any n-vertex
graph reduces to a one-vertex graph by a sequence of n - 1 edge
con tractions.)
Algorithms and Fonnulas for Minimum lmbeddings 161
If some edge incident on u has type 1 in the base imbedding (i.e., if the local
orientations at its endpoints are inconsistent), then the corresponding edge in
the derived graph is also assigned type 1. The surface sa induced by this
rotation system on the derived graph is called the "derived surface", and the
imbedding oa --+ sa is called the "derived imbedding".
This definition of the derived imbedding translates immediately into a band
decomposition. Each 0-band of the base imbedding lifts to #.9.1 different
0-bands in the derived imbedding, each of which has the same instructions (up
to .1<1-coordinates) for attaching 1-bands. Moreover, each 1-band of the base
imbedding lifts to #.9.1 different 1-bands, each with the same "twist" as the
base 1-band.
Uo. a 01 co do
u 1• a 11 c1 d1
Do. b1 do c1
v 1• bo d1 Co
Wo. bo e 11 a 01 e01
w1. b1 e0 1 a 11 e 11
0 1
Figure 4.1. Rotation projections for an imbedded voltage graph (bottom) and its derived
imbedding (top).
The derived imbedding has four faces and these boundary walks:
ao eo bl cl al el ho co
bl- -
do el ao
dl b0 - eo al
co d1 - cl d0-
Theorem 4.1.1 (Gross and Alpert, 1974; Gross, 1974). Let C be the boundary
walk of a face of size k in the imbedded voltage graph (G ~ S,a). If the net
voltage on the closed walk C has order n in the voltage group .J/1, then there are
#.Jiljn faces of the derived imbedding oa ~ sa corresponding to the region
bounded by C, each with kn sides. 0
In Example 4.1.1 the base region aebc- has length 4 and net boundary
voltage of order 2, so it lifts to one eight-sided face a 0 e 0 b0 c 1 - a 1 e 1 b 1 c0 - ,
whose boundary walk covers the boundary walk of a e b c- twice. The base
region db- e a- has length 4 and net boundary voltage of order 1, so it lifts to
two four-sided faces, d 0 b 1 - e 1 a 0 - and d 1 b0 - e 0 a!, each of whose boundary
walks covers the boundary walk of db- e a- once. The base digon c d- has
net boundary voltage of order 2, so it lifts to one four-sided face c0 d 1 - c1 d 0 - .
We emphasize the principle that faces of the derived imbedding are
computed in practice from the net voltages on boundary walks of the base
imbedding, not by the lengthy task of face tracing in the derived rotation
system. Another example further illustrates this point.
1 1 1
1 2
1 4
1
~
Figure 4.2. A voltage graph imbedded in the torus, with voltages in the cyclic group .2"6 .
166 Imbedded Voltage Graphs and Current Graphs
,o,
v
Figure 4.3.
1 v
An imbedding 8 2 -+
I 1I r r
v2 v3 v4 v0 v1
i
v2
S1 with voltages in .2'5 and the derived imbedding K 5 -+ S1 .
Proof By Theorem 4.1.1, each face of the base imbedding lifts to n faces
of the derived imbedding. The derived graph always has n times as many
vertices and edges as the base graph. Thus
(1 4 6)(2 7 3 5) (1 2 3)(4 56 7)
(1 7 5)(2 6 3 4)
Figure 4.4. A permutation voltage graph imbedded in the sphere.
Theorem 4.1.3. Let C be the boundary walk of a k-sided face in the imbedded
permutation voltage graph (G ~ S, a)n· Suppose that the net permutation
voltage on C has cycle structure ( c 1 , c 2 , • . . , en)· Then there are
cl + c2 + ... +en faces of the derived imbedding oa
~sa corresponding to the
face C, including for j = 1, 2, ... , n exactly cj faces with kj sides.
x(sa) = 7- 21 + 8 = -6
4.1.5. Orientability
Let (G ~ S, a) be an imbedded voltage graph. If the surfaceS is orientable,
then the orientability algorithm of Section 3.2 enables us to construct an
equivalent imbedding in which every edge of G has type 0. Therefore, we may
assume that every edge of the derived imbedding also has type 0, from which it
follows that the derived surface is orientable. If the base surface S is
nonorientable, the situation is not so clear. On the one hand, the derived
surface in Example 4.1.1 is nonorientable, since the closed walk a 0 b0 c0 - is
type 1. On the other hand, the derived surface might be orientable, as in the
following example.
0
Figure 4.5. Rotation proJections of a nonorientable imbedded .2"4 -voltage graph and its orienta-
hie derived imbedding.
The derived imbedding, shown on the right, has two four-sided faces, and its only
cycle consists of four edges of type 1. Thus the derived surface is the sphere; in
particular, it is orientable. If the voltages shown were in f!E3 instead, then the
derived surface would have had a cycle with an odd number of type-1 edges and
would have been nonorientable.
From this example, one may correctly suspect that the derived surface of an
imbedded voltage graph (G ~ S, a) is nonorientable, if and only if some
type-1 closed walkinG has a net voltage of odd order. In fact, one only need
check for net voltages of order 1.
Proof Let W' be a type-1 closed walk in the derived surface sa, and let
p:Ga ~ G be the natural projection. Then the image W = p(W') of the
sequence of edges in W is a type-1 closed walk, and, by Theorem 2.1.2, the net
voltage on the walk W is the identity. Conversely, let W be a type-1 closed
walk in the base imbedding G ~ S whose net voltage equals the identity.
Then the lift p- 1(W) consists of #..#copies of the closed walk W, each of
which is type 1. D
Clearly one cannot check the net voltage of every one of the infinite number
of type-1 closed walks to see if it equals the identity. Some finite collection of
"basic" cycles is needed, as in the orientability algorithm of Section 3.2, or as
in the method described in Section 2.5 for finding generators of the local group
.Jil(u). The difficulty is that the voltages assigned to basic type-0 cycles have
just as much influence over the orientability of the derived imbedding as the
voltages assigned to basic type-1 cycles. The globalness is illustrated by the
following example.
Example 4.1.6. Let B2 ~ S be any imbedding of the bouquet of two circles such
that one loop d of B 2 is type 1 and the other loop e is type 0. Let a be the
f!E2-voltage assignment given by a( d) = a( e) = 1. Since the only "basic" type-1
The Derived Imbedding 169
cycle d does not have a net voltage of 0, it might appear that the derived surface
sa is orientable. However, the closed walk de is also type I, and its net voltage is
0, since a( e) = 1. Thus the derived surface is, in fact, nonorientable.
Example 4.1.6 indicates that the derived surface is orientable when the net
voltages on type-1 closed walks are in some way "disjoint" from the net
voltages on type-0 closed walks. To make this idea more precise, a base vertex
u is needed so that the local voltage group .Ji.l(u) can be used. Let .9.1°(u)
denote the collection of net voltages on all type-0 closed walks based at vertex
u. Since the set of such walks is closed under the product of walks and the
reversing of walks, the collection .9.1°(u) forms a subgroup of the group .Ji.l(u);
this subgroup is called the "type-0 local (voltage) group".
Proof We will first prove that i and ii are equivalent. To this end, let e be
any type-1, u-based closed walk, and suppose its net voltage is b. We claim
that .Ji.l(u) = .9.1°(u) u b.9.1°(u). In support of this claim, suppose that a E
.9.1 ( u ), so that a is the net voltage on some u-based closed walk D. If D is type
0, then a E .9.1°( u) by definition. If D is type 1, then e- D is type 0, so its net
voltage b -la is in .9.1°( u ). Therefore, a E b.9.1°( u ), and the claim is proved. It
follows that .9.1°(u) has index 1 or 2 in .Ji.l(u). Since the type-1 closed walk e
is arbitrary, this also shows that .9.1°(u) has index 2 if and only if no type-1,
u-based closed walk has its net voltage in .9.1°(u). Thus, i and ii are equivalent.
Now we will show that ii and iii are equivalent. Suppose that the derived
surface sa is nonorientable. Then, by Theorem 4.1.4, there is a type-1 closed
walk e in the base graph G with net voltage equal to the identity. If e is not
already based at the vertex u, then the closed walk Den-, where D is a path
from u to the initial vertex of e, is based at u and still is type-1 with net
voltage equal to the identity. Of course, the identity is in .9.1°(u). Thus, we
have proved that ii implies iii. Conversely, suppose that some type-1, u-based
closed walk e has its net voltage in .9.1°(u). By the definition of the subgroup
.9.1°(u), some type-0, u-based closed walk D has the same net voltage as C.
Thus, en- is a type-I closed walk in the base graph, whose net voltage equals
the identity. It follows from Theorem 4.1.4 that the derived surface sa is
nonorientable. Thus iii implies ii. D
Since each of the three possible index-2 subgroups of ~2 X ~2 contains the net
voltage of a type-1loop, it follows from Theorem 4.1.5 that the derived surface is
nonorientable.
Proof It suffices to show that the subgroup !II is normal in the group .91.
This follows immediately from the observation that, if a rt; !II, then a!JB = ?A a
= .91- !II. D
To verify that this orientability test works, we first recall that the derived
surfaces for voltage assignments a and aT are the same. If their common
derived surface is orientable, then by Theorem 4.1.5, the type-0 local group
s# 0 (u) has index 2 in the local group s#(u), and no type-1 closed walk has a
The Derived Imbedding 171
u 3 u u 3 u
6 6 6 6
3
Figure 4.6. A voltage graph imbedded in the Klein bottle (the original voltages are given on the
left, the T-voltages on the right).
net voltage ind' 0 (u). Because every edge of the tree T is type 0 and has
T-voltage equal to the identity, it follows that every edge e not in the tree T
corresponds to a unique u-based closed walk having the same type as e and
net voltage equal to aT(e). Therefore, d' 0 (u) is the desired subgroup !!J.
Conversely, suppose that the subgroup !!J exists. Then aT assigns every
type-1 edge a voltage not in !!J. By the algebraic lemma (Theorem 4.1.6), it
follows that every type-1 closed walk has a net voltage not in !!J. In particular,
no type-1 closed walk has net voltage equal to the identity. Therefore, by
Theorem 4.1.4, the derived surface is orientable.
Example 4.1.8. Let (G ~ S, a) be the imbedded voltage graph given on the left
in Figure 4.6 where voltages are in the cyclic group ~12 • The base imbedding
surface S is a Klein bottle. The graph G is the same as that of Example 2.5.3,
and the tree T and root vertex u selected are again the same as in Example 2.5.3.
Local orientations are indicated by arrows around vertices. On the right of Figure
4.6 is the same imbedding with the orientation at vertex v reversed so that each
edge in the tree T is type 0. In addition the T-voltages have been given as
computed in Example 2.5.3. The local group d'(u) is {0, 3, 6, 9}. The only
possible 2-index subgroup is {0,6}. It is easily verified that aT(e) E {0,6} if
and only if the edge e is type 0. We conclude that the derived surface sa is
orientable. We also observe that the derived surface sa has three components.
Finally, since two of the boundary walks of the imbedding G ~ S have net
voltage 0 and the other two have net voltage 6, it follows that the derived
imbedding has 36 faces. Therefore, x(Sa) = 36- 84 + 36 = -12, from which
it follows that the derived surface sa consists of three orientable components of
genus 3.
It should be noted that the existence of the index-2 subgroup !!J in the test
described above is equivalent, by Theorem 4.1.6, to the existence of a homeo-
morphism cp: d'(u) ~ ~2 such that cf>(aT(e)) = 0 when e is a type-0 edge
and cf>( aT( e)) = 1 when e is a type-1 edge. This viewpoint is especially
convenient if the local group d'(u) is given by a presentation with generating
set { aT( e) 1 e E EG - T}. There really is no choice for the desired homeomor-
172 Imbedded Voltage Graphs and Current Graphs
z yz
Figure 4.7. A rotation projection for an imbedded voltage graph and a corresponding set of
T-voltages.
phism cf>, since its values on the generating set are determined. It is necessary
to check only that this assignment of Os and 1s to the generators of d(u) does
define a homeomorphism; that is, each relator for the given presentation must
have the value 0.
Example 4.1.9. Let (G ~ S, a) be the imbedded voltage graph given on the left
in Figure 4.7. The voltage group d is a finite group with a presentation of the
form
On the right in Figure 4.7 are the T-voltages for the spanning tree shown in
boldface. If we let s = xy and t = yz, then the local group d(u) has a
presentation of the form
( s, t: s 4 = t 4 = ( st) 3 = 1, ... )
Then the desired homeomorphism cf>: Jli'(u) ~ fl'2 must have cf>(s) = 0 and
cf>(t) = 1. Butthen cf>((st) 3 ) = 3cf>(s) + 3cf>(t) -:F 0. Therefore the derived surface
is nonorientable.
4.1.7. Exercises
1. Let (G ~ S, a) be the imbedded voltage graph whose rotation projection
appears in Figure 4.8a. Find the number of faces of each size in the
derived imbedding and identify the derived surface.
2. Verify that the derived imbedding of the imbedded permutation voltage
graph given in Figure 4.8b is an imbedding of the complete bipartite
graph K 4• 4 in which every face is a quadrilateral.
3. Construct an imbedding of the complete graph K 6 that has two triangles,
two 6-gons, and three quadrilaterals. (Hint: Begin with a bouquet of
three circles imbedded in the sphere. Multiple edges in the derived graph
must be identified.)
The Derived Imbedding 173
(1 2)(3 4)
(1)(2)(3)(4)
(a) (b)
Figure 4.8. Two rotation proJections of imbedded voltage graphs. (a) Ordinary voltages in .2"6 ,
(b) permutation voltages in 94.
(1, 2)
(0, 5) (0, 1)
(0, 4)
(1, 0)
Figure 4.9. A rotation projection of an imbedded voltage graph, with voltages in .2"2 X .2"8 .
For any voltage graph (G, a), the natural graph projection p: Ga-+ G is a
covering projection in the topological sense; that is, every point y E G has a
neighborhood NY such that each component of the lift p- 1(Ny) is mapped
homeomorphically by p onto NY, as described in Section 2.3. For any
imbedded voltage graph (G-+ S, a) one might expect that the natural projec-
tion p: Ga -+ G extends to a map p: sa -+ S with properties similar to those
of p. This is indeed the case, for there is an extension of p that is "almost" a
covering map.
topological argument shows that the fiber size # p - 1( z) is the same for all
nonbranch points z E S. This common value is called the "number of sheets"
of the branched covering. The "deficiency" of the branch point y, denoted
def( y ), is n - # p- 1( y ), where n is the number of sheets of the branched
covering.
Example 4.2.1. Let S and S both be the extended complex plane, that is, the
complex plane with a point at oo adjoined (the resulting surface is homeomorphic,
by stereographic projection, to the sphere). Let p: S ~ S be the function
p(z) = zn, where n ~ 1 is an integer. Since p simply multiplies the argument of
each complex number by n, the effect of pis to wrap the complex plane n times
around the origin. The restricted map p: S- {0, oo} ~ S- {0, oo} is clearly a
covering, sop: S ~ Sis a branched covering. The number of sheets is n, since
any complex number other than 0 and oo has n different nth roots. Both 0 and
oo are branch points of deficiency n - 1, and the corresponding prebranch points,
again 0 and oo, have order n.
Example 4.2.2. Let S be the extended complex plane, and let S be the set of
points {(z, w) E C X qw 2 = z(z- l)(z- 2)} with a point at oo adjoined. It
can be shown that every point of S has a neighborhood homeomorphic to the unit
disk, so that Sis a surface. Let p: S ~ S be the projection p(z, w) = z. Then p
fails to be a local homeomorphism only at the points (0, 0), {1, 0), {2, 0) and oo.
Therefore p is a branched covering with branch points at 0, 1, 2, and oo. We
observe that p is two-sheeted, since the equation w 2 = z(z- l)(z- 2) has two
solutions for w for each value of z other than 0, 1, 2, or oo.
branch point of q lies on G and at most one branch point lies in any face. Then
there is a permutation voltage assignment a in .9;, such that q is a natural
projection S" --+ S.
Example 4.2.1 Revisited. The branched covering of the extended complex plane
given by p(z) = zn could be obtained by extending the natural projection of the
permutation voltage graph ( G -+ S, a )n where G consists of a single loop d
imbedded in the sphere and a( d) = (1 2 3 · · · n ).
Example 4.2.2 Revisited. The branched covering of the Riemann surface for
w 2 = z(z - l)(z - 2) could be obtained from the imbedded voltage graph
(B 3 -+ S0 , a) given in Figure 4.11. The voltage group is f!E2 •
inside the loop and another outside the loop. The case of two branch points in
the sphere is considered in Exercise 13.
= nx(S) - L def(y) o
yeY
Example 4.2.4. The Riemann surface for the equation w 3 = z(z - 1)(z- 2)
has branch points at z = 0, 1, 2, oo, each of deficiency 2. By the Riemann-Hurwitz
equation, the Euler characteristic of the Riemann surface is 3x(S0 ) - 4 · 2 =
- 2, and we infer that the surface has genus 2.
2 + 2g loops in all
•• ••
The sharpened conclusion given here, that the branched covering can be
chosen to be two-sheeted, does not hold in higher dimensions. It is known,
however, that every closed, orientable 3-manifold is a three-sheeted branched
covering of the 3-sphere (Hilden, 1976; Montesinos, 1976). The following
nonorientable version of Alexander's theorem is also not known in higher
dimensions.
(12 4)
(1 3)(5 6) (1 3)(5 6)
(1 5)
(2 3 4 5) (2 3 4 5)
(12 4)
Figure 4.13. A permutation voltage graph imbedded in the torus.
type-0 loops are assigned the voltage 1 in fr3 and the single type-1 loop is
assigned a voltage that gives the exterior face a nonzero net voltage. Then
there are (k + 1)/2 branch points, each of deficiency 2. By the Riemann-
Hurwitz equation, the derived surface has Euler characteristic 3x(N1) -
2(k + 1)/2 = 2 - k. Since the number 2 - k is odd, the derived surface must
be nonorientable, from which we conclude that it is homeomorphic to Nk. D
4.2.6. Exercises
1. Let (G ~ S, a) 6 be the permutation voltage graph imbedding in the
torus given by Figure 4.13. Describe the branch and prebranch points of
a natural projection p: sa ~ S. Compute x(Sa) first by the Riemann-
Hurwitz equation, then by simply counting the number of vertices, edges,
and faces of the derived imbedding.
2. Find the branch points and prebranch points of the Riemann surface for
the equation w 4 = z 3 - z. What is its genus?
3. Show that the torus is a two-sheeted unbranched covering space over the
Klein bottle.
4. Generalize Exercise 3 by proving that the orientable surface S8 is a
two-sheeted unbranched covering of the nonorientable surface N8 + 1 for
all g ~ 0. (Hint: Represent the surface N8 + 1 by a bouquet of g + 1
type-1 loops.)
5. Use Theorem 3.4.5 and Exercise 4 for g = 0 to show that there are
graphs of arbitrarily large genus having two-sheeted planar coverings.
6. Give an imbedding of the Petersen graph in the projective plane such
that every face has size 5. Conclude that the !-skeleton of the dodeca-
hedron is a two-sheeted covering of the Petersen graph.
182 Imbedded Voltage Graphs and Current Graphs
7. Prove that the only surfaces that are unbranched coverings of themselves
(more than one-sheeted) are the torus and Klein bottle. Prove that the
only surfaces that are coverings of themselves with branch points are the
sphere and projective plane.
8. Prove that if S is an orientable surface, then S is a branched covering
over the sphere with exactly three branch points.
9. Prove that the nonorientable surface Nk, k > 3, is an unbranched
covering over the nonorientable surface N3 •
10. Prove that the orientable surface Sg, g > 2, is an unbranched covering
over the orientable surface S2 •
11. Show that for any imbedded voltage graph with voltage group fr2 the
sum of the net boundary voltages over all faces is 0. Conclude that any
two-sheeted branched covering must have an even number of branch
points.
12. Use Exercise 11 to prove that no surface of odd Euler characteristic is a
two-sheeted branched covering space over any surface.
13. Let p: S ~ S0 be a branched covering with two branch points in the
sphere S0 . Prove that if S is connected, it must be the sphere as well.
(a) (b)
Figure 4.14. Two placements of a surface of genus 2 in 3-space.
two of these is the third). This action on S2 clearly is pseudofree, but not free; in
particular, a 180° rotation about they axis has six fixed points. A free action of
~2 on the surface s2 is generated by the antipodal homeomorphism h(x, y, z) =
(- x, - y, - z ). The action of the group ~2 X ~2 X ~2 generated by the three
reflections through the coordinate planes is neither free nor pseudofree, since it
has entire circles of fixed points; however, this action contains both the two
previously described actions. Finally, the 3-space imbedding of the surface S2
depicted in Figure 4.14b reveals an action of the cyclic group ~3 generated by a
120° rotation about the z axis.
We observe that the various group actions in Example 4.3.1 could each be
obtained from an action on a graph at the "core" of the solid in 3-space
bounded by the surface. For instance, the core of the double torus in Figure
4.14a is a bouquet of two circles that run inside the handles. This observation
can be generalized.
Theorem 4.3.1. Let .91 be a group that acts on the graph G. Then .91 also acts
on a surface S whose genus is the cycle rank {J(G) = #E- #V + 1. If the
action of .91 on the graph G is free, then so is the action of .91 on the surfaceS.
Proof Imbed the graph G in 3-space and thicken it to form a solid, and let
S be the boundary of that solid. Obviously, the action of the group .91 on the
graph G extends to an action on the solid, from which we can extract an action
on the surface S. The action on the solid and on its boundary S is free if the
action on the graph G is free. If the graph G is a tree, then the surface S is a
sphere bounding a misshapen 3-cell. If T is a spanning tree from G, then each
edge not in T adds a handle to the surface S. Therefore, the genus of S is
{J(G), the number of edges not in the tree T. D
at the vertex v, and that his an automorphism of the graph G. We say that the
automorphism h "respects the rotation at the vertex v" if the rotation at the
image vertex h ( v) is
We say that the graph automorphism h "preserves the type of edge e" if
the image edge h (e) has the same type as the edge e.
Regular Branched Coverings and Group Actions 185
0
3 1 Figure 4.16. An imbedded voltage graph with volt-
ages in .2'8 .
natural action respects the derived rotation system and that the natural action
is free. By the corollary to Theorem 4.3.2, the conclusion of Theorem 4.3.3
holds. D
p: S-+ Sjd is locally one-to-one, except at fixed points. Since the domain is
compact, the quotient map is a branched covering. A branched covering
obtained this way is called "regular", and the group d is called the group of
"covering transformations". Note that p o IJ>a = p for any a Ed.
Example 4.2.1 Revisited. Once again we consider the complex analytic function
p(z) = zn. If p(z 1 ) = p(z 2 ), then the points z 1 and z 2 have the same modulus,
and their arguments differ by an integer multiple of 2'1Tjn. Thus, they are in the
same orbit of the pseudofree action of the cyclic group !!En on the extended
complex plane generated by a rotation about the origin through an angle of
2'1Tjn. It follows that the function p can be viewed as the quotient map of this
action, and accordingly, p is a regular branched covering.
branched covering behaves around branch points the same way that a natural
projection of an imbedded voltage graph behaves on faces. Thus we have the
following information about branch points.
This corollary provides most of the information we need for later use, but
much more can be said. For instance, the subgroup of the covering transfor-
mation group .91 that leaves fixed a particular prebranch point of y is a cyclic
group, and it is generated by a conjugate of the element b (see Exercise 13).
Many consequences of Theorem 4.3.5 result from the freedom allowed by
this theorem in the choice of the base imbedding G ~ S. Two such conse-
quences are now derived. The first is of general importance, especially for the
study of the genus of groups in Chapter 6. The second is narrower, but deeper,
and it draws upon a variety of results developed so far.
Theorem 4.3.6. If the group .91 acts pseudofreely on the connected surfaceS,
then there is a Cayley graph for .91 that is imbeddable in S.
Proof Just let a bouquet of circles assume the role of the graph G of
Theorem 4.3.5. By Theorem 2.2.3, the derived graph Ga is a Cayley graph for
.91. 0
Theorem 4.3.7. No finite group .91 acting on an orientable surface Scan have
exactly one fixed point.
Proof Since any element of the group .91 leaving a point fixed generates a
cyclic subgroup leaving that point fixed, it suffices to prove the theorem for the
case in which the group .91 is cyclic. Under that additional condition, let
S = S/.91 be the quotient surface, and let p: S ~ S be the quotient map. If
the action of .91 has only one fixed point, then the regular branched covering
p: S ~ S has exactly one prebranch point, and its order must be #.9/.
Suppose first that the quotient surface S is not the sphere. Then let G ~ S
be a one-face imbedding of a bouquet of circles containing the single branch
point of the quotient map p in its only face. The boundary walk of that face is
Regular Branched Coverings and Group Actions 189
of the form
a 1 b1 a 1 - 1 b1 - 1 a 2 b2 a 2 - 1 b2 - 1 ••• if Sis orientable
c1 c1 c2 c2 c3 c3 • • • if S is nonorientable
By Theorem 4.3.5, there is a voltage assignment a such that the quotient map
p is equivalent to a natural projection sa ~ s. If the quotient surface s is
orientable, then the net voltage on the only boundary walk is the identity,
since the group .91 is abelian. This would contradict the existence of a branch
point of p. On the other hand, if the quotient surface S is nonorientable, then
the net voltage is of the form 2a for some a E .91. It follows that #.91 cannot
be even, since then the order of the net voltage 2a and the associated
prebranch point would be at most #.91/2, that is, not equal to #.91. However
if #.91 is odd, then the derived surface would be nonorientable, by Theorem
4.1.5. Thus, in either case, there is a contradiction.
Finally, if the quotient surface S is a sphere, then let G ~ S be an
imbedding of a single loop containing the branch point of p in one of its two
faces. However, the other face has the same boundary walk and net voltage, so
it must also contain a branch point. D
Information about the number of fixed points in a group action can also be
obtained directly from the Riemann-Hurwitz equation. In the preceding
computations, if n = 5, then the number t of branch points equals 3, and the
quotient surface is the sphere. Since each branch point has deficiency 4, there
are also only three prebranch points in all. Prebranch points of a regular
covering are just the fixed points of the group of covering transformations.
Therefore, any pseudofree action of ~5 on S2 has exactly three fixed points,
as is the case in Example 4.3.2. A similar inspection of the preceding
computations shows that any pseudofree action of ~3 on S2 having the sphere
as quotient space has four fixed points, as is the case in Example 4.3.1 (the
possibility of the torus as a quotient is considered in Exercise 7). Deeper
applications of the Riemann-Hurwitz equation are deferred until Chapter 6.
4.3.6. Exercises
1. Show that every finite group acts freely on some surface.
2. Draw a band decomposition for the surface S3 that reveals an action of
~7•
3. Prove that the cyclic group ~n acts freely on the orientable surface S if
and only if n divides x(S).
4. Generalize Exercise 3 to arbitrary finite abelian groups.
5. Use the Riemann-Hurwitz equation to prove that any pseudofree action
on the projective plane has exactly one fixed point.
6. Prove that no group of even order acts pseudofreely on Nk for k odd.
(Hint: Use Exercise 4.2.12 and the fact that any group of even order has
an element of order 2.)
7. Show that every pseudofree action of ~3 on the surface S2 has exactly
four fixed points.
8. Let p: S-+ S be a regular branched covering such that the surface S is
orientable, and suppose that the covering transformation group is abelian.
Prove that p cannot have only one branch point. Show by examples that
p can have exactly one branch point if .91 is nonabelian or if S is
nonorientable. Does this contradict Theorem 4.3.7?
9. Prove that ~12 does not act pseudofreely on the surface S2 • (This is like
showing that ~16 does not act pseudofreely on S2 , but there are more
cases to consider.)
10. Complete the classification of cyclic pseudofree actions on S2 by showing
that ~n acts pseudofreely on S2 if and only if n = 2, 3, 4, 5, 6, 8, 10.
11. Prove for prime numbers n > 2 that ~n acts pseudofreely on the surface
Sg if and only if n is of the form (2g- 2 + m)/(m- k), where k is
even, k .:5; 2, and m > 1.
12. Use Exercise 11 to find for which primes n there is a pseudofree action of
~n on S5 •
Current Graphs 191
13. Prove in regard to the corollary to Theorem 4.3.5 that the subgroup of
the covering transformation group .511 that leaves fixed a particular
prebranch point of x must be cyclic and generated by a conjugate of the
element b.
(1) 3 245
(2) 4 3 51
(3) 5412
(4) 15 2 3
(5) 2134
192 Imbedded Voltage Graphs and Current Graphs
A 5
2 3 4
1 B
2 3 4
4 5 1 5 1 2 1 2 3
D 3 4 5
1
3 4 5
2 c
Figure 4.17. Five labeled polygons.
2 1
Example 4.4.1. The current graph in Figure 4.18 leads to a single generating
row for a triangular imbedding of the complete graph K 19 in the surface S20 • To
get that row (in our present notation), write the vertex v0 at the left, which is to
show which vertex has the rotation specified by the row. Then proceed as follows.
Start at any directed edge, for instance, the one labeled with current
9 modulo 19. Write that current as the subscript on the first vertex in the row.
(Since the resulting graph will have no multiple edges, the entries in the rows of
the scheme can be vertices, rather than directed edges.)
At the end of that first directed edge is a filled vertex, indicating that the
second edge is to be obtained by moving in a clockwise direction. That is, the
directed edge with current 7 is next, so write v 1 next in the generating row.
Continue in this manner, going clockwise at every solid vertex (•) and counter-
clockwise at every hollow vertex ( o ), until every edge has been traversed in both
directions. When proceeding in the minus direction on an edge, one writes the
additive inverse modulo 19 of the current shown on the plus direction. The
resulting generating row is
v0 • v9 v 1 v8 v3 v13 v15 v14 v11 v18 v4 v 11 v10 v 16 v5 v 1 v 12 v2 v6
e n•.
Observe that the rotation at vertex (/, ab) is obtained from the rotation at
vertex (/, b) by multiplying all ?A coordinates on the left by a.
Example 4.4.1 Revisited. For j = 1, ... , 9, let e1 be the edge whose plus
direction carries the current j. Gustin's method of obtaining a generating row is
clearly no more than an application of the Face Tracing Algorithm. Thus the base
imbedding G ~ S has one face whose directed boundary walk is
Denote this face by v. For j = 1, ... , 9 and i = 1, ... , 19, the derived edge
(e1 , i)+ runs from vertex (v, i) to vertex (v, i + j). Thus, the derived graph is
the complete graph K 19 • The rotation at vertex ( v, 0) is
Since the derived graph is simplicial, this rotation can also be given in vertex
form. If we let V; = ( v, i), the result is
which is the generating row given by Gustin's construction. Moreover, the full
rotation system is also the same as Gustin's scheme, since the rotation at vertex v;
is obtained from that at vertex v0 by adding i modulo 19 to all ?A coordinates.
3
7
6
2 6
1 7
3
Figure 4.19. A current graph with currents in .2"16 .
An inspection of these two rows reveals that the vertex u; is adjacent to the vertex
v1 if and only if j - i is odd, and that u; is adjacent to u1 (similarly V; and v) if
=
and only if j - i 2 mod 4. It follows that { u;: i even} U { v/ j odd} is the
vertex set for one component of the derived graph. In this component, every pair
=
of vertices is adjacent except u; and u1 , where j - i 0 mod 4 and V; and v1 ,
=
where j - i 0 mod 4. Thus, this component is isomorphic to the complete
multipartite graph K 4 •4 • 4 •4 • The other 16 vertices of the derived graph lie in a
second copy of K 4 4 4 4 .
Theorem 4.4.1 (Gross and Alpert, 1974). Let e 1• 1 ••• en•" be the rotation at
vertex v of the current graph (G ~ S, {J), and let c; be the current carried by the
direction of the edge e; that has v as its initial vertex. Let c = c1 • . • cn, and let r
be the order of c in the current group 114. Then the derived imbedding has #114/r
faces corresponding to vertex v, each of size rn, and each of the form
(el•',b),(e2• 2 ,bcl),(e3• 3 ,bC1C2) ... (en•n,bclc2 ... cn-1)
( e 1•', be), ( e 2•2 , bcc 1 ), ( e 3•3 , bcc1c 2) ...
Corollary. Let (G-+ S, /3) be a current graph such that KCL holds at every
vertex and every vertex has valence three. Then the derived imbedding is
triangular. D
Since KCL holds at every vertex in both Example 4.4.1 and Example 4.4.2,
both derived imbeddings are triangular. Thus the first gives a minimal-genus
imbedding of K 19 , whereas the second gives a minimal-genus imbedding of
K4.4.4.4·
Example 4.4.3. The current graph given in Figure 4.20 with currents in .2'6 has
one face, denoted v. By face tracing, we see that the rotation at vertex v0 of the
derived imbedding is
The order of the excess current at the bottom vertex of the current graph is 2. If
we excise the resulting three digons of the derived imbedding and reclose the
surface by identifying opposite sides, we thereby identify three pairs of multiple
edges from D; to D;+J• fori= 1, 2, 3. The final product is an imbedding of the
complete graph K 6 • The 3-valent vertex of the current graph satisfies KCL and it
generates six triangles, and the two remaining !-valent vertices together yield one
hexagon and two triangles. Thus, the Euler characteristic of the derived surface is
6 - 15 + 9 = 0. The scheme whose generating row is the given rotation for v0
with the extra v 3 deleted is precisely the table Heffter gives for his minimal-genus
imbedding of K 6 •
Example 4.4.4. Consider the current graph given in Figure 4.21 with currents in
the cyclic group .2'100 • There are two faces, which we denote u and v. In the
derived graph, there is an edge between U; and vi+l for all i, and there are no
other edges. Therefore the derived graph has two components; the vertex set of
one component is
{ u;: i even} u { v/ j odd}
Each vertex of the current graph has excess current ± 2 and valence 2, and
thereby generates a pair of 100-gons; each of these 100-gons passes through every
vertex in its component. Add a vertex inside each of the six 100-gons of one
198 Imbedded Voltage Graphs and Current Graphs
1 1
component, and add edges joining that vertex to the 100 original vertices. Then
delete all the original edges. The resulting graph is the complete bipartite graph
K 6 , 100 , and every face of the resulting imbedding is a quadrilateral. Hence the
imbedding has minimal genus.
e
E
0 Figure 4.22. A type-1 edge.
Current Graphs 199
if e is type 0
if e is type 1
Note the implication that the vertex rotations and edge types of the imbedding
G ~ S must be specified before currents are assigned (even in the orientable
case, an orientation of the imbedding surface must be chosen to specify vertex
rotations). The derived graph Gp is then defined as before.
To define the derived imbedding Gp ~ Sp, a direction for each boundary
walk must also be specified, since the Face Tracing Algorithm does not
determine a unique direction in the nonorientable case. It is also necessary to
specify at which corners a vertex rotation is reversed while tracing a face; such
a corner is called "reversed". Now suppose that the directed boundary of face
f is e 1'' . . . en••. The rotation at vertex (/, b) in the derived imbedding is then
where b; = b if f.;= + and the corner e;_ 1e; is not reversed or if f.;= - and
the corner e,e;+I is reversed, and where b; = bf3(e+)- 1 otherwise. Edge types
of the derived imbedding are defined so that if both directions of edge e
appear in the directed boundary walks of the imbedding G ~ S, then edge
( e, b) is type 0 for all b E ?A, and otherwise edge ( e, b) is type 1 for all
bE PA.
Example 4.4.5 (Youngs, 1968a). Let (G ~ S, /3) be the current graph il-
lustrated in Figure 4.23 with currents in .2"15 • As before, a filled vertex indicates
a clockwise rotation and unfilled, counterclockwise. Youngs's convention of
drawing type-1 edges broken in the middle with arrows in both directions,
suggestive of the usual " X", is used. The imbedding has one face, which we shall
call v. If we let e; denote the edge carrying current i, then the directed boundary
of face vis
1 5
6 3
2 4
Again following Youngs, reversed corners are indicated by dots. The rotation in
the derived imbedding at vertex ( v, 0), in edge form, is
where type-1 edges are given the superscript 1, as usual. The vertex form of the
rotation with ( v;, 0) = V; is
Faces of the derived graph are determined exactly as in the orientable case.
In Example 4.4.5 each of the 3-valent vertices satisfies KCL and, accordingly,
each generates 15 triangles in the derived graph. Each of the two 1-valent
vertices has an excess current of order 15 and induces in the derived im-
bedding a 15-gon whose boundary passes through each vertex. If an extra
vertex is placed in the center of each 15-gon and joined by edges to the
original vertices, the result is a triangular imbedding of the graph K 17 - K 2 •
An extra twist in an edge common to the two original15-gons allows the extra
two vertices to be joined by an edge, thereby yielding a minimal nonorientable
imbedding of K 17 •
Example 4.4.6. The current graph given by Figure 4.24 with currents in .2'13
has a single face v. Since there is an edge in the derived graph between vertices V;
and vi+J• for every i and for j = 1, 2, 3, 4, 5, 6, the derived graph is isomorphic to
the complete graph K 13 • Since KCL holds at every vertex, the derived imbedding
is triangular.
in Section 4.5. For present purposes, it suffices to know that a current graph in
a nonorientable surface has a nonorientable derived graph if the current group
has odd order.
4.4.6. Exercises
I. Give a current graph whose derived imbedding is a quadrilateral im-
bedding of the complete graph K 5 •
2. Draw a current graph that yields a triangular imbedding of the complete
graph K 7 • Give the resulting rotation system (scheme) in vertex form.
3. Find the first two rows in vertex form, of the rotation system for the
derived imbedding of Example 4.4.6.
4. Describe the derived imbedding for the current graph given in Figure
4.25a; the current group is .2"21 •
5. Show how a triangular imbedding of the complete 4-partite graph K 5 5 55
can be obtained from the current graph in Figure 4.25b. The cuire~t
group is .2"10 X .2"2 (Jungerman, 1975).
6. Obtain a triangular imbedding of K 10 - K 3 by changing the current
group in Figure 4.20.
7. Change the current group in Exercise 2 to get a triangular imbedding of
the octahedral graph 0 4 •
8. Assign currents in .2"13 to the nonorientable imbedding given in Figure
4.26a so that the derived imbedding can be used to obtain a nonorienta-
ble, triangular imbedding of K 16 - K 3 •
9. Complete the current assignment in .2"25 for the nonorientable im-
bedding given in Figure 4.26b in order to yield a nonorientable triangular
imbedding for K 25 •
10. Find the derived graph for the current graph illustrated in Figure 4.27
with current group .2"15 • (Hint: Show that the current graph has three
faces u, v, and w such that one component of the derived graph has
=
vertex set {(u, i), (v, j), (w, k) I i Omod 3, j =
1 mod 3, k 2 mod 3.} =
(5, 1) (1, 0)
(4, 1)
(0, 1)
(5,0) (1, 1)
2
Figure 4.25. Two current graphs; the cur~ent groups are, on the left, .2"21 and, on the right,
.2"w X .2"2 ·
202 Imbedded Voltage Graphs and Current Graphs
(a) (b)
Figure 4.26. Two nonorientable current graphs.
7 5 4 2
6
7 3 4
1 7 4 2
11. =
Suppose that m is odd and n 2 mod 4, where m > 2 and n > 2. Assign
currents in ~n to an m-cycle so that the derived graph can be used to
construct a quadrilateral imbedding of K m, n. (One current will be 3, and
all others will be 1.)
12. Give a nonorientable current graph whose derived graph is orientable.
The voltage and current graph constructions have obvious similarities, with
faces of a current graph playing a role of vertices of an imbedded voltage
graph and vice versa. We now give details of the formal duality.
e+*
+
I
I ..e+ • u
I
I Figure 4.28. The dual plus direction.
primal edge e in the same direction as the orientation at the initial vertex of
e+ (see Fig. 4.28).
It is important to note that the dual plus direction of e* depends as much
on the choice of orientation at the initial vertex of e + as on the choice of plus
direction for the edge e. Indeed, changing the orientation at the initial vertex
of e+ always changes the dual plus direction, but changing the plus direction
of edge e does not change the dual plus direction if the edge is type 1. One
unfortunate consequence of this dependence on local orientations is that the
dual of the dual plus direction is not necessarily the primal plus direction. This
is partly because there is no natural way to orient dual vertices; the Face
Tracing Algorithm can give different directions to a boundary walk around a
primal face depending on the choice of initial vertex for the walk. In some
cases, however, there is no way at all of orienting dual vertices that avoids this
problem.
Example 4.5.1. Figure 4.29 shows part of a locally oriented imbedding with
dual edges indicated by broken lines. The edge e is type 1 and the edged is type
0. If a clockwise orientation is chosen for the dual vertex f*, then d+ ** = d-; if
a counterclockwise orientation for the face f is chosen instead, then e + ** = e-.
~
I e
I "'
Example 4.4.1 Revisited. Paying heed to our previous warning, we first draw
the given current graph on a polygon with edges to be identified. It is readily
verified that the rotation at each vertex in Figure 4.31a is the same as that at the
corresponding vertex of Figure 4.18 (the specified orientation of the surface S2 is
clockwise). Therefore, the imbedding given in Figure 4.31a is equivalent to the
original imbedding. Its dual imbedded voltage graph is given in Figure 4.3lb.
The corners of the octagon represent a single vertex of the dual graph, and edges
of the polygon are also edges of the dual graph. It is instructive to check that the
rotation at the dual vertex, when read counterclockwise, agrees with the first row
of the rotation system for the derived graph given in Example 4.4.1, and that all
dual plus directions are correctly given.
The dual relationship between the current and voltage graph constructions
can now be given explicitly. In its original form, before the development of
Voltage-Current Duality 205
- c
5
2 2
5
----c-
Figure 4.31. The current graph of Example 4.4.1 and its dual imbedded voltage graph.
voltage graphs, the following theorem stated that the derived imbedding for
any current graph ( H ~ S, /3) is a branched covering of the dual of the
imbedding H ~ S. In its present form, the proof is a routine verification that
the definitions are satisfied, and it is omitted.
Theorem 4.5.1 (Gross and Alpert, 1974). Let ( H ~ S, /3) be a locally oriented
current graph and let (G ~ S, a) be its dual imbedded voltage graph. Then the
derived imbeddings Hp ~ Sp and Ga.~ sa are identical.
Theorem 4.5.1 explains many aspects of current graphs that were not
considered in the previous section, such as orientability and connectivity tests
for current-derived imbeddings, or were considered only briefly and without
justification, such as face tracing (Theorem 4.4.1) and excess current. These
aspects are much easier to study with voltage graphs.
Example 4.5.2 (Alpert). The nonorientable .2'28 current graph given in Figure
4.32 has one face. The derived graph is the Cayley graph for the cyclic group .2'28
with generating set {1, 2, 3, ... , 13}. Thus, it is isomorphic to the octahedral
graph 0 14 • All faces of the derived imbedding are triangles, by KCL, except for
two 28-gons corresponding to the two !-valent vertices. If an extra vertex is
placed at the center of each 28-gon and is joined by edges to each of the original
28 vertices, then the result is a triangular imbedding of 0 15 • But what is the
• 5 7 3 9 1
•
Figure 4.32. A nonorientable current graph with currents in .2'28 .
206 Imbedded Voltage Graphs and Current Graphs
orientability type of the surface? The edges carrying currents -10, 11, -12, and
13 are traversed twice in the same direction in tracing the face of the current
graph imbedding. Therefore, the corresponding loops in the dual imbedded
voltage graph are type 1. Since the voltages -12 and -10 are in the only
index-2 subgroup of .2'28 , the derived imbedding is nonorientable. This current
graph can be generalized to yield nonorientable triangular imbeddings for all
=
octahedral graphs On, n 0 mod 3, n > 3 (see Exercise 10).
Example 4.5.3. At the bottom left of Figure 4.33 is a current graph in the
sphere with current group .2'4 • At the bottom right is the dual imbedded voltage
graph. At the upper right is the derived imbedding, the 3-cube. At the upper left
is the dual derived imbedding, the 3-octahedron. Both base graphs can be viewed
as quotient graphs under the .2'4 action generated by a 90° rotation about the
vertical axis. For the cube (derived graph) the action is free; for the octahedron
(dual derived graph) the action has fixed points at the top and bottom vertex.
Let (H--+ S, {J) be a current graph, and let (G--+ S, a) be its dual
imbedded voltage graph. Since the surface projection p: sa --+ S may wrap a
face of the derived imbedding many times around a face of the imbedded
voltage graph, as in Example 4.5.3, the natural projection p: Hp* --+ H of the
dual derived graph onto the current graph is not, in general, a covering map.
Indeed, if r is the order of the excess current at vertex v of the current graph
H, then p: Hp* --+His locally r-to-1 near each vertex of p- 1(v).
Voltage-Current Duality 207
duality
1 1
duality
The graph map p: Hp* ~ H is not a branched covering, but rather a folded
covering, because the set of points where the map fails to be a local homeo-
morphism has dimension not 2, but 1, less than the complex H. Parsons et al.
(1980) and Jackson et al. (1981) [also Gross et al. (1982)] use the term
"wrapped quasi-covering" in their papers that analyze the construction, and
they develop it for direct use in graph-imbedding problems. We shall call it
simply a "wrapped covering".
Even if KCL holds at very vertex of a current graph so that the dual
derived graph does not cover the current graph, it is still not easy to determine
the structure of the dual derived graph. The following example illustrates the
problem in one case and presents an ad hoc solution.
Example 4.5.4 (White, 1974). Suppose that .J/1 is an abelian group and
X= { x 1 , ••• , x 2 m} is a set of distinct generators for .J/1 such that X; * x1- 1 for
all i and j. Let H ~ S be the imbedding of a bouquet of ~m circles in the surface
of genus m given by the rotation
Let fJ be the current assignment such that {J(e/) =X;, for i = 1, ... , 2m.
Figure 4.34 illustrates the case m = 2, where the edges of the octagon form the
dual voltage graph.
208 Imbedded Voltage Graphs and Current Graphs
Clearly the derived graph is the Cayley graph C( .91, X). We claim this
imbedding is self-dual; that is, the dual derived graph Hp * is also the Cayley
graph C(.Jil, X). Since KCL holds at the one vertex of the current graph H,
the graph Hp * is a regular covering of H, and accordingly, it must be a Cayley
graph for the group .91. However, it is not at all obvious that the generating set
for this Cayley graph is X. To show that it is, we give an explicit labeling of
vertices and coloring of edges that makes Hp * the desired Cayley graph.
By Theorem 4.4.1, for each a E .91, exactly one of the #.91 faces of the
derived imbedding, which we label a, has a directed boundary walk of the
form
The directed edge ( e 1 , a)- then forms part of the boundary walk of the face
labeled ax 2 -I, and the edge (e 2 , a)+ lies on the directed boundary of the face
labeled ax 1 - 1. If the vertex dual to face a is also labeled a, then the dual edge
( e 1 , a)* has its plus direction from dual vertex a to dual vertex ax 2 , and the
plus direction of ( e 2 , a)* goes from dual vertex ax 1 - to dual vertex a. The
other pairs of edges e 2 ;_ 1 and e 2 ;, for i > 1, behave similarly. Thus, if the
plus direction of the dual edge (e2i_ 1 , a)* is colored x 2 , and the minus
direction of the dual edge (e 2 ;, a)* is colored x 2;_ 1 , then the result is a Cayley
color graph for the group .J/1 and generating set X.
If the abelian group .J/1 in Example 4.5.4 is the cyclic group .2'4 m + 1 and
X= {1, 2, ... , 2m}, then the result is a self-dual imbedding of the complete
graph K 4 m+ 1 • The graph Kn has a self-dual imbedding only if n 0,1 mod4 =
Voltage-Current Duality 209
(Exercise 3); Stahl (1979) and Pengelley (1975) have obtained such self-dual
imbeddings of K 4 m+l (see Exercise 11). If the group .511 is ~4 m+ 2 and
X= {1, 2, ... , 2m} again, then the result is a self-dual imbedding of the
octahedron graph 0 2 m+I· Gross (1978) uses a variation of this construction to
compute the least number of edge crossings of certain octahedron graphs
imbedded in certain surfaces (Exercises 12-14).
Although it is difficult to determine the structure of the dual derived graph,
it is possible to give a readily applied criterion to decide when a dual derived
graph of a simplicial current graph is simplicial. The proof of the following
theorem is a straightforward application of Theorem 4.4.1 and is omitted.
Example 4.5.5. Let ( H ~ S, fJ) be the current graph given in Figure 4.35 with
the current group ~n X ~n. The excess currents at vertices v 1, v2 , v3 , v4 are
(1, 0), (1, 1), (- 2, 1), and (0, - 2), respectively. As long as the number n is odd,
no multiple (consider the group operation to be addition) of (1, 1) or ( -2, 1) has
=
a zero coordinate. If s(1, 1) = t(- 2, 1), then - 2t t mod n, so 3t 0 mod n. =
=
Thus, if n 1 or 5 mod 6, then the adjacent vertices of H satisfy the condition of
Theorem 4.5.2. Therefore the dual derived graph is simplicial. Since each vertex
of the graph H has excess current of order n, there are n vertices of the dual
derived graph corresponding to each V;, i = 1, ... , 4. Since no vertex correspond-
ing to V; is adjacent to another corresponding to V; and each dual derived vertex
has valence 3n, the dual derived graph is the complete multipartite graph
Kn. n, n. n. The imbedding is triangular and, therefore, minimal, because the dual
voltage graph is regular of valence 3.
0 1
2
Figure 4.36. A voltage graph and two current graphs.
Voltage-Current Duality 211
(0, 1)
(0, 0) (1, 1)
graph. On the right of Figure 4.36 is another current graph with currents in
f!En 12 , illustrated for the case m = 7. By Theorem 4.5.2, the dual derived graph
has no multiple edges or loops. Since the top and bottom vertices satisfy KCL,
they generate n/2 + n/2 vertices, each of valence m. Thus, they form one part
of the complete bipartite graph K m, n. The excess current at each of the
2-valent vertices has order n/2 in f!En 12 . Therefore, these vertices generate m
vertices of valence n. Since every face of the current graph imbedding is a
quadrilateral, so is every face of the derived imbedding.
A quadrilateral imbedding of Km n for m and n both even can also be
given as a dual derived graph. In this case, the construction is especially
elegant. The current group is f!Em 12 X f!En 12 , and the current graph is given in
Figure 4.37. It is left to the reader to verify that the dual derived graph is
Km,n·
Since K m, n has no cycles of length less than 4, it follows that
=
Proof Consider first the irregular case m 3 mod 4, n 1 mod 4. Then =
K m _ 1 n has a quadrilateral imbedding, which for n = 9 looks like Figure 4.38
(left) ~ear a vertex u in the ( m - 1) part of K m _ 1 n. Split the vertex u into
two vertices v and w and join each of these vertices to the nearest "hair' of
the surrounding n vertices as shown in the center of Figure 4.38. Extra edges
from v and w are needed to obtain Km n; this in turn will necessitate some
extra handles. The imbedding surface for Km-l,n has genus (m- 3)(n- 2)/4,
whereas the desired imbedding for Km n has genus f(m - 2)(n - 2)/41, which
= =
for m 3 mod4 and n 1 mod4 is' the same as ((m- 2)(n- 2) + 1)/4.
Thus, there is a leeway of (n- 1)/4 handles that can be used to put in the
extra edges at vertices v and w. The right side of Figure 4.38 shows how to
add the handles: each labeled face at vertex v is joined by a tube to the
212 Imbedded Voltage Graphs and Current Graphs
like-labeled face at vertex w. Four edges are then drawn on tube 1, two from v
and two from w, as illustrated in Figure 4.39. Three edges are drawn on tube
=
2, two from v and one from w. For general n 1 mod4, each tube carries
four edges except the bottom tube, which carries three.
=
Similar constructions work for m 3 mod 4 and n 0 or 3 mod 4; n/4 =
=
handles are added for n 3 mod 4. Again, the same construction works for
m = = =
1 mod 4, n 0 mod 4. Finally, for m 1 mod 4 and n 1 mod 4, a =
quadrilateral imbedding for K m _ 3• n is obtained first. Three vertices are split,
and (n - 1)/4 handles are added near each split vertex, as before. Since
f(m- 2)(n- 2)/41 = (m- S)(n- 2)/4 + 3(n- 1)/4 form= n 1 mod4, =
this yields the desired imbedding of K m n. All other combinations of con-
gruence classes mod 4 are covered by inte~changing m and n. D
4.5.5. Exercises
1. Draw the current graph whose dual voltage graph is given on the left in
Figure 4.6.
2. Draw the current graph whose dual voltage graph is given in Figure 4.1.
3. Prove that if Kn has a self-dual imbedding, then n 0 or 1 mod4. =
Voltage-Current Duality 213
chr(S) 5: l
7 + /49 - 24c
2
J
which is now known as the "Heawood inequality". The value of the expression
on the right-hand side is called the "Heawood number" of the surfaceS and is
denoted H(S).
The determination of the chromatic numbers of the surfaces other than the
sphere is called the "Heawood problem". Its solution, mainly by Ringel and
Youngs (1968), gave topological graph theory the critical momentum to
develop into an independent research area. The solution is that, except for the
Klein bottle, which has chromatic number 6, the chromatic number of every
surface equals the Heawood number. For example, the projective plane has
chromatic number 6 and the torus has chromatic number 7, exactly their
Heawood numbers. The idea of the proof is to imbed in each surface a
complete graph whose number of vertices equals the Heawood number of the
surface.
Since the sphere has Euler characteristic c = 2, its Heawood number is 4.
However, Heawood's argument for other surfaces cannot be used to establish
four as an upper bound for the chromatic number of the sphere. Although the
sphere is the least complicated closed surface, and although some of the most
distinguished mathematicians attempted to solve the problem, the chromatic
number of the sphere was the last to be known. This last case was resolved
when Appel and Haken (1976) established that chr(S0 ) = H(S0 ) = 4, which is
called the Four-Color Theorem. Since the proof of the Four-Color Theorem is
fully explained elsewhere, quite lengthy, and not topological in character, we
confine our attention to the Heawood problem.
215
216 Map Colorings
The original solution to the Heawood problem occupies about 300 journal
pages, spread over numerous separate articles, and it requires several different
kinds of current graphs, whose properties are individually derived. Ringel
(1974) has condensed this proof somewhat. Although Ringel considers every
case, it remains useful to refer to the original papers for complete details of
some of the more difficult cases (such as "orientable case 6").
The introduction of voltage graphs and topological current graphs unifies
and simplifies the geometric part of the solution. However, the construction of
appropriate assignments of voltages or currents remains at about the same
level of difficulty as originally. The present review of the Heawood problem
concentrates on a representative sample of the cases whose solutions are most
readily generalizable to other imbedding problems.
The first step in calculating the chromatic numbers of all the closed surfaces
except the sphere is to derive the Heawood inequality. The second step is to
use Heawood's inequality to reduce the Heawood problem to finding the genus
and the crosscap number of every complete graph. A computation of the genus
=
of each complete graph Kn such that n 7 modulo 12 illustrates the basic
approach to completing the second step.
for #F, which we substitute into the previous inequality. This yields the
inequality
By Theorem 1.1.1, the sum of the valences is equal to 2#E. Thus, the average
valence is 2#Ej#V. Substituting the upper bound 3#V- 3c for #E, we
conclude that
6c
average valence( G) ~ 6- #V 0
2 3 4 5 6 0 2
Figure 5.2. An imbedding K 7 -+ S1 is obtained by pasting the left side to the right side and then
pasting the top to the bottom with a 2/7 twist.
Example 5.1.3. Like the torus, the Klein bottle N2 has Euler characteristic
c = 0. By the same argument as for the torus it is proved that chr(N2 ) 5: 7.
Unlike the torus, however, the Klein bottle has chromatic number 6, one less
than its Heawood number, because-as we established in Theorem 3.4.6-the
crosscap number of K 1 is 3. This anomaly was discovered by Franklin (1934).
Theorem 5.1.3 (Heawood, 1890). Let S be a closed surface with Euler char-
acteristic c 5: 1. Then
chr(S) 2 - 7 chr(S) + 6c
is nonpos1ttve. To this end, let G be a graph imbedded in S such that
chr(G) = chr(S), and such that G is chromatically critical. From Theorem
5.1.1, it follows that
6c
average valence( G) 5: 6 - # V
The Heawood Upper Bound 219
6c
chr( S) - 1 < 6 - -
- #V
( chr(S) -
1 - v49 - 24c ) ( 1 + v49 - 24c )
2 chr(S) - 2 :;;; 0
For c :;;; 0, the value of the expression 7 - v49 - 24c is less than or equal to
0. Thus, the value of the first factor is positive. It follows that the value of the
second factor is nonpositive. Since chr(S) is an integer, the conclusion follows.
D
chromatically critical, we can color the graph G - u with five colors, say the
colors e1, ••. , es. If any two of the five neighbors of u had the same color, then
some color e; would not be applied to any neighbor of u. Accordingly we
might extend the coloring of G - u to a coloring of G by assigning colore; to
u. Thus, we suppose that the five neighbors u1, ... , Us of u are assigned the
colors e 1, ••• , es, respectively, corresponding to the order in which they are
encountered in a clockwise traversal around u. This is illustrated in Figure 5.3
For any two colors e; and e1 , consider the subgraph of G- u containing
every vertex that is assigned color e; or e1 . Every component of this subgraph
is called a "ere1 Kempe chain".
First, suppose that the vertex u3 is not contained in the ece 3 Kempe chain
that contains the vertex u1 • One might recolor the vertices of this ece 3 Kempe
chain by reversing the roles of e1 and e3 , that is, by recoloring every eccolored
vertex e3 and recoloring every e3-colored vertex e1 . This results in a proper
coloring of the graph G - u such that u1 and u3 are both assigned color e1 ,
while vertices u2 , u4 , and us are still assigned colors e2 , e4 , and es respec-
tively. We can extend this revised coloring of G - u to a coloring of G by
assigning color e3 to u.
Alternatively, if the vertex u3 is in the same ece3 Kempe chain as u1 , then
u2 and u4 cannot be in the same e2-e 4 Kempe chain, because of the Jordan
curve theorem. Thus, we might revise the coloring of G- u by reversing the
assignment of e 2 and e4 to the vertices in the e2-e 4 Kempe chain that contains
u4 • Then color e4 would be assigned to no neighbor of u, and so it could be
assigned to u, yielding a 5-coloring of G. D
Tables 5.1 and 5.2 show the calculated values of the Heawood number for
some orientable surfaces and for some nonorientable surfaces, respectively.
From the formula for the Heawood number, one sees that it must increase
as the genus or crosscap number increases, but at a rate proportional to the
square root of the genus or crosscap number. Tables 5.1 and 5.2 confirm this.
Thus, every sufficiently large integer is the chromatic number of at least one
orientable surface and of at least one nonorientable surface. Moreover, the
same chromatic number may serve for several different surfaces of consecutive
genera or of consecutive crosscap numbers, and the number of consecutive
surfaces that share the same chromatic number tends to increase along with
the chromatic number.
One reads from Table 5.1 that H(S20 ) = H(S21 ) = H(S22 ) = 19. Now
suppose that there were a graph G imbeddable in S20 such that chr(G) = 19.
It would follow that chr(S20 ) ~ 19. Moreover, since any such graph that is
imbeddable in s20 is also imbeddable in s21 and s22' it would also follow that
chr(S21 ) ~ 19 and chr(S22 ) ~ 19. By Theorem 5.1.3, the chromatic number of
each of these three surfaces is also at most 19, their common Heawood
number. Thus, they would all have chromatic number exactly 19. In fact, this
is true, because the complete graph K 19 is imbeddable in S20 , which is proved
in Example 4.4.1. In general, complete graphs are used in this manner to show
that the chromatic number of a surface is equal to its Heawood number. To
formalize the process, we define the numbers
I(n) = r
(n- 3)(n- 4)
12
l and
-
l(n) =
r(n-3)(n-4)l
6
Theorem 5.1.5. Assume that n ~ 7 and that y(Kn) ~ I(n). Then for any
orientable surface Sg such that I(n) ~ g < I(n + 1), the equation chr(Sg) =
H(Sg) = n holds.
222 Map Colorings
Proof Substitution of /(n) and l(n + 1) into the formula for the Heawood
number establishes by direct calculation that H(S1<n>) = n, that H(S1(n+l)) =
n + 1, and that for a surface Sg such that /(n) 5: g < l(n + 1), H(Sg) = n.
By Theorem 5.1.3, chr(Sg) 5: H(Sg) = n. Under the stated assumption, on the
other hand, the complete graph Kn is imbeddable in every such surface Sg, so
that chr(Sg) :;:: n = H(Sg). 0
Theorem 5.1.6. Assume that n > 8 and that y(Kn) 5: J(n). Then for
any nonorientable surface Nk such that I(n) 5: k < I(n + 1), the equation
chr(Nk) = H(Nk) = n holds.
I(n) -
- r (n- 3)(n- 4)
12
1-
-
(n- 3)(n- 4)
12
for n =0, 3, 4, or 7 modulo 12
A review of the proof of Theorem 5.1.7 reveals that this equality is obtained if
and only if #F = t#E, which happens, of course, if and only if the graph Kn
triangulates the surface of genus I ( n ).
In general, the quantity /(n)- (n- 3)(n- 4)/12 measures by how much
an orientable imbedding Kn ~ S1<n> would fail to be a triangulation, in the
following sense. For each positive integer i, let F; denote the set of i-sided
faces of the imbedding. Then
5.1.6. Exercises
1. Let G be a 16-vertex graph with four vertices each of valences 4, 5, 6, and
7. Prove that the graph G is not planar.
2. Construct a 5-regular planar graph with as few vertices as possible.
3. Suppose that G is a connected chromatically critical graph such that
chr( G) = 3. Is G necessarily an odd cycle?
4. Construct a chromatically critical graph G such that y(G) = 1 and
chr(G) = 5.
5. Prove that the removal of an edge from K 7 yields a graph of crosscap
number 2.
6. Construct an imbedding of K 6 in a Mobius band.
7. Let G be the edge-complement in K 8 of a 1-factor. What is the genus
of G?
8. What is the chromatic number of the graph G of Exercise 7?
9. Suppose that y(G) = 1 and that chr(G) = 7. Prove that G contains a
sub graph isomorphic to K 7 •
#V = 1 #E = 6s + 3 and #F = 4s + 2
Figure 5.4. Imbeddings 8 9 --+ ~ and 8 15 --+ S3 . These are base graphs for imbeddings K19 --+ S20
and K31 --+ S64 .
Alternatively, one may observe that the global Kirchhoff voltage law corre-
sponds to a system of 4s + 2 linear equations in 6s + 3 unknown edge-voltages,
which would seem to have many solutions. However, the requirement that
each of the voltages 1, ... , 6s + 3 be used exactly once corresponds to an
inequality ·
n(x;- xi)* 0
I, j
3s+2
Figure 5.5. This assignment of voltages in .2"12,+ 7 to the imbedded bouquet B6,+ 3 -+ Ss+l
satisfies KVL and yields as a derived graph a triangular imbedding K12 ,+ 7 -+ S1<12 s+?)·
other edge. Since the voltage 2 is assigned to the outer edge of the second
triangle from the top, in the same direction as the voltage 3s + 1, that triangle
will satisfy KVL if its other edge is assigned the voltage 3s + 3. In this manner
the voltages 1, ... , 2s are used to coil from the middle voltage 3s + 2 in the
middle range outward to the limiting voltages 2s + 2 and 4s + 2. Figure 5.6
shows a combinatorial abstraction of the coil.
2s
4s +2
2s-1
Figure 5.6. The voltages 1, ... ,2s are used to coil outward from 3s + 2 toward 2s + 1 and
4s + 2.
228 Map Colorings
5 voltages
in~l9
3LSJ3 2
voltages
in~7
2 2
(a)
5
(b)
voltages
in;2"31
3 3
8
(c)
Figure 5.7. Imbedded voltage graphs that yield triangular imbeddings of (a) K7 , (b) K19 , and
(c) K3,.
• clockwise
11 14 12 o counterclockwise
4 3 2
6 9 7
5
Figure 5.8. A Gustin current graph for K31 .
• clockwise
11 14 o counterclockwise
6 9
5
Figure 5.9. Pictorial representation of the Face Tracing Algorithm.
230 Map Colorings
• clockwise
o counterclockwise
4s+3 6s+2 5s+5 5s+1 5s+4 5s+2
2s 2s-1 4 3 2
. . . ---¢----c>-----o--r--4
2s+2 4s+ 1 3s+4 3s 3s+3 3s+1
2s + 1
Figure 5.10. A general current graph for orientable case 7 of the Heawood problem.
It would not make sense to depend on the largest odd factor of 12s + 4,
since 12s + 4 might be a power of 2, for example if s = 5 or 85. However,
there is a tactic that does help, namely, to double some 1-factors of K 12 s+4·
"Doubling" a set of edges means to insert an additional adjacency along each
edge in the set, that is, so that it has the same two endpoints.
If one 1-factor of K 12 .+ 4 were doubled, then the resulting graph would
have 12s + 4 vertices and (6s + 2)(12s + 3) + (6s + 2) edges, that is,
(6s + 2)(12s + 4) edges. This looks promising, since assigning the voltages
1, ... , 6s + 2 modulo (12s + 4) to the edges of the bouquet B6.+ 2 yields as a
derived graph K 12.+ 4 with a doubled 1-factor. The edge with voltage 6s + 2
leads to the double edges in the derived graph, because it is a loop with a
voltage of order 2.
One does not want a triangular imbedding of K 12 s+4 with a doubled
1-factor. Instead, the idea is to derive an imbedding so that each pair of
doubled edges bounds a digon, but that all other faces are three-sided. Then
each digon may be excised and its two sides sewn together, that is, identified.
The result would be a triangular imbedding K 12.+4 ~ S1<12.+ 4 >. The number
of digons in the derived imbedding would be 6s + 2, and the number of
triangular faces would be (6s + 2)(8s + 2).
According to Theorem 4.1.1, such a derived imbedding could be obtained
from an imbedded voltage graph with one vertex, 6s + 2 edges, and 4s + 2
faces. Of these faces, 4s + 1 should be 3-sided and satisfy KVL. The other
face should be a monogon whose net voltage has order two. An easy calcula-
tion reveals that the base surface would have Euler characteristic 1 - 2s, an
odd number, so the base surface would be nonorientable. Figure 5.11 shows
such an imbedded voltage graph for K 4 , when s = 0.
Some experimentation reveals that it is not easy to generalize Figure 5.11.
One of the problems that would arise is the necessity of proving that the
derived surface is orientable. Of course, this is obvious for Figure 5.11,
because the derived surface has Euler characteristic 2. However, every non-
positive even Euler characteristic is shared by an orientable surface and a
nonorientable surface.
These problems are completely circumvented if one doubles three 1-factors
of K 12 s+4• thereby obtaining a total of (6s + 3)(12s + 4) edges. This time the
derived imbedding has 18s + 6 digons, to be excised, and (6s + 2)(8s + 2)
triangular faces. A plausible base imbedding for this configuration might have
one vertex, 6s + 3 edges, and 4s + 4 faces. Three of these faces could be
(3, 0)
(3, 1) (3, 1)
(3, 0)
Figure 5.12. Voltages in .2"8 X .2"2 for an imbedding 8 9 ..... S1 . The derived graph is K16 with
three doubled 1-factors, and the imbedding surface is S13 .
(6, 1)
(10, 1)
(6,1)
Figure 5.13. An imbedding bouquet 8 21 -+ S3 with voltages in .2'20 X .2'2.
voltages (3, 0), (6, 0) and (9, 0) coil among the voltages whose first coordinate is
congruent to 1 modulo 3, while the voltages (3, 1) and (6, 1) coil among the
voltages whose first coordinate is congruent of 2 modulo 3. The voltage (9, 1) is
used to patch the coils together. The current graph in Figure 5.14 shows the
coiling and patching clearly. The general solution to orientable case 4 was first
achieved by Terry et al. (1970), and is left as an exercise. (See Exercise 11.)
(0, 1)
(7 (3, 1) (3, 0) (9, 0)
(~
(6, 1) (9, 1) (6, 0)
6s
Figure 5.15. This base graph has 2s loops at each of its three vertices and 4s + 1 edges between
each pair of vertices. The derived graph consists of three isomorphic copies of K 12 , + 3 •
1 13
13 3 16
13
l<'igure 5.16. A three-orbit .2"27 -current graph that yields a triangular imbedding K 27 -+ S46 .
is used to show that the current graph has three edge orbits, each of which
contains each nonzero current exactly once.
Orientable case 0 was first solved by Terry et al. (1967). The complication
here is that if the number 12s has no large odd factors, then by Theorem 5.2.1,
the complete graph K12 , has no small quotients. The solution here involves
the doubling of many 1-factors and the use of nonabelian voltages or currents.
In particular, if 2k is the largest power of two that divides 12s, then 2k - 1
1-factors are doubled.
5.2.6. Exercises
1. Number the vertices of an eight-sided polygon in consecutive cyclic order
from 0 to 7. For each pair of vertices that is identified in pasting opposite
edges together, write a transposition. Note that fori = 1, ... , 7, there is a
transposition (i, i + 3 modulo 8). How many vertex orbits are there in the
permutation group generated by these transpositions?
2. Suppose that opposite sides of a 4n-sided polygon are pasted together,
and let the image of the polygon rim be regarded as a graph imbedded in
the resulting surface. Prove that the graph has only one vertex. (Hint: See
Exercise 1.)
3. Suppose that opposite sides of a 4n + 2 sided polygon are pasted
together as in Exercise 2. How many vertices does the graph have?
4. Try to complete the partial voltage assignment shown so that the derived
imbedding is K 19 ~ S20 • Paste opposite sides together.
voltages
in.2"19
236 Map Colorings
voltages
9 1
in~I9
voltages
3 ln.2'i4
Figure 5.17. A voltage graph in N2 that ultimately yields a triangular imbedding K 15 -+ N22 •
on the other side of the digon boundary remains the same as before the
matching.
The two crosscaps in Figure 5.17 convert the apparently spherical imbed-
ding surface in which the figure is drawn into the Klein bottle N 2 • The edge
identifications convert the graph into the bouquet B 1 . Four of the regions are
three-sided and satisfy KVL in the group ~14 • In particular, 4 + 5 + 5 = 14
and 2 + 6 + 6 = 14. There are also two monogons, one with boundary
voltage 7, the other with boundary voltage 1.
The derived graph is K 14 with a doubled 1-factor, and the derived imbed-
ding has 56 = 14 · 4 triangular faces, seven digons, and one 14-sided face. The
Euler characteristic of the derived surface is 14 - 98 + 64 = -20. Since the
even numbers 0, 2, 4, 6, 8, 10,12 form the only subgroup of index 2 in ~14 , and
since the base edge carrying voltage 6 is orientation-reversing, the derived
surface is nonorientable, by Theorem 4.1.4.
Excising the seven digQns and closing up the holes converts the derived
imbedding into an imbedding K 14 ~ N22 such that every face is triangular,
except for one 14-sided face whose boundary circuit is a Hamiltonian cycle.
Insert a 15th vertex into the interior of that face and adjoin it to each of the
other 14 vertices. The result is a triangular imbedding K 15 ~ N22 •
5 6
4 2
7 1
Figure 5.18. The Youngs nonorientable current graph to produce the imbedding K 15 --+ N22 •
4s 2 4 6 4s-4 4s-2
6s + 1 2s + 1 2s - 1 2s + 3 2s - 3
••• 3 4s-1 1
Figure 5.19. The Youngs nonorientable current graph that yields K12 ,+ 3 ..... Ni(l 2 s+J). The
currents are in -2'"12,+2.
4s+2 2 4 6 4s-2 4s
6 6
3
Figure 5.21. An imbedding 8 11 ..... N4 with voltages in .2'"23 .
240 Map Colorings
2 4 6s-2 7 3s-2 3s+4
::::~·I ~~,,.. :. ..
6s-3 5 6s-1 8 3s-1
6
3s+5
3
Figure 5.22. The Youngs current graph for nonorientable case 0. The current group is .2'12 s _ 1 •
The rectangular sheet in Figure 5.21 is made into a torus in the usual
manner, by matching opposite sides together. Then two crosscaps are inserted,
to obtain N4 • The Kirchhoff voltage law holds on every face except the
monogon.
In general, to derive the imbedding K 12 • ~ Ni(l2s)• the base graph is the
bouquet B6 • _ 1 , and the base surface is N2 ., formed by inserting two crosscaps
into the orientable surface s._ 1 • The base imbedding has one monogon and
4s - 1 triangular faces. The voltages 3, 6, ... , 6s - 6 coil among the voltages
4, 7, 10, ... , 6s - 2 and also among the voltages 5, 8, 11, ... , 6s - 1. The
Youngs current graph is shown in Figure 5.22.
5.3. 7. Exercises
1. Using Figure 5.19, draw a voltage graph from which a triangular imbed-
ding K 27 --+ N92 can be obtained.
Additional Adjacencies for Irregular Cases 241
Figure 5.23. An imbedding 8 14 -+ N5 with voltages in .2"28 . The derived graph yields an
imbedding K 28 -+ N1 00 .
2. For the current graph of Figure 5.20, calculate exactly how many faces of
each size occur in the derived graph.
Kl2s+5- e~ Sl(l2s+5)-l
Figure 5.24. A three-orbit .2'27 -current graph that ultimately yields a triangular imbedding
Kz9- Kz ..... Ss4·
K12s+3 ~ SI(l2s+s)-1
would be for a graph with 12s + 3 vertices and (12s + 3)(6s + 1) edges. There
would be two (12s + 3)-gons and (12s + 3) · 4s triangles. One might hope
that it covers an imbedded voltage graph with one vertex, 6s + 1 edges, 4s
triangles, and two monogons. Unfortunately, no such 1-vertex voltage graph
or corresponding 1-orbit current graph for case 5 is known. The imbedding is
achieved, instead, with a 3-orbit current graph closely related to case 3, as
illustrated for K 20 in Figure 5.24, in which x and y denote the vertices with
excess current.
Once we have the imbedding K l2s + 3 ~ S 1<12 • + S) _ 1 , we install a vertex in
the interior of the two complete (12s + 3)-gons and adjoin it to each of
the 12s + 3 vertices on the region boundary. This yields an imbedding of
K 12 s+s - e into S 1<12 .+ 5>_ 1 • An additional handle is added for the missing
edge, thereby achieving our goal, an imbedding Kns+s ~ S 1<12 s+S>·
0 0 0
2 5 4 3 6
1
Figure 5.25. A .2"Tvoltage graph with base imbedding B3 -+ So and derived imbedding K 7 -+ S3 •
The three complete 7-gons of the derived imbedding are shown at the right.
2 3
2 3
2 3
:: I~ I,
)I Jl Jl
• • • 2s+ 1
X
.... • -4
3s+2 3s+3 3s+ 1 4s+2
Figure 5.29. A .2'12 s+ 7-current graph used in the general solution to orientable case 10. The
vertices with excess currents are labeled x, y, and z.
Additional Adjacencies for Irregular Cases 245
and then a handle is added to create room for the missing 3-cycle.
The rotation graph for orientable case 6 is a fascinating variation on the
one for case 3, and the assignment of currents presents interesting new
difficulties. Analogous to case 10, the near-final objective is a triangular
imbedding
Orientable cases 11, 2, and 8 require special ingenuity. In case 11, the
branched covering space is a mostly triangular imbedding of K 12 s+6• which
has five exceptional faces, each a complete (12s + 6)-gon. Figure 5.30 shows
the current graph used when s = 2 and 12s + 11 = 35. When new vertices are
inserted into the (12s + 6)-gons and adjoined to the vertices on the boundary
cycles, the result is a triangular imbedding
Kl2s+ll- Ks ~ SI(l2s+llJ-2
13 11 7 10
12
7 7
J sing two handles to create room to add the rmssmg ten edges of K 5 is
ubstantially more difficult than the analogous problem of using one handle in
ase 10 to add K 3 • Cases 2 and 8 are like case 11 in some important ways, but
1ey include complicated additional irregularities.
K12s+3 ~ Nl(l2s+5)-1
with two complete (12s + 3)-gons and all other faces three-sided. Figure 5.31
contains the voltage graph to be used when s = 1 and K 12 s+s = K 11 •
A new vertex is inserted into the interior of each of the two complete
(12s + 3)-gons and adjoined to each vertex of the (12s + 3)-gon boundary.
The result is a triangular imbedding
Figure 5.32 illustrates the vicinity of the vertex 0, where we have assigned x
as the label on the new vertex installed into the 15-gon covering the monogon
6 6
Figure 5.33. Since the vertices x and y have vertex 0 as a common neighbor, there is a way to
adJoin them if only one more crosscap is added.
with voltage 1, and y as the label on the other new vertex, for the special case
of K 11 •
We shall now demonstrate that only one additional crosscap is needed to
create room for the missing edge xy. Figure 5.33 illustrates the appropriate
surgical procedure. First, split the vertex 0 into the two vertices o+ and o-, so
that the vertex x and all the vertices on one side of the path yOx are adjoined
to 0-, whereas the vertex y and all the vertices on the other side of the path
yOx are adjoined too+.
Next, delete the interior of a disk whose boundary passes through both o+
and 0- and parametrize that boundary by the unit circle so that 0 + and 0-
are antipodal points. Also, draw edges from x and y, respectively, to another
pair of antipodal points. Finally, reclose the surface by identifying all pairs of
antipodal points on the parametrized boundary. The antipodal identification
adds a crosscap, it completes an edge between x and y, and it reunites vertices
0 + and 0- into a single vertex 0 adjacent to every other vertex. Thus, the
result is an imbedding K 17 ~ Ni(I?). This technique easily generalizes to all of
nonorientable case 5.
5.4.6. Exercises
1. Prove that graphs of the form K 12s+s- e or of the form K 12 s+l0- K 3
have no nontrivial quotients.
248 Map Colorings
2. Trace the edge orbits in Figure 5.24, in order to demonstrate that there are
three orbits and that each nonzero current occurs exactly once on each
orbit.
3. Use the pattern of Figure 5.24 to develop a current graph for the general
solution of orientable case 5.
4. Draw the imbedded voltage graph used for K 22 •
5. Use Figures 5.28 and 5.29 to construct a general solution to the ad-
ditional-adjacency problem for case 10.
6. Describe the face-size distribution for the voltage graph dual to Figure
5.30 and the order of the net voltage on each face.
6
The Genus of a Group
Abelian groups are an obvious first target in a general attack on the genus of a
group. Indeed, White (1972) initiated the modem study of the genus of a
group by computing the genus of certain abelian groups, using methods
discussed in Section 3.5. Jungerman and White (1980), whose work is now
described, subsequently succeeded in determining the genus of "most" other
abelian groups.
Before we proceed, a comment about generating sets is in order. A
generating set X for a group s# can be irredundant without being the
249
250 The Genus of a Group
O ---J)
....__.,.,-"'---/ /
,-
______
..,...
\ I
Figure 6.1. The Cayley graph C(.9'3 ,
"-
X) and two quotients. Edges corresponding to generator
(1 2 3) are solid, and generator (1 2), dashed.
minimum in size over all generating sets for .91. For example, consider the
group .91= ~6 and the irredundant generating set {2, 3}. If the generating
set X has "minimum" size, then the number #X is called the "rank" of the
group .91.
In computing the genus of a group, one need consider only irredundant
generating sets, because discarding extra generators serves only to delete edges
from the associated Cayley graph. Obviously, deleting edges cannot increase
the genus of a graph. On the other hand, as we shall see, it is possible that the
genus of a group is not attainable by using any of its minimum-size generating
sets. (See Exercise 6.4.16.) This was first observed by Sit (1980), in response to
a problem of Gross and Harary (1980).
Theorem 6.1.1. The natural voltage assignment a on the Schreier coset graph
C = S(s#: !!J, X) yields a derived graph with #[.5#: !!J] components, each
isomorphic to the Cayley graph C(s#, X).
Proof Suppose that !!J1 = !!J, !!J2 , ••• , !!Jn are the right cosets of !!J. Then
the n#s# vertices of the derived graph ca are the pairs (!!J;, a), where a E s#
and 1 5: i 5: n. There is an edge from the vertex ( !!J; , a) to the vertex ( !!J1 , b)
if and only if there is a generator x E X such that !!J1 = !!J;x and b = ax. We
define the vertex part of a graph map f from the Cayley graph C( s#, X) to the
derived graph ca by the rule f(a) =([a], a), where [a] denotes the right coset
of !!J that contains a. Also, let f take the edge of C(s#, X) from a to ax to
the edge in co. from ([a], a) to ([a]x, ax). Since [a]x = [ax], we infer that the
resulting map is a graph map. It clearly takes the set of directed edges
originating at the vertex a of C(s#, X) one-to-one onto the set of directed
edges originating at the vertex ([a], a) of ca. Therefore, f is a covering map
from C( s#, X) onto a component of ca. Since f is also one-to-one on the
whole vertex set of C(s#, X), it follows that f is a graph isomorphism onto
that component. The theorem follows, since the components of the derived
graph of an ordinary voltage graph are mutually isomorphic. 0
Proof Apply Theorem 6.1.1 with the kernel of q, as the subgroup !!J, so
that the Schreier coset graph S( .5#: !!J, X) is isomorphic to the Cayley graph
C(fl, q,(X)). o
conditions:
i. In each directed face boundary there is exactly one edge assigned the
current x and one the current x-I for every generator x E X, and all
currents or their inverses are in X.
ii. There is a subgroup ?A contained in .J/1 and a one-to-one correspondence
between the faces of the imbedding G --+ S and the cosets of ?A, such that
if the edge e lies in faces corresponding to cosets ?A 1 and ?A 2 , then
?A 1a( e •) = ?A2 , where e • is the direction of e given by the directed
boundry of the face for ?A 1 •
Then each component of the derived graph Ga. is isomorphic to the Cayley graph
C(.J/1, X).
Proof Conditions i and ii imply that the dual graph for the imbedding
G --+ S is the Schreier coset graph S(.J/1: ?A, X) and that the dual voltage
assignment is the natural one. The corollary then follows from Theorem 6.1.1.
D
The following two examples further develop the quotients of Figure 6.1 in
the context of Theorem 6.1.1 and its corollaries. The first example concerns the
left quotient, which is regular, and the second the right quotient, which is not.
Example 6.1.1. On the left of Figure 6.2 is a voltage graph imbedded in the
sphere with voltages in the symmetric group .51'3 generated by X= { x, y }, where
x = (1 2 3) and y = (1 2). The given voltage graph is the Schreier coset graph
S( .9'3 : ?A, X) where ?A is the subgroup generated by (12 3). Alternatively, since
?A is normal in .51'3 with quotient map <P: .9'3 --+ f!E2 such that <P(x) = 0 and
<P(y) = 1, the voltage graph may be viewed as the Cayley graph C(f!E2 , {0, 1}).
The vertices have been labeled by the cosets ?A and PAy (rather than 0 and 1) to
maintain the Schreier coset viewpoint. By Theorem 6.1.1, the derived imbedding
for this voltage graph consists of two copies of the Cayley graph C( .9'3 , X). If one
prefers to view the voltage graph as a Cayley graph, use the first corollary to
Theorem 6.1.1 instead. On the right of Figure 6.2 is the dual current graph. The
-- '
O _!_n
I / ...... \
I :dy \
x I I
'---'~
• .. i 'x • :y
X
y 1Y I
~ I
Figure 6.2. An imbedded voltage graph and its dual current graph.
The Genus of Abelian Groups 253
I
ly
l
y y
l -,'l
X
I 1,. X= (12 3)
1
X X
r
I
I
I
X \X
-- y
/
y = (12)
reader should verify that the edge and face labels satisfy conditions i and ii of the
second corollary.
Example 6.1.2. On the left of Figure 6.3 is a current graph imbedded in the
torus with currents in the symmetric group 9"3 generated by X= { x, y }, where
x = (1 2 3) andy = (1 2). Clearly condition i of the second corollary to Theorem
6.1.1 is satisfied. If the faces are labeled as illustrated, by cosets of the subgroup
!!4 generated by the permutation (1 2), then condition ii is also satisfied. Thus the
derived graph for this current graph consists of two copies of the Cayley graph
C( 9"3 , X). On the right of Figure 6.3 is the dual voltage-graph imbedding, which
the reader should recognize as the right quotient in Figure 6.1, the Schreier coset
graph S(Y'3 : !!4, X).
In terms of the dual current graph, this table gives the currents encountered in
a trip around each face boundary. In the early literature, these are sometimes
called the "logs of the circuits". In effect, the table of logs is just a condensed
rotation system for the derived imbedding of the Cayley graph C( s#, X). That
is, edges incident on a vertex are identified by their "color" x or x- 1, x E X,
and the rotation at vertex a, where a is in coset !!4,, is simply the log for
circuit !!4,.
254 The Genus of a Group
such that m; divides mi+l fori= 1, ... , r- 1, and m 1 > 1. Each element of
.J/1 is written as an r-tuple whose coordinates correspond to the canonical
product form. The "canonical generating set" for .J/1 consists of the r elements
that have 1 as one coordinate and 0 for all others.
Cyclic groups !!En and abelian groups of the canonical form f!E2 X !IEm
have planar Cayley graphs, and groups with canonical form f!Em 1 X f!Em 2 ,
m 1 > 2, have genus 1 (see Exercise 6). Also, if the canonical form of an
abelian group has any f!E2-factors, then its genus is easily computed using
methods of Section 3.5. Therefore, in the rest of this section, we consider only
abelian groups of rank r > 2 having no f!E2-factors in their canonical form.
The following theorem establishes a lower bound on the genus of such an
abelian group that has no f!E3-factors in its canonical form either. The bound
given is simply the genus of an imbedding of C( .J/1, X), where X is minimum
size and all faces of the imbedding are quadrilaterals. At first glance this
bound appears to be obvious, since any cycle of length 2 or 3 in a Cayley
graph for an irredundant generating set must come from a generator of order 2
or 3, and the canonical generating set has no elements of order 2 or 3 under
the given restrictions. However, there are other irredundant generating sets,
and they can have elements of order 2 or 3. The argument needed to eliminate
such generating sets is actually rather delicate.
Theorem 6.1.2 (Jungerman and White, 1980). Let the abelian group .J/1 have
the canonical form !!Em X · · · X !!Em such that r > 1 and m; > 3 for all i. Then
I '
y(.J/1);;::: 1 + #.J/1· (r- 2)/4.
Proof Let X be a generating set for .J/1 such that y(C(.J/1, X))= y(.J/1).
As we have already observed, it may be assumed that X is irredundant. It
cannot be assumed that X is canonical or even that #X= r.
Let s be the number of generators in X of order at least 4, let t 2 be the
number of order 2, and let t 3 be the number of order 3. We first claim that
s + t 2 ;;::: r and s + t 3 ;;::: r. To show that s + t 3 ;;::: r, let !14 be the subgroup of
.J/1 consisting of all elements of order 2. Then b E !14 if and only if the ith
coordinate of b is 0, when m; is odd, and either 0 or m;/2, when m; is even.
Let <f>: .J/1-+ !2 be the natural quotient homomorphism whose kernel is !14.
Clearly, the group !2 is isomorphic to !!En X · · · X !!En , where n; = m; if m; is
I '
odd and n; = m;/2 if m; is even. In particular, the group !2 has rank r, since
m; > 2 for all i. Therefore, the set </>(X) must contain at least r nonidentity
The Genus of Abelian Groups 255
elements, because <P( X) generates fl. Since every element of order 2 is sent by
<P to the identity, it follows that s + t 3 ~ r as claimed. A similar argument
using the subgroup of elements of order three establishes the other inequality
s + t 2 ~ r.
Now suppose that C( .J/1, X) ~ S is an orientable imbedding. Let /; be the
number of i-sided faces. Then
because digons and triangles arise only from elements of order 2 or 3 in the
irredundant generating set X. Since '£if;= 2#£, it follows that
Therefore
Hence
x(S) = #V- #E + #F
~ #V- #E/2 + #.Jil(t 2 + t 3/3)/4
= #.91- #.9/(s + t 2 + t 3 )/2 + #.Jil(t 2 + t 3/3)/4
~ #.91(1 - (2s + t 2 + t 3 )/4)
Theorem 6.1.3 (Jungerman and White, 1980). Let the abelian group .91 have
the canonical form .2'm X • · · X .2'm , where r > 1 and every m; > 3, and such
1 '
that either every m; is even or r > 3. Then y(J/1) = 1 + #.9/(r- 2)/4,
whenever the number on the right-hand side of this equation is an integer.
256 The Genus of a Group
a
Figure 6.4. A voltage graph and its dual current graph.
Proof The proof is broken down into various overlapping cases. In each
case, a generating set X is given such that #X= r and a quadrilateral,
orientable imbedding for the Cayley graph C( d, X) is constructed. In all but
Case 1, the imbedding is obtained by a voltage-current-graph construction,
using a quotient graph of C( d, X) corresponding to a certain subgroup 114.
=
Case 3. r 2 mod 4, m 1 odd.
Let X be the canonical generating set for d. Let 114 be the subgroup
containing every element of d whose sum of coordinates is congruent to
0 mod m 1 . Then the quotient group dj 114 is isomorphic to ~m 1 , and each
directed edge of the Schreier coset graph S( d: 114, X) originating at vertex j
The Genus of Abelian Groups 257
--n TI: Q
r 1 2 3 r- 2
-1
m-1
9d 1 r r-1 r-2
9dm-}, 1 -1
• r
2. r-2
•• •• ••
• •
r-1
•••
•••
Figure 6.5. The current graph for Case 3. Hollow vertices have counterclockwise rotation and
solid vertices clockwise rotation.
.c?itu,
2 1 2 3 4 3 4 5 6 5 6 7 8 r-1 r-2
••• xx~ ~
1
2 1 2 3 4 3 4 5 6 5 6 7 8 r-1 r-2 -
!JdoJ, .c?itwi
u~u,
2 1 r 2 3 4 3 4 5 6
-1
••• -1
r-1 r-2 '.,.,..c?itw,
"'j+ I
1 4
n~u,
2 r 2 3 r 4 3 5
6
1
• • • 1
8ito11 2 3
8ito1 1 " -
2
, c.. ..., ... , '+ ,J 6
·nt
•••
r-1
1
r-2 ~-!Jdw~
Figure 6.6. The current graph for Case 4. Rotations at hollow and solid vertices as in Figure 6.5.
The Genus of Abelian Groups 259
It can be verified from this rotation scheme that all faces are quadrilaterals
satisfying KVL. For example, start with edge 1 (i.e., x 1 ) at vertex !?.4m_ 3 • That
edge terminates at vertex !14m_ 2 , where we obtain the next edge, 6. That edge
takes us to vertex !JBm-t· There we obtain edge I, which returns us to vertex
!?.4m_ 2 , where we find edge 6 awaiting to return us to !?.4m_ 3 • The next edge is 1
again, so the face is complete (since m - 3 is even, the rotation for !?.4m_ 3
is given by the first row). The face is a quadrilateral with net voltage
x 1 + x 6 - x 1 - x 6 = 0.
Figure 6.7. The current graph for Case 5. Rotations at hollow and solid vertices as in Figure 6.5.
260
The Genus of Abelian Groups 261
j = m- 1:
(j,O,O). 76176541543232
(),0,1). 76176541543232
(),1,0). 45145671672323
(),1,1). 45145671672323
As in previous cases, it can be verified from the scheme above that every
face is a quadrilateral satisfying KVL. For example, starting with edge 1 at
vertex (m- 3,0, 1), one obtains next the edge 7 at vertex (m- 3, 1, 1), then
edge I at vertex (m - 2, 1, 1), then edge 7 at vertex (m - 3, 1, 1), then edge 1
at vertex ( m - 3, 0, 1) again. We notice that m - 3 is even, so that rotations
come from the first and third groups.
(j,O). 76171632325454
(),1). 61716745452323
(),2). 61716745452323
(),3). 76171632325454
262 The Genus of a Group
j odd, j *m- 3:
(j,O). 61716745452323
(J, 1). 76 I 11 63 2 3 254 54
(},2). 76171632325454
(},3). 61716745452323
j = m- 2:
(j,O). 54514767163232
(},1). 67167451452323
(J' 2). 61 76 7 4 1 5 4 5 2 323
(},3). 54154761763232
j = m- 1:
(j,O). 54514761763232
(},1). 61767415452323
(},2). 67167451452323
(},3). 54514767163232
(},2). 891897817823234545
Theorem 6.1.4 (Mohar et al., 1985; Brio and Squier, 1986). The genus of the
group fr3 X !Z3 X fr3 is 7. D
and for about a decade, it was commonly suspected that the genus was 10,
based on the reasoning that a lower genus could not be realized with a
The Genus of Abelian Groups 263
6.1.4. Exercises
1. Construct a voltage-graph imbedding (or current graph) for Case 1 of
Theorem 6.1.3 using the quotient group .2"2 X · • · X .2"2 of rank r.
2. Verify that conditions i and ii hold for the current graph in Case 2 of
Theorem 6.1.3. Give the table of logs for this current graph.
3. Check that KCL holds at every vertex of the current graph given for
Case 3 of Theorem 6.1.3. Verify at least three rows in the associated table
of logs given in the text.
4. Same as Exercise 3, but for Case 4 of Theorem 6.1.3.
5. Same as Exercise 3, but for Case 5 of Theorem 6.1.3.
6. Prove that y(.2"m X .2"n) = 1, where m divides n and m > 2 (see also
Exercise 6.4.14 for the case m = n = 3).
7. Use a one-vertex voltage graph imbedded in the sphere to obtain a
toroidal imbedding of a Cayley graph for .2"3 X .2"3 •
8. Let X = {x, y} be an irredundant generating set for .2"3 X .2"3 • Give a
rotation scheme for an imbedding of C( .2"3 X .2"3 , X) that has six
triangles and at least one quadrilateral. (Hint: Draw a rectangle xyx- 1y- 1
bordered by a pair of x 3 triangles and a pair of y 3 triangles. This gives
you the full rotation at the four vertices of the rectangle and part of the
rotation at four more vertices. Now complete the rotation system so that
the remaining x 3 and y 3 circuits are faces.) Given that y(.2"3 X .2"3 ) = 1,
the imbedding constructed in this way must be toroidal. Why? What are
the sizes of the faces?
9. Use one-vertex voltage graphs to construct three different imbeddings of
a Cayley graph for .2"3 X .2"3 X .2"3 in the surface S 10 : one with 9
triangles, one with 18, and one with 27.
10. Let X = {x, y, z} be an irredundant generating set for .2"3 X .2"3 X .2"3 •
Let !14 be the subgroup generated by z. The Schreier coset graph
264 The Genus of a Group
In attempting to determine the genus of a given Cayley graph C(.Jil, X), one
strategy is to suppose that there exists a minimum-genus imbedding that
somehow reflects the symmetry of the graph. In particular, the natural action
of the group .91 on the Cayley graph C( .s#, X) might extend to an action of .91
on the imbedding surface. If the natural action does so extend, then we call the
imbedding "symmetric". Furthermore, if the natural action is orientation
preserving, then we call the imbedding "strongly symmetric" and, otherwise,
"weakly symmetric". The "symmetric genus" a(.s#) (resp., "strong symmetric
genus" a 0 ( .91 )) is then defined to be the smallest number n such that the
surface sn contains a symmetric (resp., strong symmetric) imbedding of a
Cayley graph for the group .91. A main result of this section is that if the group
.91 acts on an orientable surface S, then there is a symmetric imbedding in S
of a Cayley graph for .s#. It follows from this result that the symmetric genus
could be defined equivalently without any reference at all to generating sets or
Cayley graphs.
The Symmetric Genus 265
Example 6.2.1. The dihedral group !i)4 is given by the presentation (x, y:
x 4 = y 2 = 1, yxy- 1 = x- 1 ), with generating set X= {x, y}. Consider the
planar imbedding of the Cayley graph C(!i)4 , X) given on the left in Figure 6.8.
Edges corresponding to x are solid, and edges corresponding toy are dashed.
Every vertex has the (clockwise) rotation xx- 1y- 1y.
'~"'...
, 'I
1:,4/ I
I
I
I
~
i..,~ ~~,
'
Figure 6.8. An imbedding of a Cayley graph in the sphere.
266 Tile Genus of a Group
Theorem 6.2.1. An imbedding C(.J/1, X) --+ Sis strongly symmetric if and only
if the rotation at every vertex is the same. If an imbedding C( .J/1, X) --+ S is
symmetric, then the rotation at every vertex is the same or reversed.
6.2.2. Reflections
If the group .91 acts pseudofreely on the surface S, then the methods of
Chapter 4 involving branched coverings and voltage graphs suffice to describe
the action of .91 on S. However, some group actions on surfaces are not
pseudofree. Under the most general conditions, the fixed point set of a surface
homeomorphism can be very complicated, and one might fear that nonfree
actions are completely intractable. Fortunately, the behavior of a surface
homeomorphism of finite order is fairly simple, which is reassuring since our
interest is in finite groups of homeomorphisms. For example, the main
theorem of this section asserts that an orientation-preserving surface homeo-
morphism of finite order greater than 1 has only a finite number of fixed
points.
We have already called attention, in passing, to some group actions that are
not pseudofree, such as the action discussed just above of fl'2 X fl'4 on the
sphere, which involves reflection across the equator. For the sphere, reflection
in the equator is, in effect, the only finite-order homeomorphism having an
infinite number of fixed points. For a surface of higher genus, it is easy to
visualize a similar reflection using a plane in 3-space to split the surface in
"half". There are, however, other "reflections" that are not so easily visual-
ized.
Example 6.2.2. Let S be the torus given by the rectangle with sides to be
identified, as in Figure 6.9. Note that after the top and bottom are identified, to
form a cylinder, the left and right sides are identified by a half-twist. Let f be a
reflection of the rectangle about the center broken line. Then f induces a
homeomorphism h on the torus S. The broken line becomes a circle C left
pointwise fixed by h, whereas the line v u v becomes a circle C' that is taken to
itself by h but is not left pointwise fixed-it is turned 180°, taking u to v and v to
u. The circles C and C' split the torus S into halves that are interchanged by h.
On the right of Figure 6.9 is the torus with edges identified. The symbols "F"
u u
G1~
l I
I
u I u
I
I
u
F : 1u
Figure 6.9. A reflection of the torus.
The Symmetric Genus 269
and "G" are given, together with their images under the reflection, to show how
the homeomorphism h acts on the torus.
i. h is an orientation-reversing involution;
ii. S is the union of two connected surfaces with boundary S0 and S 1 such
that h(S0 ) = S1 and S0 n S1 is a collection of disjoint circles im-
bedded in S, at least one of which is left pointwise fixed by h.
The surfaces S0 and S1 are called the "halves" of the reflection h, and the
components of S0 n S1 are called the "dividing circles".
It can be shown (Exercise 3) that x(S) = 2x(S0 ). Since there are only
finitely many surfaces with boundary having a given Euler characteristic, the
choice of halves for reflections on a given surface is limited. For a torus, there
are only two kinds of reflections: the one given in Example 6.2.2 and the
obvious reflection that has the same halves as Example 6.2.2 but leaves both
dividing circles pointwise fixed. Reflections for a higher-genus surface are
considered in Exercise 4. The main result of this section can now be stated.
Proof The proof in the general topological case, due mostly to Kerekjarto
(1921), with later corrections by Eilenberg (1934), is very difficult and is not
considered here. If one assumes that the given homeomorphism is an automor-
phism of a graph imbedding that behaves "piecewise linearly" on the faces,
then the proof follows from easily verified facts about linear transformations
(see Exercises 5-10). D
z z
y r
X
Figure 6.10. A surface of genus 2 and a quotient surface.
The orientation-preserving subgroup .91° consists of the identity and the 180°
rotations ab, ac, be about the z, y, x axes, respectively.
To visualize the quotient surface Sjs# 0 and the quotient action of s#js# 0 ,
we proceed as follows. The surface S is divided into eight pieces by the three
coordinate planes. We keep two adjacent pieces to form a surface S' (with
boundary) and discard the rest of S. On the right of Figure 6.10 is shown the
case in which S' consists of all points of S satisfying x .::;; 0 and y .::;; 0. The
restriction of the quotient map p: S ~ S/.91° to S' still maps onto Sjs# 0 ,
since the subsurface S' contains a point from every orbit of the action of .91°.
Moreover, the restriction of the covering projection p to the interior of S' is a
homeomorphism, because that part of S' contains only one point from each
orbit. On the other hand, the quotient map identifies pieces of the boundary of
S'. Rotation ac about the y axis identifies the top and bottom halves of the
circle in the yz-plane on the left and the semicircle in the yz-plane on the
right, as indicated by double arrows. Rotation be about the x axis identifies
the top and bottom halves of the semicircle in the xz-plane, as indicated by
single arrows.
Thus the quotient map closes off the holes in S' to form a sphere for the
quotient surface S/.91°. The branch points of the covering p: S ~ Sjs# 0 are
indicated by heavy dots: r, s, t correspond to fixed points of rotation ac about
the y axis, u to fixed points of rotation ab about the z axis, and v to fixed
points of rotation be about the x axis. The sphere Sjs# 0 has an equator given
by the arcs rs (semicircles identified), st (half of a dividing circle for reflection
in the xy-plane), tu (quarter-circles identified), uv (quarter-circles identified),
and vr (half of a dividing circle for reflection in the xy-plane). The action of
the quotient group s#js# 0 on the quotient surface Sjs# 0 is simply reflection
in this equator.
Theorem 6.2.3 (Sabidussi, 1958). Let the group .91 act on the graph G so that
the action on the vertex set is free and transitive. Then G is a Cayley graph for
.J/1, in which edges in the same orbit of the action correspond to the same
generator of .J/1. Moreover, any edge left invariant by a nonidentity element of .91
corresponds to an involution in the generating set.
Proof Choose any vertex v, and label it by the group identity 1. For each
a E.!#, label the vertex <l>a(v) by a. If <Pu(v) = <Pb(v), then <Pab_,(v) = v,
which implies that ab - 1 = 1, since .J/1 acts freely on the vertex set. Thus, there
are no conflicts in the labeling. Moreover, every vertex receives a label, since
the action of .91 is transitive on vertices.
Next, choose an edge e incident on the vertex v, and suppose that its other
endpoint u is labeled with the group element x. Then assign the label x to
every edge in the orbit of e. In addition, if <Px(e) * e, then for each a E.!#,
direct the edge <Pa( e) so that its initial vertex is <Pa< v ); if <PA e) = e, do not
assign any directions to the edges labeled x. Observe that the edge <Pa( e),
which is labeled x, leads from vertex <Pa(v), which is labeled a, to the vertex
<Pa(u) = <Pa(<PAv)) = <PaAv), which is labeled ax. The only way a conflict in
assigning directions could arise would be if there were a group element b such
that <Pu{e) = <Pb(e) and <Pa(v) * <Pb(v). But then <Pab_,(e) = e, which implies
that <Pab_, ( v) = u, which in tum implies that ab - 1 = x; and therefore, we
have <PA e) = e, which implies that no direction should be assigned, thereby
ending the conflict. After edges in the orbit of the edge e are labeled and
directed, choose another edge incident on v, and then label and direct the
edges in its orbit. Continue in this manner until all edges incident on v have
been labeled. Since the action of group .91 is transitive on vertices, it follows
that all edges of the graph G have been labeled. As has already been observed,
this labeling transforms G into a Cayley color graph for .J/1.
The given labeling certainly assigns the same generator to every edge in the
same orbit. We still must show that if e is labeled x and if <Pa(e) = e for some
nonidentity element a, then x is an involution. So suppose that the initial
endpoint of e is labeled b. We may infer that ab = bx and that a(bx) = b,
since <Pa switches the endpoints of e. It follows that b = a(bx) = (ab)x =
bx x, and therefore that x 2 = 1. 0
The Symmetric Genus 273
Theorem 6.2.4 (Tucker, 1983). Let the finite group .511 act on an orientable
surface S. Then there is a Cayley graph for .511 cel/ularly imbedded in S so that the
natural action of .511 on the Cayley graph extends to the given action of .511 on the
surfaceS.
f(w)
w=f(w)
Figure 6.11. The imbedding of half of Gin T (black dots are branch points).
274 The Genus of a Group
circle C;, choose a finite set W; of points inC;, such that each component of
C; - W; contains exactly one branch point; if C; contains no branch points, let
W; consist of a single point in C;. Then, for each point w in W;, imbed an arc
in T running from v tow, and, if f(w) * w, then imbed another arc from v to
f(w) (see the right of Figure 6.11). Let G0 be the resulting collection of loops
and arcs in T, and let G = G0 U /( G0 ). Then G is a graph with two vertices v
and f(v) (the points in ware now just midpoints of edges). By construction, it
follows that f(G) = G and that the imbedding of G is cellular with at most
one branch point inside each face. D
Corollary. Let .91 be a finite group. Then the symmetric genus o(.91) is the
minimum genus over all surfaces on which .91 acts, and the strong symmetric
genus o 0 (.91) is the minimum genus over all surfaces on which .91 acts preserving
orientation. D
Example 6.2.4. Let .91 be a finite group acting with reflections on the surface S
such that the quotient surface Sj.91° is the sphere and such that the covering
p: S-+ Sjd 0 has three branch points in S/.91°, one of which lies on the
dividing circle of the reflection f of the quotient action of .91j.91° on S j.91°. Since
f takes branch points to branch points, it follows that the remaining two branch
points must be paired by f. In particular, they have the same order. The
appearance of graph G is then as illustrated in Figure 6.12 The generator z
corresponding to the edge e through the dividing circle off must be an involution,
by Theorem 6.2.3, because at least one edge of its preimage p - 1( e) passes
through a circle left pointwise-fixed by a reflection of .91. Since the order of the
net voltage on each face equals the order of the branch point inside that face, the
group .91 can be presented
(x, z: z 2 = 1, xq = 1, (xzx- 1z )' = 1, ... ),
where r is the order of the branch point on the dividing circle and where q is the
order of the other paired branch points.
The Symmetric Genus 275
X
v u
y y
s
X
Figure 6.13. A quadrilateral in an imbedding of a Cayley graph C( JJI, X), where JJI is abelian.
Since the imbedding is symmetric, each of these rotations must be the same
or reversed. Since the reverse of ... xy . .. is ... yx . .. , the rotation at t is the
same as the rotation at s. By similar reasoning, the rotations at u and v must
be the same as those at t and u. From the information given at s, t, and u,
this rotation must be ... xyx- 1y- 1 • • • • The information given at vertex v
implies that x immediately follows y- 1 in this same rotation, but that is
impossible because z and z- 1 must appear somewhere in the rotation. We
276 The Genus of a Group
Example 6.2.5. Consider the pattern of bricks shown in Figure 6.14, which the
reader has probably seen in a courtyard or patio. Imagine the pattern extending
infinitely in all directions in the plane. What rigid motions, that is, Euclidean
isometries of the plane, take this pattern to itself? First, one can simply
"translate" the whole pattern, so that, for instance, the set of four outlined
bricks around point u displaces those around point v. Second, one can "rotate"
the pattern 90° about the point u (note that this rotation takes the entire global
pattern onto itself, not just the four local bricks). Third, one can "reflect" the
pattern about a horizontal or vertical line such as k or I. Finally, one can
translate half the distance from u to v, at the same time reflecting about the line
through u and v (this last type of motion is called a "glide"). The whole pattern
can be generated as the orbit under rigid motions of a single half-brick like that
shaded near the point w in Figure 6.14.
The collection of all Euclidean isometries leaving the pattern in Figure 6.14
invariant forms a group ~. called the "symmetry group" of the pattern. This
group has two important properties:
k-- 1- - - k
u ~w
Another way of stating property 2 is that for any point u, there is a minimum
distance d such that every isometry in the group C(l either moves u at least the
distance d or does not move u at all. Any group of isometries of the Euclidean
plane satisfying properties 1 and 2 is called a (two-dimensional) "Euclidean
space group" (or alternatively a Euclidean "crystallographic group"). Every
pattern in the Euclidean plane having translational symmetry in more than
one direction and having no infinitesimal symmetry (N .B.: a pattern of vertical
stripes has infinitesimally small vertical translations as symmetries) has a
Euclidean space group as its symmetry group. Conversely, every Euclidean
space group has patterns for which it is the symmetry group: one first finds a
"fundamental domain" like the half-brick in Figure 6.14, next draws any motif
on this domain, and then replicates the motif throughout the plane using the
given space group.
The history of space groups, from the art of the Alhambra and of M. C.
Escher to the crystallography of Federov (1891) to the mathematics of
Bieberbach (1911), is too rich to elaborate here. The reader is urged to read
Weyl (1952) or Schattschneider (1978) for an informal account. Lyndon (1985)
and Coxeter and Moser (1957) provide the mathematical details. The im-
portant facts we need about Euclidean space groups are summarized in the
following theorem.
Theorem 6.2.5. There are exactly 17 Euclidean space groups (up to group
isomorphism), 5 of which contain only orientation-preserving isometries. The
collection !T of all translations contained in a given Euclidean space group C(l is a
278 The Genus of a Group
normal subgroup of C(l, is isomorphic to !ZX :Z, and has as quotient C(ljY either
the trivial group or one of the cyclic groups fr2 , fr3 , fr4 , fr6 or one of the
dihedral groups !!)2 , !!)3 , !!)4, !!)6 •
Proof Omitted. D
Theorem 6.2.6. Let C(l be a Euclidean space group and ..#' any normal subgroup
of finite index in C(l. Then the quotient group C(/j..ff acts on the torus or on the
sphere.
Alternatively, if ..#' 0 * ..#', then ..#'~ 0 = ~. since the subgroup ~ 0 has index
2 in ~. By the fundamental isomorphism theorem for quotient groups, we
know that ~ 0/..#' 0 is isomorphic to ..#'~ 0/..ff. Since ~ 0/..#' 0 acts on Sand
..#'~ 0 = ~. it follows that ~/..#' acts on the torus or on the sphere. D
One can also define "spherical space groups" using spherical geometry and
"hyperbolic space groups" using hyperbolic geometry. In both cases we
require the group to be discrete. For the sphere, which is a compact surface,
this automatically means the group is finite. For the hyperbolic plane, we
require the quotient space under the group action to be compact, in analogy to
condition 1. In either case, there is no subgroup analogous to the translation
subgroup !T for Euclidean space groups.
as the fundamental triangle and half the opposite; the forward and backward
letter Fs tell which is which. The group ~ is the symmetry group of the pattern
and hence is a Euclidean space group.
These are in fact the only relators needed. We sketch a proof. First, note that
the dual graph to the triangular pattern for a ( p' q' r) triangle group re is a
Cayley graph for re, since re acts on the dual graph transitively, fixing no
vertex (i.e., fixing no triangle in the original pattern). Any relator in the
generators corresponds to a closed walk in this Cayley graph. Since the Cayley
graph is imbedded in the plane, every cycle, and hence very closed walk, can
be expressed in terms of the face boundary cycles, which simply correspond to
the relators (xy)P = (yz)q = (xz)' = 1. Thus, every relator can be expressed
in terms of the already given relators. (We are being vague about what we
mean by "expressed", but this can be made rigorous.)
Given a full (p, q, r) triangle group re, the subgroup ?A generated by the
rotations xy, yz, and xz at the three vertices of the fundamental triangle is
called the "ordinary" (p, q, r) triangle group. Since (xy)(yz) = xy 2z = xz,
this group is generated by u = xy and v = yz alone. It has the presentation
Moreover, this subgroup has index 2 in the full triangle group [the map
f:re--+ ?!'2 given by f(x) = f(y) = f(z) = 1 has kernel ?A], and thus ?A
coincides with the orientation-preserving subgroup re 0 .
We summarize these facts in the following theorem.
6.2.9. Exercises
1. Prove that an imbedding of a Cayley graph C(~. X) is weakly symmet-
ric if and only if there is a subgroup !11 of index 2 in the group ~ such
that every vertex of the Cayley graph labeled by an element of !11 has the
same rotation and all other vertices have the reversed rotation.
2. Let S be a surface with r boundary components. Then x(S) is defined
to be x(S) - r, where S is the closed surface obtained by spanning each
boundary component with a disk. Find all orientable surfaces S with
boundary such that x(S) = -5.
3. Let S0 be one-half of a reflection of the surface S. Prove that x(S) =
2x(So)·
4. Find all reflections of the surface of genus 2.
5. Given simplexes CJ in R m and T in R n' a map f: CJ ~ T is said to be
"linear" iff takes vertices of a to vertices of -r, and iff is the restric-
tion to a of an affine map F: Rm ~ Rn [i.e. F(x) = L(x) + b, where
L: R m ~ R n is a linear transformation and b E R n]. Find a linear map
from the simplex a in R 2 having vertices (0,0), (1,0), (0, 1), to the
simplex -r in R 3 having vertices (1, 2, 1), (3, 5, 2), (0, 2, 8). Prove that a
linear map between simplexes is uniquely determined by what it does to
vertices.
6. Given simplicial complexes K and L, a continuous map f: IKI ~ILl is
"piecewise linear" if there are subdivisions K', L' of K, L such that f
restricted to each simplex of K' is linear ( K' is a subdivision of K if
IK'I = IKI and every vertex of K is a vertex of K'). Let K be the
simplicial complex in R 2 consisting of the 2-simplex with vertices (0, 0),
(0, 1), and (1, 0) (and all its faces). Find a piecewise-linear map
f: IKI ~ IKI that leaves fixed the two legs of the right triangle IKI and
takes the point (1/3, 2/3) to the point (2/3, Ij3).
7. The set of fixed points of a piecewise-linear homeomorphism can be quite
complicated. Let L be the two-dimensional simplicial complex in R 2
formed by four right triangles with vertices at (0, 0), ( ± 1, 0), (0, ± 1). Use
the function f of Exercise 6 to construct a piecewise-linear homeomor-
phism g: ILl~ ILl that leaves fixed all points of ILl on the x axis, on the
y axis, or in the second quadrant. Sketch ILl and its set of fixed points.
Does g have finite order?
8. Prove that if h: R 2 ~ R 2 is a linear transformation of finite order leaving
the x axis fixed, then either h is the identity or h is a reflection across the
x axis.
9. Let g be an orientation-preserving, piecewise-linear homeomorphism
g: ILl ~ ILl, where ILl is a surface. Use Exercise 8 to prove that if g has
finite order and leaves some 1-simplex of L fixed, then g is the identity.
10. Use Exercise 9 to prove Theorem 6.2.2 for piecewise-linear homeomor-
phisms.
Groups of Small Symmetric Genus 283
11. Prove that t¥a"b = t¥at/lt,, where t/1 defines the natural quotient action of
the quotient group dj!!l on the quotient surface Sj!!l.
12. Suppose that the group .s;l acts on a surfaceS with reflections so that the
quotient surface Sj.s;/ 0 is a surface of genus 2. Suppose that the quotient
action of dj.s;/ 0 on Sjd 0 is a reflection with a single dividing circle
and that the branch points of p: s--+ Sj.s;/ 0 have orders and locations
as given by Figure 6.16. Find a partial presentation for the group .s;l by
constructing a 2-vertex voltage graph as in the proof of Theorem 6.2.4.
13. Prove that a Cayley graph imbedding C( .s;l, X) --+ S is strongly symmet-
ric if and only if it is the derived imbedding of a 1-vertex imbedded
voltage graph.
14. Prove that a 0 Ul'2 X .2'4 ) = 1 (see also Theorem 6.3.1).
15. List all the rotations used in the quadrilateral imbedding given in Section
3.5 for the Cayley graph C( d, X), where .s;l = .2'10 X .2'10 X .2'10 and
X= { x, y, z }, a canonical generating set for d.
16. Give an example of an involution on the torus that is not a reflection.
17. Let .'T be a discrete group of translations of the Euclidean plane
containing translations in independent directions. Choose any point P.
Let Q be the closest point to P in the orbit of P, and let R be the closest
point to P in the orbit of P but not on the line through P and Q. Prove
that the translations by vectors PQ and PR generate .'T.
18. Let ct' be a Euclidean space group, and let r E ct' be a rotation about
some point P. Let Q be the closest point to P in the orbit of P. Prove
that if r is a rotation through an angle of less than '1T j3, then r( Q) is
closer to Q than Q is to P. Why does this contradict the choice of Q?
Use a similar argument to show that r is not a rotation of order 5 either.
A first project in studying the genus of groups is to identify all the groups of
genus 0, then maybe all the groups of genus 1. We shall see that the analogous
problem for symmetric genus happens to be a quite tractable problem, and
that preliminary consideration of the symmetric genus gives remarkably accu-
rate insight into the general case, whose derivation is otherwise rather lengthy.
Study of the symmetric genus is additionally rewarding, in that it also reveals
rich connections with Euclidean and non-Euclidean geometry. Moreover, the
moderately easy success in determining all the groups of symmetric genus 0 or
284 The Genus of a Group
The key observation about the order of a branch point is that if the regular
branched covering p: S -+ Sj.Jil corresponds to an imbedded voltage graph,
then the order of each branch point is the order of the net voltage on the face
containing that branch point.
Theorem 6.3.1. For a finite group .J/1, the strong symmetric genus a 0 ( .J/1) is 0
if and only if either the group .J/1 is isomorphic to one of the cyclic groups fl'n or
to one of the dihedral groups f)n, or else the group .J/1 has one of the following
presentations and the given order:
Proof If a 0 ( ..l11) = 0, then the group J1l acts on the sphere S preserving
orientation. By the corollary to Theorem 6.2.2, the action is pseudofree. Let
p: S --+ S j..l11 be the associated regular branched covering, with branch set B.
From the second form of the Riemann-Hurwitz equation, we observe that
from which it follows that J1l is the trivial group. We conclude that there are
either two or three branch points.
Suppose, on one hand, that there are two branch points. Then imbed the
1-vertex, 1-loop graph B 1 in the sphere Sj..l11 so that the loop separates the
branch points. By Theorem 4.3.5, the covering p: S --+ Sj..l11 is the derived
covering for some voltage assignments on the imbedded bouquet B 1 into the
group ..l11. Since B 1 has only one edge, the group J1l must be generated by a
single element, that is, the group J1l is cyclic.
Now suppose, on the other hand, that there are three branch points. Let
q 5: r 5: s be the orders of those branch points. The only values for the triple
(q, r, s) that leave the right-hand side of the Riemann-Hurwitz equation
positive are (2, 2, n ), (2, 3, 3), (2, 3, 4), (2, 3, 5). Imbed the bouquet B2 into the
quotient surface Sj..l11 as in Figure 6.17 (in which heavy dots represent the
branch points of the given orders). Again it follows from Theorem 4.3.5 that
the covering p: S --+ Sj..l11 is obtainable by assigning the two loops voltages
x, y E J11, as shown in Figure 6.17. Since the order of the net voltage on each
face equals the order of the enclosed branch point, we infer that the group J1l
may be presented
If (q, r, s) = (2,2, n), then J1l is the dihedral group f!)n· The other three
possibilities for (q, r, s) yield the three presentations given in the theorem.
s
•
Figure 6.17. An imbedded voltage graph.
Groups of Small Symmetric Genus 287
Moreover, in each of these cases the order of .J/1 is determined exactly by the
Riemann-Hurwitz equation. For example, if (q, r, s) = (2, 3, 5), then
so #.J/1 = 60.
We have shown that if o 0 (.Jil) = 0, then .J/1 is as described in the theorem.
The converse, that if .J/1 is isomorphic to !!En or f!)n or has presentations (a),
(b), or (c), then .J/1 acts on the sphere, follows directly from the voltage-graph
constructions already given in this proof. D
Corollary (Maschke, 1896). The finite group .J/1 has strong symmetric genus 0
if and only if it is isomorphic to one of the groups, !!En, !!)n, .J/14 , ~, or .J/15 •
These groups correspond respectively to the ordinary (1, n, n ), (2, 2, n ), (2, 3, 3),
(2, 3, 4), (2, 3, 5) triangle groups.
Theorem 6.3.2. Suppose that the finite group .J/1 acts on the sphere, reversing
orientation. The action has no reflections if and only if .J/1 is isomorphic to ~2 n
288 The Genus of a Group
for some n > 1. The action has reflections if and only if either d is isomorphic to
f) n , ~2 X ~n, ~2 X f) n or d has one of the following presentations and the
given order:
and therefore we conclude there is at most one branch point, lest the right side
become negative. If the order of the branch point is n, then #d= 2n.
Moreover, with a single branch point in the projective plane, the covering
p: S --+ Sjd can be obtained from an imbedded voltage group with a single
type-1 loop. Therefore the group d is cyclic.
Suppose d acts with reflections. The orientation-preserving subgroup d 0 ,
however, still acts pseudofreely. As in Theorem 6.3.1, the quotient surface
SId 0 is again the sphere, and the regular covering p: S --+ S 1d 0 has two or
three branch points. Suppose there are two branch points. As in Theorem
6.3.1, they must have the same order, call it n. The quotient group djd 0 acts
on Sjd 0 as a single reflection. Either both branch points lie on the dividing
circle of this reflection or one lies on one side and one lies on the other side.
The regular covering p: S--+ Sjd 0 can then be obtained by a voltage graph
as illustrated on the left of Figure 6.18 for the first case, and the right of
Figure 6.18 for the second case. In both cases, edges crossing the reflection line
correspond to reflections in the action of d on the sphere S, and hence the
voltages on those edges have order 2. Presentations for the group d are given
by setting the order of the net voltage on each face equal to the order of the
enclosed branch point. In the first case, we get the presentation
I I
I
¢ + 0+-0
X I X
I I
I I
I I
Figure 6.18. Voltage graphs in the case of two branch points.
I
I
X I X
0-f-0 I
I
I
Figure 6.19. Voltage graphs in the case of three branch points.
and thus .91 is the dihedral group p)n. In the second case, we get
The second presentation is, of course, the dihedral group p)n. In the first
presentation, since y commutes with both x and z, and since the subgroup
generated by x and z is the dihedral group f)n, the group .91 must be ~2 X f)n
290 The Genus of a Group
(unless the subgroups (y) and (x, z) have a nontrivial intersection, in which
case y is a redundant generator and .J/1 is simply the dihedral group Edn)·
For the other possible values of (q, r, s), we obtain the presentations
(a)-(d) given in the statement of the theorem. Presentations (a) and (b)
correspond to the two ways of locating branch points when (q, r, s) = (2, 3, 3),
whereas (c) and (d) correspond to the only possible branch-point locations
from (2, 3, 4) and (2, 3, 5). The subgroup that must be of index two in each
presentation is just the orientation-preserving subgroup, which is also the
isotropy group for the given voltage graph. The order of .J/1 is determined as
given in each case, because #.Jil= 2#.91° and #.91° is determined by
Theorem 6.3.1.
Since all the voltage graph constructions given in this proof depend only on
.J/1 having the desired presentation, the conditions .J/1 must satisfy are not only
necessary but also sufficient in order that .J/1 act on the sphere as described.
D
It can be shown (Exercise 6) that the subgroup (x, yxy) in presentation (a)
and the subgroup (xy, yz) in presentations (b)-(d) must have index at most
two for purely algebraic reasons. Since these subgroups are just the groups
(a)-(c) of Theorem 6.3.1, they are automatically finite without any extra
relations. Therefore the presentations (a)-(b) of Theorem 6.3.2 are also
automatically finite without any extra relations. This means again that only
one group is defined by each presentation with the given subgroup restriction.
In Exercise 5 the reader is asked to find the appropriate generating sets to
present ~2 X .J/14 by (a), ~ by (b), ~2 X~ by (c), and ~2 X .915 by (d).
Again, the reader should recognize the full triangle groups appearing in the
proof of the theorem and in presentations (b)-(d). On the other hand, the
pre sen ta tion
(2, 2, n ), the hybrid (2, 3, 3), the full (2, 3, 4), and the full (2, 3, 5) triangle
groups.
Theorem 6.3.3. The group d acts on the torus and preserves orientation if and
only if d has one of the following presentations
The group d acts on the torus and reverses orientation, but without reflec-
tions, if and only if d has one of the following presentations, with the given
subgroup !14 having index 2 in d:
The group .91 acts on the torus with reflections if and only if .91 has one of the
following presentations, with the given subgroup !14 having index 2 in .91:
Proof Suppose .91 acts on the torus S preserving orientation. Then .91
acts pseudofreely. By the Riemann-Hurwitz equation
It follows that x(S/.91) ~ 0 and that S/.91 is the torus or the sphere. If S/.91
is the torus, then there can be no branch points. Presentation (a) is thus
obtained from a voltage assignment for a bouquet of two circles cellularly
imbedded as usual in the torus Sjs#. If Sjs# is the sphere, then, by the
Riemann-Hurwitz equation, there are either four branch points all of order 2
or three branch points. Figure 6.20 illustrates two ways to imbed a bouquet of
three circles in the sphere so that each face of the imbedding encloses one
branch point of order two. The imbedded voltage graph on the left yields
presentation (b 1 ), and the one on the right, (b 2 ). If there are three branch
points of order q:;;; r:;;; s, then, from the Riemann-Hurwitz equation, the
possible values for ( q, r, s) are (3, 3, 3), (2, 4, 4), (2, 3, 6). Presentations (c), (d),
Groups of Small Symmetric Genus 293
z
X
• •
Figure 6.20. Two imbedded voltage graphs for the case of four branch points.
I/
"IE
X '
\
I • • I
\ y I
'' ,. /
y X
y
Figure 6.21. Some imbedded voltage graphs.
u.,.....___:=:;::::~-... u
X :
i
X
z
Iy
i u--==:::;:::-....,. u
Figure 6.22. Two voltage graphs imbedded in the torus.
wise fixed; it is illustrated on the left in Figure 6.22. The other possible
reflection is the one discussed in Example 6.2.2 and is illustrated on the right
in Figure 6.22 (the dotted line down the middle of the rectangle is the dividing
circle left pointwise fixed by the reflection, whereas the left and right sides of
the rectangle form the dividing circle that is taken to itself with a half-rota-
tion). In both cases, the figure shows a 2-vertex voltage graph imbedded in the
torus, invariant under the given reflection, as in the proof of Theorem 6.2.4.
Voltages on edges crossing a dividing circle left pointwise fixed are involu-
tions. These two cases yield presentations (h) and (i).
Finally, suppose that the quotient surface S/.91° is the sphere. As in the
orientation-preserving case, the covering p: S --+ Sjs# 0 has either four branch
points of order 2 or else three branch points. With four branch points of order
2, there are three ways the points can be located with respect to the dividing
circle of the quotient reflection: all four on the circle, two on and two off, all
four off. Figure 6.23 illustrates the corresponding imbedded voltage graphs;
two possible voltage graphs are shown for the case of no branch points on the
dividing circle. These imbedded voltage graphs yield presentations (j), (k), (1 1)
and (1 2 ), respectively. The last five presentations, (m)-(q), correspond to the
case of three branch points, as in presentations (a)-(d) of Theorem 6.3.2, only
here the branch-point orders are (3, 3, 3), (2, 4, 4), or (2, 3, 6).
The converse, that any group .91 having one of the presentations (a)-(q)
with the given subgroup restrictions must act on the torus as described,
Groups of Small Symmetric Genus 295
X X
X X
y y
Figure 6.23. Imbedded voltage graphs for four branch points of order 2.
Corollary. If the finite group .91 acts on the torus, then .91 is a quotient of a
Euclidean space group. The presentations a-q of Theorem 6.3.3 correspond
respectively to the groups p1, p2, p3, p4, p6, pg, pgg, pm, em, pmm, cmm,
pmg, p3m1, p31m, p4m, p4g, p6m.
Theorem 6.3.4. If the group .91 has one of the presentations (f 1 )-(q), then the
given subgroup !!J has index at most 2, and if !!J = .91, then .91 also has one of
the presentations (a)-(e). Thus, if .91 has any of the presentations (a)-(q), with
or without subgroup restrictions, then o( .91) :;;; 1. Finally, if .91 has one of the
presentations of Theorem 6.3.3, then .91 has a normal, 2-generator, abelian
subgroup !T such that s#j!T is trivial or one of the following groups:
fl'2(b, f, h, i) I!JJ2 (g, j, k, I)
fl'3( c) EP 3 (m, n)
fl'4(d) I!lJi o, p)
fl'6 (e) I!JJ6(q)
Theorem 6.3.5 (Hurwitz). Suppose the group .91 acts on the surface of genus
g > 1. Then #.91:;;; 168(g- 1).
Groups of Small Symmetric Genus 297
Proof Let p: S--+ Sjd 0 be the regular covering associated with the
pseudofree action of the orientation-preserving subgroup d 0 • By the
Riemann-Hurwitz equation,
The value of 1/q + 1/r + ljs- 1 next nearest to 0 is - f4, occurring for
(q, r, s) = (2, 3, 8). Thus the proof of Theorem 6.3.5 leads to the following
corollaries.
Corollary. Suppose o 0 (d) > 1. Then #d> 48(o 0 (d)- 1) if and only if
#d= 84(o 0 (d)- 1) and d has the presentation
Conder (1980) has shown that the symmetric group Y, has a presentation
of the form given in the second corollary for all n > 167. Thus o(Y,) =
n!j168 + 1, n > 167. He has also shown that the alternating group dn has a
presentation of the form given in the first corollary for all n > 167. Thus
o(dn) = o 0 (dn) = n!j168 + 1, n > 167. Conder (1984) has also computed
the symmetric genus of all the other symmetric and alternating groups. Glover
and Sjerve (1986) have computed the strong symmetric genus of the simple
groups PSL(2, p ), where p is a prime.
298 The Genus of a Group
Theorem 6.3.6 (Refined Hurwitz Theorem). Suppose o(.91) > 1 and #.91>
24( o ( .91) - 1). Then one of the following holds:
Moreover, if #.91> 48(o(.91)- 1), then .91 is either a quotient of the ordinary
(2, 3, 7) triangle group or a proper quotient of the full (q, r, s) triangle group,
where (q, r, s) is (2, 4, 5), or (2, 3, s) with 7 5: s 5: 11.
6.3.6. Exercises
1. Prove that the group (x, y: y 2 = xn = yxyx = 1) has order 2n by pro-
ving that the subgroup ( x) is normal of index 2. Prove that the group
(x,y: x 2 = y 2 = (xyy = 1) has order 2n by proving that the subgroup
( xy) is normal of index 2.
2. Prove that the group (x, y: x 3 = y 2 = (xy) 3 = 1) has order 12 by show-
ing that ?A, PAy, ?Ayx- 1 are all of the right cosets of the subgroup
!II= (x). (Hint: Use yxy = x- 1yx- 1 and yx- 1y = xyx.)
3. Prove that the group (x, y: x 4 = y 2 = (xy) 3 = 1) has order 24 by show-
ing that !II, !!Byx', !!Byx 2y, r = 0, 1, 2, 3, are all of the right cosets of the
subgroup !II= (x).
4. Prove that the group (x, y: x 5 = y 2 = (xy) 3 = 1) has order 60 by show-
ing that !II, !!Byx', ?Ayx 2yx', ?Ayx 2yx- 2y, r = 0, 1, 2, 3, 4, are all of the
right cosets of the subgroup ?A= (x). (For example, !!Byx- 2y is in this
list of cosets because !!Byx 2yx is in the list, and ?Ayx 2yx = ?Axyxxyx =
!!Byx-1yyx-1y = ?Ayx-2y.)
Groups of Small Symmetric Genus 299
5. Find generators for fr2 X .914 , 94, fr2 X 94, fr2 X .915 that give, respec-
tively, presentations (a)-(d) of Theorem 6.3.2.
6. Prove that the subgroup !14 = (xy, yz) has index at most 2 in the group
.91= (x, y, z: x 2 = y 2 = z 2 = 1, ... ). Do the same for the subgroup
!14 = (x, yxy) in the group .91= (x, y: y 2 = 1, ... ).
7. Prove for presentations (f 1 )-(g 1 ) of Theorem 6.3.3 that the given sub-
group !14 is of index at most 2.
8. Same as Exercise 7 for presentations (h)-(1 2 ) of Theorem 6.3.3.
9. Prove that if !14 = .91 in presentations (f 1 ), (f 2 ), (h), or (i) of Theorem
6.3.3, then .91 has presentation (a).
10. Prove that in Theorem 6.3.3 if !14 = .91 in presentations (g 1 ), (g 2 ), (j), (k),
(1 1 ), then .91 has presentation (b 1 ), and if !14 = .91 in presentation (1 2 ),
then .91 has presentation (b 2 ).
11. Show that presentations (m)-(q) of Theorem 6.3.3 correspond, respec-
tively, to presentations (c), (c), (d), (d), (e) when !14 = .91.
12. Verify that the normal subgroup !T of Theorem 6.3.4 is given by
(x- 1y, xy- 1 ) for presentation (c), (xyx, x 2 y) for (d), and
(xyx- 1y, x- 1yxy) for (e).
13. Verify that the normal subgroup !T of Theorem 6.3.4 is !14 itself in
presentations (f 1 ), (h), (i).
14. Verify that the normal subgroup !T of Theorem 6.3.4 is given by
(xyx- 1y-I, x- 1yxy) for presentation (g 2 ), (xz, yw) for (j), (xyxy, yxz)
for (k), and (x, yz) for (1 2 ).
15. Verify that the normal subgroup !T of Theorem 6.3.4 is given by
(yxyz, xyzy) for presentations (m) and (o), (x- 1yx- 1y, xyxy) for (n)
and (p), and ([yz, xy], [zy, xy]) for (q).
16. Determine o(.P) and o- 0 (2), where 2 = (i, j: i 4 = 1, i 2 = ) 2 ), the
quaternion group of order 8.
17. Find o( .91) and o 0 ( .91) for all groups of order #.91 < 16.
18. * Determine o 0 ( .91) for all groups .91 such that o( .91) = 0.
19. * Compute o( .91) and o 0 ( .91) as for many groups as you can in Coxeter
and Moser's (1957) list of all groups of order #.91 < 32.
20. Prove that the group obtained by adjoining to the ordinary (m, m, n)
triangle group the reflection in the perpendicular bisector to the side
opposite the wjn angle, has presentation
Now that the lists of all groups of symmetric genus 0 or 1 have been compiled
in Section 6.3, it is time to ask about genus rather than symmetric genus. Is it
possible that every group of genus 0 or 1 appears in the lists of Section 6.3?
Since y(..l11).::;; a(J11) for any finite group ..l11, every group in these lists has
genus at most 1. On the other hand, from the example at the end of Section
6.2, the inequality y( ..l11) .::;; a( J11) is strict for most abelian groups, so one
would not expect these lists to be complete. In this section, perhaps surpris-
ingly, it will be shown that if y(..l11).::;; 1, then y(..l11) = a(J11), with three
exceptions. Thus, with the addition of three extra presentations, the lists of
Section 6.3 classify all groups of genus 0 or 1. At the same time, it is shown
that Hurwitz's theorem applies to genus as well as to symmetric genus, that is,
#J11.::;; 168[y(..l11)- 1] if y(..l11) > 1. It follows that the number of groups of a
given genus g > 1 is finite.
The classification of groups of genus 0 or 1 begins with Maschke (1896),
who shows that if y( ..l11) = 0, then J1l is isomorphic to one of the groups !!En,
!!) n, J114 , ~, or J115 or to the direct product of f!E2 with any of these groups,
from which it follows that a(..l11) = y(..l11) when y(..l11) = 0. Baker (1931)
extended some of Maschke's work to the torus, but it was Proulx (1977, 1978)
who did all the massive work needed to classify groups of genus 1. Tucker
(1983, 1984a) clarified the connections of Proulx's results to group actions on
surface and space groups. He used those connections to prove that three of the
four group presentations not fitting into Proulx's list of general classes of
Euclidean space group quotients are truly exceptional. The imbedded-Cayley-
graph version of Hurwitz's theorem was proved first by Tucker (1980) and was
later refined by Tucker (1984b). This last paper includes also a proof of
Proulx's classification. Finally, this last paper opens the way to the determina-
tion by Tucker (1984c) of the only group of genus 2.
6.4.1. An Example
The basic idea behind the results of this section is fairly simple. Suppose that
the group J1l has a Cayley graph imbedded in a surface whose genus is small
compared with #..l11 (in particular, genus 0 or 1). Then this imbedding must
have many faces. If it has many faces, there must be some faces of small size.
A face of small size gives a short closed walk in the Cayley graph, which in
turn corresponds to a short relator in the group ..l11. Upon inspection, the
group presentations appearing in the lists in Section 6.3 all seem to have few
generators and short relators. If the imbedding forces enough short relators in
Groups of Small Genus 301
.J/1, then .J/1 will have one of the listed presentations. As example will illustrate
the details of this argument in a particular case.
y y
X V X y V X
y X
rd ) r-1
( - - r- d #V + rx(S) ~ L (r- i)f;
2 i=1
r#E - r#V + rx ( S) = r L /,
i=l
Therefore,
r-1
r#E- r#V + rx(S) ~ 2#£ + L (r- i)/;
i= 1
y z y z y
X X
X
<Dx
z y
X ®-;
z
X
X X X
y
X X
X X
y
<D ®
Figure 6.25. A subgraph of the Cayley graph C(d', X) and a partial imbedding of C(.JJI, X).
hand, if a pair of antipodal x-edges is removed from each x-cycle, the resulting
graph can be imbedded in the plane; in effect, the crossing of the z-diameters
can be undone with a twist at the expense of two x-edges crossing (see the
right of Figure 6.25). The four remaining x-edges can be reattached using
two handles (hitting the plane at spots marked 1 and 2 in Figure 6.25). Thus
y(C(.J<I, X))= 2.
Theorem 6.4.2 (Proulx, 1978; Tucker, 1980, 1984a). Let C(.!<l, X) --+ S be an
imbedding of an irredundant Cayley graph in a surfaceS. If #.!<1> -84x(S),
then .!<1 has one of the presentations listed in Theorems 6.3.1-6.3.3 or one of
these presentations:
To give some idea of the power of this theorem, we list three immediate
corollaries.
Groups of Small Genus 305
Corollary I. If y(d) > 1, then #d:5. 168(y(d)- 1). Thus there are only
finitely many groups of a given genus g > 1.
Corollary 2. If y(d) :5. 1, then either dis one of the three groups (a), (b), (c)
of Theorem 6.4.2 or else o(d) :5. 1. In particular, the lists of Section 6.3
together with (a),(b),(c) classify all groups of genus 0 or 1.
Corollary 3. If the group d has a Cayley graph imbeddable in the Klein bottle,
then it has a Cayley graph imbeddable in the torus.
This last corollary does not say that any Cayley graph imbeddable in the
Klein bottle is imbeddable in the torus. In fact, for redundant Cayley graphs
this is not true.
Example 6.4.2 Revisited. The two cycles with diameters formed by x-z edges
imbed in a pair of projective planes so that the x-cycles bound faces (view the
projective plane as a disk with antipodal points identified). If the two projective
planes are then joined by a tube to form a Klein bottle, the remaining y- and
w-edges can be added around the tube. Thus the Cayley graph C(d, X) imbeds
in the Klein bottle but not in the torus.
In general, Theorem 6.4.2 and its corollaries apply to the genus of a group,
not the genus of a Cayley graph. Part of the difficulty is that in the proof of
Theorem 6.4.2, the given generating set X occasionally must be changed to
obtain one of the presentations in the Section-6.3 lists; thus one obtains a
Cayley graph for d imbeddable in the torus, but it might not be the original
Cayley graph C( d, X). In fact, this difficulty can be overcome, but it involves
some intricate group theory that would take us far afield (Tucker, 1984b).
Thus, in Theorem 6.4.2, without extra hypotheses, it can be shown that the
306 The Genus of a Group
given Cayley graph C(.Jl/, X) imbeds in the torus or sphere whenever #.9/>
- 84x(S). It also follows that there are only finitely many irredundant Cayley
graphs of a given genus g > 1, that an irredundant Cayley graph imbeddable
in the Klein bottle is imbeddable in the torus, and that the lists of Section 6.3
together with presentations (a)-(c) of Theorem 6.4.2 classify all generating sets
and groups for irredundant Cayley graphs of genus 0 or 1.
The other difficulty in applying the results of this section to the genus of a
Cayley graph, rather than the genus of a group, is the restriction of irre-
dundancy. Example 6.4.2 shows that this is a serious problem. It is conjectured
that this example is, in effect, the only example that makes redundant Cayley
graphs different.
The proof of Theorem 6.4.2 occupies the remainder of this chapter. By the
second corollary to Theorem 6.4.1, only Cayley graphs of valence 3 or 4 need
be considered. The proof is broken into two main arguments: valence 4 and
valence 3. These are, in turn, broken into cases and subcases depending on the
nature of the given generating set X.
The proof of Theorem 6.4.2 involves much manipulation of generators and
relators. The study of groups given by generators and relators (usually called
combinatorial group theory) has many famous logical pitfalls, but here the
obvious intuitive approach is fine. Some agreement on terminology is, never-
theless, in order. A "word" in a generating set X for the group .9/ is a finite
sequence of elements of X and inverses of elements of X. The "length" of the
word is the number of terms in the sequence. A "relator" is a word the
product of whose terms is the identity; by abuse of language, we shall refer to
the product itself as a relator. Any conjugate or inverse of a relator is a relator,
as is any cyclic permutation. A relator is "reduced" if no cyclic permutation of
it can be shortened by using another relator to replace one subword by
another word. For example, if xyxy is a relator, then yzxzyx would not be a
reduced relator, since zxzyxy = zxzx- 1• A relator of length k is called a
k-relator.
Since (xy) 2 is the only 4-relator, there can be at most two faces of size 4
meeting at any vertex, and hence 4/4 ~ 2#Jii'. Since / 3 ~ #J/1'/3, it follows
that #J/1'> -42x(S). Thus if #J/1'> -42x(S), there must be a k-relator,
k ~ 6, not involving x, that is, y has order at most 6, as claimed.
and therefore #.9.1::::;, - 20x(S). On the other hand, if four faces of size 4 meet
at a vertex, then it is easily verified that, no matter what the rotation at that
vertex, the group .9.1 must have presentation (b 1 ), (h), or (1 2 ) of Theorem 6.3.3.
Case III. X= { x, y, z, w }, x 2 = y 2 = z 2 = w 2 = 1.
As in Case Ia, if #.9.1> - 20x(S) there must be a vertex at which four
faces of sizes 4 meet. By irredundancy, each reduced 4-relator involves two
generators. Thus there is some cyclic ordering of the generators such that
(xy) 2 ,(yz) 2 ,(zw) 2 , and (wx) 2 are all relators, which gives us presentation (j)
of Theorem 6.3.3.
contradicting the assumed order of x (in any case x 2 = x- 1 so that the relator
yxy = x 2 makes .5!1 the dihedral group ~3 ). A similar argument handles
yxyx 2 • We conclude that / 3 = f 4 = f 5 = 0. By Theorem 6.4.1, with r = 7,
#.Jil/2 + 7x ( s) 5;, / 6
has presentation (e) of Theorem 6.3.3 again. Finally, the relator (xyx- 1y) 3
gives us presentation (n) of Theorem 6.3.3.
We claim that if#..#> -84x(S), then xz must have order 6 or less, and
hence ..# has presentation (b), (c), or (d) of Theorem 6.3.2 or (q) of Theorem
6.3.3. The claim is equivalent to showing that there is a relator of length 13 or
less not involving y. Indeed, suppose that every relator of length 13 or less
involved y. By the analysis of previous cases, we can assume that 13 =Is=
1 7 =Is= 0 and that the only 4-relator is (xy) 2 and only 6-relator is (yz) 3•
Suppose we could show that 19 = 0, that every 10-relator uses at least four y 's,
and that every 11-relator uses at least three y's (by irredundancy, any relator
using y must use it twice). There are only #..# sides of edges labeled y
available for building faces. Counting occurrences of y, we would have
4#..# + 14x ( S) 5: 1014 + 816 + 4110 + 31n + 2112 + 113 5: 814 + 516 + #..#
X-
Y-..,-
1 2 3 4 5
Figure 6.26. A toroidal imbedding of group (a): (x, y: x 3 = y 3 = xyxy-lx-ly- 1 = 1). Vertices
with the same number are to be identified.
namely, (x, yxy) (see Exercises 4 and 5). The orders of the groups (a)-(c) are
then easily computed from the order of any one of them; the order of (c) is
computed in Exercise 6. None of these groups has symmetric genus 1;
Theorem 6.3.4 is used in Exercise 7 to show that the group (a) does not act on
the torus, and therefore (b) and (c) cannot either, since (a) is contained in both
(b) and (c). Also, as a perhaps curious coincidence, group (c) is the classical
group GL(2, 3) of 2 X 2 invertible matrices over the field with three elements,
which means that group (a) must be the subgroup SL(2, 3) of matrices with
determinant 1 (see Exercise 8). In summary, we have the following result.
Theorem 6.4.3 (Tucker, 1984a). The groups (a)-(c) of Theorem 6.4.2 are the
only groups .J/1 such that y( .J/1) = 1 and a( .J/1) > 1. In addition, group (a) is a
subgroup of both (b) and (c), and (c) is the group GL(2, 3).
11 12
Figure 6.27. A toroidal imbedding of group (b): (x, y: x 3 = y2 = (xy) 3 (x- 1y) 2 = 1). Vertices
with the same number are to be identified.
Theorem 6.4.5 (Tucker, 1984b). Suppose that y( .!#) > 1 and that
#..r#> 24(y(..r#)- 1). Then either the group ..r# is as described in Theorem
6.3.6, with the word "proper" omitted throughout, or else ..r# has a presentation
Groups of Small Genus 315
10
v-~13
,,.;-~ p~,
:( }>--<{ p~,
::}>--<{ }>--<{ ;::
::( )>--<{ p.-~!0
14'--_A k_/. 7
15 ~
17 \._
18."---
19
A2
-;N' 1
y. 3
5
of the form
(x, y: x 3 = y 2 = 1, w = 1, ... )
where w is a word in x, x-I, and y of length 14, 16, 18, 20, or 22. If
#..# > 48( y( ..#) - 1), then ..# is as described in Theorem 6.3.6, without the
word "proper".
Proof Omitted. 0
Conder (1987) has considered all possible words w. For w = (xy) 4 (x- 1y)\
Conder (1986) has shown that #..#> 24(y(..#)- 1) and ..# is not described
in Theorem 6.3.6. Thus the exceptional presentations of this type really do
provide examples of conflicts between y(..#) and o(..#), even when the order
of ..# is large compared with y( ..#). The role of the word "proper" in all of
this is not understood. It is known that improper quotients have Cayley-graph
imbeddings in the surface of "correct" Euler characteristic, but that surface
must be nonorientable. The genus of such groups is unknown; the alternating
groups .Jiln, n > 167, are an example of such groups, due to Conder (1980).
316 The Genus of a Group
lbeorem 6.4.6 (Tucker, 1984c). There is exactly one group of genus 2. It has
order 96, and it has the presentation
Proof That the given group has an imbedding in the surface of genus 2 (a
symmetric imbedding at that!) follows from the given presentation. The reader
can verify that the subgroup generated by y' = xy and x' = yz is the
exceptional group of genus 1 with presentation (c) in Theorem 6.4.2. In
particular, since this group does not act on the torus, neither does the given
group. Since the given group does not have order 24 or order 48, it is not
exceptional either. Thus the given group must have genus 2. An exhaustive
analysis of groups of order between 24 and 168, using the restrictions of
Theorem 6.4.5, shows that there are no other groups of genus 2. D
The group given in Theorem 6.4.6 probably does not look familiar. Coxeter
(1982) provides a full account of its origins in geometry and combinatorics; it
is, for example, the automorphism group of the generalized Petersen graph
G(8, 3) (see Frucht et al., 1971).
Finally, it is possible to adapt the methods used in the proof of Theorem
6.4.2 to give a partial answer to another fundamental question about the genus
of a group. Recall Babai's theorem that if !!4 is a subgroup of d, then
y(!!A).::;; y(d). One would expect that the same holds true for quotients: to
wit, that if 9. is a quotient of d, then y(9.).::;; y(d). This is clearly true for
symmetric and strong symmetric genus, by the natural action of the quotient
group on the quotient surface. Unfortunately, taking the quotient of a Cayley
graph can increase its genus. For instance, in Example 6.4.2 (due to Wormald),
the genus-2 Cayley graph for the presentation
Of course, the first presentation is redundant, and in fact the underlying group
~2 X ~2 n has genus 0, but this example does indicate that some subtle
questions about generating sets are involved, which is not the case for the
corresponding subgroup problem.
Groups of Small Genus 317
6.4.7. Exercises
1. Given graphs G and Hand edges d in G and e in H, define Gd • He to
be the graph obtained by deleting edges d and e and adding a pair of
edges joining the endpoints of d to the endpoints of e. Note that if G
and H are regular of valence n, then so is G d • He . Prove that if
y(G- d)= y(G) and y(H- e)= y(H), then y(Gd *He)= y(G) +
y(H).
2. Use Exercise 1 to construct infinitely many regular graphs of valence n,
3 5: n 5: 5, and genus g ~ 0 (Proulx, 1983).
3. Construct infinitely many regular graphs of valence 6 and genus g > 0
using Exercise 1 (Proulx, 1983). (Hint: First construct infinitely many
valence-6 triangulations of the torus.)
4. Prove that the subgroup (x, yxy) of group (b) of Theorem 6.4.2 is
isomorphic to group (a).
5. Prove that the subgroup (x, yxy) of group (c) of Theorem 6.4.2
is isomorphic to group (a) (let u = x, v = yx- 1y, and consider
uvuv - 1 u- 1v- 1 ).
6. In group (c) of Theorem 6.4.2, prove that (xy) 4y(xy) 4y = 1. Show then
that xy has order 8 and (xy) 4 is central. Conclude that group (c) has
order 48, by considering the quotient by the normal subgroup ((xy) 4 ).
7. Let .511 be the group (a) of Theorem 6.4.2 and suppose that a( .511) = 1.
Use Theorem 6.3.4 to prove that then .511 must have a normal subgroup
JV with quotient !Z3 and that both xy- 1 and y- 1x are in JV. Then
show that [xy-1, y- 1x] = 1 implies that x- 1y, has order 2 and that
therefore .511 has order 12, a contradiction.
8. Show that
x=(_~ -U
and
Mohammed Abu-Sbeih
-and T. D. Parsons (1983). Embeddings of bipartite graphs, J. Graph Theory 7, 325-334.
(210)
Jin Akiyama
- and M. Kano (1985). Factors and factorization of graphs-a survey, J. Graph Theory 9,
1-42. (This entire issue of the JOurnal is devoted to factorizations of graphs.) (79)
James W. Alexander
(1920). Note on Riemann spaces, Bull. Amer. Math. Soc. 26, 370-372. (180)
(1924). An example of a simply connected surface bounding a region which is not simply
connected, Proc. Nat. Acad. Sci. USA 10, 8-10. (100)
Seth R. Alpert
(1973). The genera of amalgamations of graphs, Trans. A mer. Math. Soc. 178, 1-39. (154)
(1975). Two-fold triple systems and graph imbeddings, J. Combin. Theory Ser. A 18, 101-107.
(105)
-and J. L. Gross (1976). Components of branched coverings of current graphs, J. Combin.
Theory Ser. B 20, 283-303. (87)
see also J. L. Gross
Dana Angluin
-and A. Gardiner (1981). Finite common coverings of pairs of regular graphs, J. Combin.
Theory Ser. B 30, 184-187. (80)
Kenneth Appel
- and W. Haken (1976). Every planar map is four-colorable, Bull. Amer. Math. Soc. 82,
771-712. (38, 215)
Dan Archdeacon
(1981). A Kuratowski theorem for the projective plane, J. Graph Theory 5, 243-246. (53)
(1984). Face colorings of embedded graphs, J. Graph Theory 8, 387-398. (46)
(1986). The orientable genus is nonadditive, J. Graph Theory 10, 385-401. (154)
-and P. Huneke (1985). On cubic graphs which are irreducible for non-orientable surfaces,
J. Combin. Theory Ser. B 39, 233-264. (53)
Louis Auslander
-with T. A. Brown and J. W. T. Youngs (1963). The imbedding of graphs in manifolds,
J. Math. and Mech. 12, 629-634. (140, 141)
319
320 References
Laszlo Babai
(1976). Embedding graphs in Cayley graphs, Probl. Comb. Theorie des Graphes, Proc. Conf.
Paris-Orsay (J.-C. Bermond et al., eds.), CNRS Paris, 13-15. (17)
(1977). Some applications of graph contractions, J. Graph Theory I, 125-130. (92ft')
- with W. Imrich and L. Lovasz (1972). Finite homeomorphism groups of the 2-sphere,
Colloq. Math. Janos Bolyai 8, Topics in Topology, Keszthely (Hungary). (287)
R. P. Baker
(1931). Cayley diagrams on the anchor ring, Amer. J. Math. 53, 645-669. (300)
Hyman Bass
see J.-P. Serre
Joseph Battle
- with F. Harary, Y. Kodama, and J. W. T. Youngs (1962). Additivity of the genus of a
graph, Bull. Amer. Math. Soc. 68, 565-568. (152)
Lowell W. Beineke
seeS. Stahl
K. N. Bhattacharya
(1943). A note on two-fold triple systems, Sankhya 6, 313-314. (105)
Ludwig Bieberbach
(1911). Uber die Bewegungsgruppen der euklidische Raume I, Math. Ann. 70, 297-336; II, 72
(1912), 400-412. (277)
Norman L. Biggs
-with E. K. Lloyd and R. J. Wilson (1976). Graph Theory, I736-I936, Clarendon Press,
Oxford. (35, 79)
R H Bing
(1954). Locally tame sets are tame, Ann. of Math. 59, 145-158. (18)
Garrett Birkhoff
see S. MacLane
Donald W. Blackett
(1967). Elementary Topology: A Combinatorial and Algebraic Approach, Academic Press,
Orlando. (110)
Bela Bollobas
- and P. Catlin (1980). Hadwiger's conjecture is true for almost every graph, European J.
Combin. I, 195-199. (40)
W. W. Boone
(1954). Certain simple unsolvable problems of group theory, Nederl. Akad. Wetensch. Indag.
Math. I6, 231-237 and 492-497; I7, 252-256 and 571-577; I9, 22-27 and 227-232. (14)
Andre Bouchet
(1982a). Constructions of covering triangulations with folds, J. Graph Theory 6, 57-74.
(175, 210)
(1982b). Constructing a covering triangulation by means of a nowhere-zero dual flow,
J. Combin. Theory Ser. B 32, 316-325. (210)
(1983). A construction of a covering map with faces of even lengths, Ann. Discrete Math. I7,
119-130. (210)
References 321
Matthew G. Brin
-and C. C. Squier (1986). On the genus of Z3 X Z3 X Z3 , Europ. J. Combin., to appear.
(262)
Edward M. Brown
(1969). The Hauptvermutung for 3-complexes, Trans. A mer. Math. Soc. 144, 173-196. (18, 98)
T. A. Brown
see L. Auslander
Robin P. Bryant
-and D. Singerman (1985). Foundations of the theory of maps on surfaces with boundary,
Quart. J. Math. Oxford Ser. (2) 36, 17-41. (281)
W. Burnside
(1911). Theory of Groups of Finite Order (1st ed., 1897). Cambridge Univ. Press, Cambridge.
(14, 265)
James W. Cannon
(1984). The combinatorial structure of cocompact discrete hyperbolic groups, Geom. Dedicata
16, 123-148. (14)
Paul Catlin
(1979). HaJOS graph-coloring conjecture: Variations and counterexamples, J. Combin. Theory
Ser. B 26, 268-274. (41)
see also B. Bollobas
Augustin-Louis Cauchy
(1813). Recherches sur les polyedres-premier memoire, J. de /'Ecole Poly. 9 (Cah. 16), 68-86.
[Extract in Biggs, Lloyd, and Wilson (1976).] (27)
Arthur Cayley
(1878). The theory of groups: Graphical representations, A mer. J. Math. I, 174-176. (10)
Marston D. E. Conder
(1980). Generators of the alternating and symmetric groups, J. London Math. Soc. (2) 22,
75-86. (305, 315)
(1984). Some results on quotients of triangle groups, Bull. Austral. Math. Soc. 30, 73-90. (297)
(1986). A note on Cayley graphs, J. Combin. Theory Ser. B 40, 362-368. (315)
(1987). Three-relator quotients of the modular group, Quart. J. Math. Oxford, to appear. (315)
Harold Scott MacDonald Coxeter
(1961). Introduction to Geometry, Wiley, New York. (279)
(1982). The group of genus two, Rend. Sem. Mat. Brescia 7, 219-248. (316)
- and W. 0. J. Moser (1957). Generators and Relations for Discrete Groups (4th ed., 1980).
Springer-Verlag, Berlin and New York. (14, 277, 291, 295)
Richard W. Decker
-with H. H. Glover and J.P. Huneke (1981). The genus of the 2-amalgamations of graphs,
J. Graph Theory 5, 95-102. (154)
-with H. H. Glover and J.P. Huneke (1985). Computing the genus of the 2-amalgamations
of graphs, Combinatorica 5, 271-282. (154)
Max Dehn
(1911). Uber unendliche diskontinuierliche Gruppen, Math. Ann. 72, 413-421. (10, 14)
322 References
G. Demoucron
- - with Y. Malgrange and R. Pertuiset (1964). Graphs planaires: Reconnaissance et
construction de representations planaires topologiques, Rev. Fram;aise Recherche Operation-
neUe 8, 33-47. (52)
Gabriel A. Dirac
--and S. Schuster (1954). A theorem of Kuratowsk.i, Nederl. Akad. Wetensch. lndag. Math.
16, 343-348. (264)
Richard A. Duke
(1966). The genus, regional number, and Betti number of a graph, Canad. J. Math 18,
817-822. (134)
(1971). How is a graph's Betti number related to its genus?, A mer. Math. Monthly 78, 386-388.
(136)
W. Dyck
(1882). Gruppentheoretische Studien, Math. Ann. 20, 1-45. (265)
Jack R. Edmonds
(1960). A combinatorial representation for polyhedral surfaces (abstract), Notices A mer. Math.
Soc. 7, 646. (113, 117, 192)
(1965). On the surface duality of linear graphs, J. Res. Nat. Bur. Standards Sect. 8, 121-123.
(119, 135)
Samuel Eilenberg
(1934). Sur les transformations periodique de Ia surface de sphere, Fund. Math. 22, 28-41.
(269)
Arnold Emch
(1929). Triple and multiple systems, their geometric configurations and groups, Trans. Amer.
Math. Soc. 31, 25-42. (105)
Paul Erdos
--and S. Fajtlowicz (1981). On the conjecture of Hajos, Combinatorica 1, 141-143. (41)
Leonhard Euler
(1736). Solution problematis ad geometriam situs pertinentis, Comment. Acad. Sci. /. Petro-
politanae 8, 128-140. (34)
(1750). Letter to C. Goldbach [excerpted in Biggs, Lloyd, and Wilson (1976)). (27)
Cloyd L. Ezell
(1979). Observations on the construction of covers using permutation voltage assignments,
Discrete Math. 28, 7-20. (88)
Siemion Fajtlowicz
seeP. Erdos
Istvan Fary
(1948). On straight line representation of planar graphs, Acta Sci. Math. (Szeged) 11, 229-233.
(48)
E. S. Federov
(1891). Zapiski Mineralogischeskogo lmperatorskogo S. Petersburgskogo Obstrechestva (2) 28,
345-390. (277)
Ion S. Filotti
--with G. Miller and J. Reif (1979). On determining the genus of a graph in 0( v0 <G>) steps,
Proc. 11th Annual ACM Symp. Theory of Computing, 27-37. (53)
References 323
Ralph H. Fox
(1957). Covering spaces with Singularities, Algebraic Geometry and Topology: A Symposium in
Honor of S. Lefschetz, Princeton University Press, Princeton, 243-257. (175)
Philip Franklin
(1934). A six color problem, J. Math. Phys. 16, 363-369. (137, 218)
Roberto Frucht
(1938). Herstellung von Graphen mit vorgegebener abstrakten Gruppe, Compositio Math. 6,
239-250. (70)
- with J. E. Graver and M. Watkins (1971). The groups of generalized Petersen graphs,
Proc. Camb. Philos. Soc. 70, 211-218. (59, 316)
Merrick L. Furst
-with J. L. Gross and L. McGeoch (1985a). Finding a maximum-genus graph imbedding,
J. Assoc. Comp. Mach., to appear. (133, 146)
-with J. L. Gross and R. Statman (1985b). Genus distributions for two classes of graphs,
J. Combin. Theory Ser. B, to appear. (147)
see also J. L. Gross
A. Gardiner
see D. Angluin
Brian L. Garman
(1979). Voltage graph embeddings and the associated block designs, J. Graph Theory 2,
181-187. (105)
Henry H. Glover
- with P. Huneke and C. Wang (1979). 103 graphs that are irreducible for the projective
plane, J. Combin. Theory Ser. B 27, 332-370. (53)
- and D. Sjerve (1985). Representing PSL 2 (p) on a Riemann surface of least genus,
Enseignement Math. 31, 305-325. (297)
see also R. W. Decker
Chris Godsil
-and W. lmrich (1985). Embedding graphs in Cayley graphs, Graphs and Combinatorics, to
appear. (17)
Jack E. Graver
see R. Frucht
Hubert Brian Griffiths
(1976). Surfaces (2nd ed., 1981). Cambridge University Press, Cambridge. (110)
Jonathan L. Gross
(1974). Voltage graphs, Discrete Math. 9, 239-246. (57, 165, 176)
(1977). Every connected regular graph of even degree is a Schreier coset graph, J. Combin.
Theory Ser. B 22, 227-232. (78)
(1978). An infinite family of octahedral crossing numbers, J. Graph Theory 2, 171-178.
(209, 214)
-and S. R. Alpert (1973). Branched coverings of graph imbeddings, Bull. A mer. Math. Soc.
79, 942-945. (194)
- and S. R. Alpert (1974). The topological theory of current graphs, J. Combin. Theory Ser.
B 17,218-233. (117,165,194, 196, 205)
324 References
B. Kerekjarto
(1919). Uber die endlichen topologische Gruppen der Kugelflache, Amsterdam Akad. Verst. 28,
555-556. (287)
(1921). Uber die periodische Transformationen der Kreischeibe und der Kugelflache, Math.
Ann. 80, 36-38. (269)
E. Ward Klein
see J. L. Gross
Felix Klein
(1911). Elementar-Mathematik vom hoheren Standpunkte aus, Vol. 2, Teubner, Leipzig. (26)
Donald Erwin Knuth
(1969). The Art of Computer Programming, Vol. 2, Addison-Wesley, Reading, Massachusetts.
(161)
Yukihiro Kodama
see J. Battle
Denes Konig
(1916). Uber graphen und ihre Anwendung auf Determinantheorie und Mengenlehre, Math.
Ann. 77, 453-465. (80)
S. Kundu
(1974). Bounds on the number of disjoint spanning trees, J. Combin. Theory Ser. B 17,
199-203. (145)
Kasimierz Kuratowski
(1930). Sur le probleme des courbes gauches en topologie, Fund. Math. 15, 271-283. (28)
Henry M. Levinson
(1970). On the genera of graphs of group presentations, Ann. New York Acad. Sci. 175,
277-284. (249, 264)
-and B. Maskit (1975). Special embeddings of Cayley diagrams, J. Combin. Theory Ser. B
18, 12-17. (265)
S.-A.-J. Lhuilier
(1811). Memoire sur Ia polyedrometrie, Ann. de Mathematiques 3, 169-189. [Extract in Biggs,
Lloyd, and Wilson (1976).) (27)
J. B. Listing
(1861). Der Census raumlicher Complexe oder Verallgemeinerung des Euler'schen Satzes von
den Polyedern, Abh. K. Ges. Wiss. Gottingen Math. CliO, 97-182. (27)
E. Keith Lloyd
see N. L. Biggs
Samuel J. Lomonaco
see J. L. Gross
Laszlo Lovasz
see L. Babai
Roger C. Lyndon
(1985). Groups and Geometry (London Math. Soc. Lecture Note Series, No. 101), London
Math. Soc., Cambridge. (277, 279)
- - and P. E. Schupp (1977). Combinatorial Group Theory, Springer-Verlag, Berlin and New
York. (14)
References 327
C. Maclachlin
(1965). Abelian groups of automorphisms of compact Riemann surfaces, Proc. London Math.
Soc. 15, 699-712. (265)
Saunders MacLane
(1937). A structural characterization of planar combinatorial graphs, Duke Math. J. 3,
460-472. (54)
-and G. Birkhoff (1967), Algebra, Macmillan, New York. (254)
Wilhelm Magnus
(1974). Non-Euclidean Tessalations and Their Groups, Academic Press, Orlando. (279)
--with A. Karrass and D. Solitar (1966). Combinatorial Group Theory, Wiley-Interscience,
New York. (14)
Y. Malgrange
see G. Demoucron
H. Maschke
(1896). The representation of finite groups, A mer. J. Math. 18, 156-194. (287, 290, 300, 314)
Bernard Maskit
see H. Levinson
Lyle McGeoch
see M. L. Furst
Martin Milgram
--and P. Ungar (1977). Bounds for the genus of graphs with given Betti number, J. Combin.
Theory Ser. B 23, 227-233. (136)
Gary Miller
see I. S. Filotti
John Milnor
(1961). Two complexes which are homeomorphic but combinatorially distinct, Ann. of Math.
74, 575-590. (18, 98)
(1965a). Lectures on the h-Cobordism Theorem (Princeton Mathematical Notes), Princeton
University Press, Princeton. (109)
(1965b). Topology from the Differentiable Viewpoint, University Press of Virginia, Charlottes-
ville. (185)
August F. Mobius
(1865). Uber die Bestimmung des Inhaltes eines Polyeders, Ber. K. Sachs. Ges. Wiss. Leipzig
Math.-Phys. Cl17. (25)
Bojan Mohar
- - with T. Pisanski, M. Skoviera, and A. T. White (1985). The cartesian product of three
triangles can be imbedded into a surface of genus 7, Discrete Math. 56, 87-89. (262)
Edwin E. Moise
(1952). Affine structures in 3-manifolds, V, Ann. of Math. 56, 96-114. (18)
J. Montesinos
(1976). Three-manifolds as 3-fold branched covers of S 3 , Quart. J. Math. Oxford 27, 85-94.
(180)
William 0. J. Moser
see H. S. M. Coxeter
328 References
James R. Munkres
(1975). Topology: A First Course, Prentice-Hall, Englewood Cliffs, New Jersey. (95)
Seiya Negami
(1983). Uniqueness and faithfulness of embedding of toroidal graphs, Discrete Math. 44,
161-180. (147)
(1985). Unique and faithful embeddings of projective-planar graphs, J. Graph Theory 9,
235-243. (147)
M. H. A. Newman
(1954). Elements of the Topology of Plane Sets of Points, Cambridge Univ. Press, Cambridge.
(100)
P. Niggli
(1924). Die Flachensymmetrien homogener Diskontinuen, Z. Kristallogr. Mineralog. Petrog.
Abt. A 60, 283-298. (278)
Edward A. Nordhaus
- with B. M. Stewart and A. T. White (1971). On the maximum genus of a graph,
J. Combin. Theory Ser. B 11, 258-267. (132)
-with R. D. Ringeisen, B. M. Stewart, and A. T. White (1972). A Kuratowski-type theorem
for the maximum genus of a graph, J. Combin. Theory Ser. B 12,260-267. (132, 148)
P. S. Novikov
(1955). On the algorithmic unsolvability of the word problem in groups (in Russian), Trudy
Mat. lnst. Steklov 44, lzdat. Akad. Nauk. SSSR, Moscow. (14)
Christos D. Papakyriakopoulos
(1943). A new proof of the invariance of the homology groups of a complex, Bull. Soc. Math.
Grece 22, 1-154. (18, 98)
Torrence D. Parsons
-with T. Pisanski and B. Jackson (1980). Dual imbeddings and wrapped quasi-coverings of
graphs, Discrete Math. 31, 43-52. (207, 210)
see also M. Abu-Sbeih, J. L. Gross, and B. W. Jackson
David Pengelley
(1975). Self-dual orientable embeddings of K., J. Combin. Theory Ser. B 18, 46-52. (209)
R. Pertuisset
see G. Demoucron
Julius Petersen
(1891). Die Theorie der regularen Graphen, Acta Math. 15, 193-220. (35, 36)
Tomaz Pisanski
(1980). Genus of cartesian products of regular bipartite graphs, J. Graph Theory 4, 31-42.
(155, 159)
(1982). Nonorientable genus of cartesian products of regular graphs, J. Graph Theory 6,
31-42. (156)
see also J. L. Gross, B. W. Jackson, B. Mohar, and T. D. Parsons
Georg Polya
(1924). Uber die Analogie der Kristallsymmetrie in der Ebene, Z.Kristallogr. Mineralog.
Petrog. Abt. A 60, 278-282. (278)
References 329
Viera Krnanova Proulx
(1978). Classification of the toroidal groups, J. Graph Theory 2, 269-273. [Extraction of results
from Ph.D. thesis (1977), Columbia University.] (300, 304)
(1981). On the genus of symmetric groups, Trans. A mer. Math. Soc. 266, 531-538. (318)
(1983). On the genus of five- and six-regular graphs, J. Graph Theory 7, 231-234. (317)
Tibor Rado
(1925). Uber den Begriff der Riemannschen Flache, Acta Litt. Sci. Szeged 2, 101-121. (95)
John Reif
see I. S. Filotti
Richard D. Ringeisen
(1972). Determining all compact orientable 2-manifolds upon which Km. • has 2-cell imbed-
dings, J. Combin. Theory Ser. B 12, 101-104. (148)
(1978). Survey of results on the maximum genus of a graph, J. Graph Theory 3, 1-13. (142)
see also E. A. Nordhaus
Gerhard Ringel
(1955). Uber drei kombinatorische Probleme am n-dimensionalen Wurfel und Wurfelgitter,
Abh. Math. Sem. Univ. Hamburg 20, 10-19. (30, 159)
(1959). Farbungsprobleme auf Flachen und Graphen, Springer-Verlag, Berlin and New York.
(30, 236ff)
(1961). Uber das Problem der Nachbargebiete auf orientierbaren Flachen, Abh. Math. Sem.
Univ. Hamburg 25, 105-127. (192, 230, 234)
(1965). Das Geschlecht des vollstandigen paaren Graphen, Abh. Math. Sem. Univ. Hamburg
28, 139-150. (30)
(1974). Map Color Theorem, Springer-Verlag, Berlin and New York. (113, 216, 240, 245)
-and J. W. T. Youngs (1968). Solution of the Heawood map-coloring problem, Proc. Nat.
A cad. Sci. USA 60, 438-445. (38, 215)
Neil Robertson
-and P. D. Seymour (1983). Graph minors, I: excluding a forest, J. Combin. Theory Ser. B
35, 39-61. (53)
-and P. D. Seymour (1984). Graph minors, III: planar tree-width, J. Combin. Theory Ser.
B 36, 49·-64. (53)
- and P. D. Seymour (1986a). Graph minors, V: excluding a planar graph, J. Combin.
Theory Ser. B 41,92-114. (53)
-and P. D. Seymour (1986b). Graph minors, VI: disjoint paths across a disc, J. Combin.
Theory Ser. B 41, 115-138. (53)
- and P. D. Seymour (1987). Graph minors, II: algorithmic aspects of tree width,
J. Algorithms, to appear. (53)
Ronald H. Rosen
see J. L. Gross and F. Harary
Gert Sabidussi
(1958). On a class of fixed-point free graphs, Proc. Amer. Math. Soc. 9, 800-804. (69, 272)
Doris Schattschneider
(1978). The plane symmetry groups: their recognition and notation. A mer. Math. Monthly 85,
439-450. (277)
330 References
W. Scherrer
(1929). Zur Theorie der endlichen Gruppen topologischer Abbildungen von geschlossenen
Flachen in sich, Comment. Math. Helv. 1, 69-119. (287)
Otto Schreier
(1927). Die Untergruppen der freien Gruppen, Abh. Math. Sem. Univ. Hamburg 5, 161-183.
(10)
Paul E. Schupp
see R. C. Lyndon
Seymour Schuster
see G. A. Dirac
Jean-Pierre Serre
--with H. Bass (1977). Groups discretes: arbres, amalgams et SL2 , Cours redige avec Ia
collaboration of H. Bass, Asterisque 46. (14)
Paul D. Seymour
seeN. Robertson
David Singerman
see R. P. Bryant and G. A. Jones
Cho Wei Sit
(1976). Quotients of complete multipartite graphs, Pacific J. Math. 63, 531-538. (236)
(1980). Ph.D. thesis, Columbia University. (250, 318)
Denis Sjerve
see H. H. Glover
Martin Skoviera
see B. Mohar
Stephen Smale
(1961). Generalized Poincare's conjecture in dimensions greater than four, Ann. of Math. 74,
391-406. (109)
Donald Solitar
see W. Magnus
Craig C. Squier
seeM. G. Brin
Saul Stahl
(1978). Generalized embedding schemes, J. Graph Theory 2, 41-52. (113, 136)
(1979). Self-dual embeddings of Cayley graphs, J. Combin. Theory Ser. B 27, 92-107. (209)
(1980). Permutation-partition pairs: a combinatorial generalization of graph embeddings,
Trans. Amer. Math. Soc. 259, 129-145. (154)
(1981). Permutation-partition pairs II. Bounds on the genus of amalgamations of graphs,
Trans. A mer. Math. Soc. 271, 175-182.
- - and L. W. Beineke (1977). Blocks and the nonorientable genus of graphs, J. Graph
Theory 1, 75-78. (155)
--and A. T. White (1976). Genus embeddings for some complete tripartite graphs, Discrete
Math. 14, 279-296. (87, 170)
References 331
John Stallings
(1971). Group Theory and Three-Dimensional Manifolds, Yale University Press, New Haven.
(14)
Richard Statman
see M. L. Furst
Bonnie M. Stewart
see E. A. N ordhaus
Robert E. Trujan
see J. E. Hopcroft
Charles M. Terry
-with L. R. Welch and J. W. T. Youngs (1967). The genus of K 12., J. Combin. Theory 2,
43-60. (235)
-with L. R. Welch and J. W. T. Youngs (1970). Solution of the Heawood map-coloring
problem-case 4, J. Combin. Theory 8, 170-174. (233)
Carsten Thomassen
(1980). Planarity and duality of finite and infinite graphs, J. Combin. Theory Ser. B 29,
244-271. (42)
H. Tietze
(1910). Einige Bemerkerungen uber das Problem des Kartenfarbens auf einseitigen Flachen,
Jahresber. Deutsch. Math.-Verein. 19, 155-159. (217)
Marvin D. Tretkoff
(1975). Covering space proofs in combinatorial group theory, Comm. Algebra 3, 429-457. (14)
(1980). A topological approach to the theory of groups acting on trees, J. Pure Appl. Algebra
16, 323-333. (14)
Albert W. Tucker
(1936). Branched and folded coverings, Bull. Amer. Math. Soc. 42, 859-862. (175)
Thomas W. Tucker
(1980). The number of groups of a given genus, Trans. Amer. Math. Soc. 258, 167-179.
(300, 304)
(1983). Finite groups acting on surfaces and the genus of a group, J. Combin. Theory Ser. B 34,
82-98. (265, 273, 300)
(1984a) On Proulx's four exceptional toroidal groups, J. Graph Theory 8, 29-33. (300, 304,
313)
(1984b). A refined Hurwitz theorem for imbeddings of irredundant Cayley graphs, J. Combin.
Theory Ser. B 36, 244-268. (300, 305, 314)
(1984c). There is one group of genus two, J. Combin. Theory B 36, 269-275. (300, 316)
see also J. L. Gross
William Tutte
(1947). The factorizations of linear graphs, J. London Math. Soc. 22, 107-111. (79)
(1949). On the imbedding of linear graphs in surfaces, Proc. London Math. Soc. (2) 51,
474-483. (46)
(1960). Convex representation of graphs, Proc. London Math. Soc. 10, 304-320. (48)
Peter Ungar
see M. Milgram
332 References
V. G. Vizing
(1964). On an estimate of the chromatic class of a p-graph (in Russian), Diskret. Analiz 3,
25-30. (160)
Chin San Wang
see H. H. Glover
Mark E. Watkins
see R. Frucht
Lloyd R. Welch
see C. M. Terry
Hermann Weyl
(1952). Symmetry, Princeton University Press, Princeton. (277)
Arthur T. White
(1970). The genus of repeated cartesian products of bipartite graphs, Trans. Amer. Math. Soc.
151, 393-404. (15511')
(1972). On the genus of a group, Trans. Amer. Math. Soc. 173, 203-214. (249, 256)
(1974). Orientable imbeddings of Cayley graphs, Duke. Math. J. 41, 353-371. (207)
(1978). Block designs and graph imbeddings, J. Combin. Theory Ser. B 25, 166-183. (105)
(1973). Graphs, Groups and Surfaces (rev. ed., 1984), North-Holland, Amsterdam. (93)
see also M. Jungerman, B. Mohar, E. A. Nordhaus, and S. Stahl
Hassler Whitney
(1932). Non-separable and planar graphs, Trans. Amer. Math. Soc. 34, 339-362. (54)
(1933). A set of topological invariants for graphs, Amer. J. Math. 55, 231-235. (49)
Robin J. Wilson
see N. L. Biggs
Nguyen Huy Xuong
(1979). How to determine the maximum genus of a graph, J. Combin. Theory Ser. B 26,
217-225. (133, 14411')
J. W. T. Youngs
(1963). Minimal imbeddings and the genus of a graph, J. Math. and Mech. 12, 303-315. (117)
(1967). The Heawood map coloring conJecture, Graph Theory and Theoretical Physics
(F. Harary, ed.), Academic Press, Orlando, 313-354. (194)
(1968a). Remarks on the Heawood conjecture (nonorientable case), Bull. Amer. Math. Soc. 74,
347-353. (199, 237)
(1968b). The nonorientable genus of Kn, Bull. Amer. Math. Soc. 74, 354-358. (237)
see also L. Auslander, G. Ringel, and C. M. Terry
Bibliography
Pointers are given to further results on topics discussed in the text and also to
several major research areas mentioned scarcely or not at all. A few useful
monographs are cited. We have favored papers with topological content and
tended to exclude papers not yet published.
Michael Albertson
-and J. P. Hutchinson (1977). The independence ratio and genus of a graph, Trans. A mer.
Math. Soc. 226, 161-173.
-and J. P. Hutchinson (1978). On the independence ratio of a graph, J. Graph Theory 2,
1-8.
-and J.P. Hutchinson (1980). On six-chromatic toroidal graphs, Proc. London Math. Soc.
( 3) 41, 533-556.
- and J. P. Hutchinson (1980). Hadwiger's conJecture for graphs on the Klein bottle,
Discrete Math. 29, 1-11.
V. B. Alekseev
-.and V. S. Gonchakov (1976). Thickness of arbitrary complete graphs, Mat. Sb. 101,
212-230.
Amos Altschuler
(1973). Construction and enumeration of regular maps on the torus, Discrete Math. 4,
201-217.
I. Anderson
-and A. T. White (1978). Current graphs and bi-embeddings, J. Graph Theory 2, 231-239.
Dan Archdeacon
(1986). The non-orientable genus is additive, J. Graph Theory 10, 363-383.
Konhei Asano
(1986). The crossing number of K1• 3• • and K 2• 3. "' J. Graph Theory 10, 1-8.
M. Behzad
- with G. Chartrand and L. Lesniak-Foster. Graphs and Digraphs, Prindle, Weber, and
Schmidt, Boston.
Lowell W. Beineke
- and F. Harary (1969). The thickness of the complete graph, Canad. J. Math. 21,
850-859.
-with F. Harary and J. W. Moon (1964). On the thickness of the complete bipartite graph,
Proc. Cambridge Philos. Soc. 60, 1-5.
333
334 Bibliography
David S. Cochran
-and R. H. Crowell (1970). H2 (G') for tamely embedded graphs, Quart. J. Math. Oxford
Ser. (2) 21, 25-27.
H. Coldeway
see H. Zieschang
R. Connelly
seeM. Brown
John H. Conway
- and C. MeA. Gordon (1983). Knots and links in spatial graphs, J. Graph Theory 7,
445-453.
Richard H. Crowell
see D. S. Cochran
A. K. Dewdney
(1972). The chromatic number of a class of pseudo-manifolds, Manuscripta Math. 6, 311-320.
Gabriel A. Dirac
(1953). The coloring of maps, J. London Math. Soc. 28, 476-480.
(1957). Short proof of a map colour theorem, Canad. J. Math. 9, 225-226.
Dragomir Z. Djokovic
(1974). Automorphisms of graphs and coverings, J. Combin. Theory Ser. B 16, 243-247.
W. Dorfler
(1978). Double covers of hypergraphs and their properties, Ars Combin. 6, 293-313.
Allan L. Edmonds
- with J. H. Ewing and R. S. Kulkarni (1982). Regular tesselations of surfaces and
( P, Q, 2) triangle groups, Ann. of Math. 116, 113-132.
John H. Ewing
see A. L. Edmonds
M. Farzan
-and D. A. Waller (1977). Antipodal embeddings of graphs, J. London Math. Soc. (2) 15,
377-383.
Massimo Ferri
(1979). Crystallisation of 2-fold branched coverings of S 3 , Proc. Amer. Math. Soc. 73,
271-276.
-with C. Gagliardi and L. Graselli (1986). A graph-theoretical representation of PL-mani-
folds-a survey on crystallizations, A equationes Math. 31, 121-141.
Steve Fisk
(1977). Geometric coloring theory, Adv. ,in Math. 24, 298-340.
Carlo Gagliardi
(1981). Extending the concept of genus to dimension n, Proc. A mer. Math. Soc 81,473-481.
see also M. Ferri
V. S. Gonchakov
see V. B. Alekseev
Cameron MeA. Gordon
see J. H. Conway
336 Bibliography
L. Grasselli
seeM. Ferri
Jack Graver
- - and M. Watkins (1977). Combinatorics with Emphasis on the Theory of Graphs, Springer-
Verlag, New York.
Jonathan L. Gross
(1975). The genus of nearly complete graphs-case 6, Aequationes Math. 13,243-249.
Branko Grunbaum
(1970). Polytopes, graphs, and complexes, Bull. Amer. Math. Soc. 76, 1131-1201.
Richard K. Guy
(1969). The decline and fall of Zarankiewicz's theorem, Proof Techniques in Graph Theory
(F. Harary, ed.), Academic Press, Orlando.
(1972). Crossing number of graphs, Graph Theory and Applications, Springer-Verlag, New
York, 111-124.
- - and L. W. Beineke (1966). The coarseness of the complete graph, Canad. J. Math. 20,
888-894.
Frank Harary
- - and W. Tutte (1965). A dual form of Kuratowski's theorem, Canad. Math. Bull. 8,
17-20.
see also L. W. Beineke
Nora Hartsfield
(1986). The toroidal splitting number of the complete graph Kn, Discrete Math. 62, 35-47.
Peter Hoffman
- - and B. Richter (1984). Embedding graphs in surfaces, J. Combin. Theory Ser. B 36,
65-84.
Joan P. Hutchinson
--and George F. McNulty (1983). Connected graphs of genus g with complementary orbits,
Discrete Math. 45, 255-275.
see also M. 0. Albertson
T. Ito
(1981). On a graph of O'Keefe and Wong, J. Graph Theory 5, 87-94.
Bradley W. Jackson
- and G. Ringel (1984). Coloring island maps, Bull. Austral. Math. Soc. 29, 151-165.
- - and G. Ringel (1984). The splitting number of complete bipartite graphs, Arch. Math.
42,178-184.
--and G. Ringel (1984). Heawood's empire problem on the plane, J. Reine Angew. Math.
347, 148-153.
- - and G. Ringel (1985). Heawood's empire problem, J. Combin. Theory Ser. B 38,
168-178.
Lynne D. James
- - and G. A. Jones (1985). Regular orientable imbeddings of complete graphs, J. Combin.
Theory Ser. B 39, 353-367.
Gareth A. Jones
see L. D. James
Bibliography 337
Paul C. Kainen
- and A. T. White (1978). On stable crossing numbers, J. Graph Theory 2, 181-187.
see also F. Bernhart
Daniel J. Kleitman
(1970). The crossing number of K 5•• , J. Combin. Theory 9, 315-323.
Hudson V. Kronk
(1969). An analogue to the Heawood map-coloring problem, J. London Math. Soc. (2) I,
550-552.
Ravi S. Kulkarni
see A. L. Edmonds
Gerard Lallament
--and D. Perrin (1981). A graph covering construction of all finite complete biprefix codes,
Discrete Math. 36, 261-272.
Frank Thomson Leighton
(1982). Finite common coverings, J. Combin. Theory Ser. B 33, 231-238.
Linda Lesniak-Foster
see M. Behzad
Sostenes Lins
-and A. Mandel (1985). Graph-encoded 3-manifolds, Discrete Math. 57, 261-284.
Bernard Liouville
(1978). Sur le genre d'une somme cartesienne de graphes, Problemes combinatoires et theorie
des • * h • graphes, Colloques Internal. CNRS 260, Paris, 275-277.
Richard J. Lipton
-and R. E. Trujan (1979). A separator theorem for planar graphs, SIAM J. Appl. Math.
36, 177-189.
--and R. E. TaiJan (1980). Applications of a planar separator theorem, SIAM J. Comput.
9, 615-627.
Heidi Mahnke
(1972). The necessity of non-abelian groups in the case 0 of the Heawood map-coloring
problem, J. Combin. Theory B I3, 263-265.
Joseph Malkevitch
(1970). Properties of planar graphs with uniform vertex and face structure, Mem. Amer.
Math. Soc. 99.
Arnaldo Mandel
seeS. Lins
George F. McNulty
see J. P. Hutchinson
L. S. Melnikov
see 0. V. Borodin
Bojan Mohar
(1986). A common cover of graphs and 2-cell embeddings, J. Combin. Theory Ser. B 40,
94-106.
John W. Moon
see L. W. Beineke
338 Bibliography
Bruce P. Mull
- - and R. G. Rieper and A. T. White (1987). Enumerating 2-cell imbeddings of complete
graphs, Proc. A mer. Math. Soc., to appear.
Ronald C. Mullin
(1965). On counting rooted triangular maps, Canad. J. Math. 17, 373-382.
L. Nebesky
(1981). Every connected, locally connected graph is upper embeddable, J. Graph Theory 5,
197-199.
Seiya Negarni
(1987). Enumeration of proJective-planar embeddings of graphs, Discrete Math., to appear.
Lee P. Neuwirth
(1966). Imbedding in low dimension, lllinois J. Math. 10, 470-478.
Torrence D. Parsons
(1980). Circulant graph imbeddings, J. Combin. Theory SerB 29, 310-320.
Dominique Perrin
see G. Lallament
Giustina Pica
- - with T. Pisanski and A. G. S. Ventre. The genera of amalgamations of cube graphs,
Glasnik Mat. 19, 21-26.
Tomaz Pisanski
- with J. Shawe-Taylor and J. Vrabec (1983). Edge-colorability of graph bundles, J.
Combin. Theory Ser. B 35, 12-19.
see also G. Pica
L. Bruce Richmond
see E. A. Bender
Bruce Richter
- and H. Shank (1984). The cycle space of an embedded graph, J. Graph Theory 8,
365-369.
see also P. Hoffman
Robert G. Rieper
see B. P. Mull
Richard D. Ringeisen
see L. W. Beineke
Gerhard Ringel
see B. Jackson
Thomas L. Saaty
(1967). Two theorems on the minimum number of intersections for complete graphs, J.
Combin. Theory 2, 571-584.
Edward F. Schmeichel
- - and G. S. Bloom (1979). Connectivity, genus, and the number of components in
vertex-deleted subgraphs, J. Combin. Theory Ser. B 27, 198-201.
Herbert Shank
see B. Richter
Bibliography 339
John Shawe-Taylor
see T. Pisanski
Jozef Siran
- - and M. Skoviera (1985). Quotients of connected regular graphs of even degree, J.
Combin. Theory Ser. B 38, 214-225.
Martin Skoviera
(1986). A contribution to the theory of voltage graphs, Discrete Math. 61, 281-292.
see also J. Siran
Z. Skupien
(1966). Locally Hamiltonian and planar graphs, Fund. Math. 58, 193-200.
Saul Stahl
(1983). The average genus of classes of graph embeddings, Congr. Number. 40, 375-388.
Richard P. Stanley
(1981). Factorization of permutations into n-cycles, Discrete Math. 37, 255-262.
H. Joseph Straight
(1979). Cochromatic number and the genus of a graph, J. Graph Theory 3, 43-51.
Robert E. T3.1Jan
see R. J. Lipton
William T. Tutte
(1973). What is a map? New Directions in the Theory of Graphs (F. Harary, ed.), Academic
Press, Orlando, 309-325.
(1986). From topology to algebra, J. Graph Theory 10, 331-338.
see also F. Harary
Brian Ummel
(1978). The product of nonplanar complexes does not imbed in 4-space, Trans. A mer. Math.
Soc. 242, 319-328.
Aldo G. S. Ventre
see G. Pica
Andrew Vince
(1983). Combinatorial maps, J. Combin. Theory Ser. B 34, 1-21.
(1983). Regular combinatorial maps, J. Combin. Theory Ser. B 35, 256-277.
E. Vogt
see H. Zeischang
Joze Vrabec
see T. Pisanski
Derek A. Waller
(1976). Double covers of graphs, Bull. Austral. Math. Soc. 14, 233-248.
see also F. W. Clarke and M. Farzan
Mark Watkins
see J. C. Graver
Arthur T. White
(1983). Ringing the changes, Math. Proc. Camb. Philos. Soc. 94, 203-215.
see also I. Anderson and B. G. Mull
340 Bibliography
Nick C. Wormald
(1981). Counting unrooted planar maps, Discrete Math. 36, 205-225.
Joseph Zaks
(1974). The maximum genus of cartesian products of graphs, Canad. J. Math. 26, 1025-1035.
Heiner Zieschang
--with E. Vogt and H. Coldeway (1980). Surfaces and Planar Discontinuous Groups (Lecture
Notes in Mathematics, Vol. 835), Springer-Verlag, New York.
Table of Notations
GRAPHS
341
342 Table of Notations
GROUPS
GENERAL NOTATION
• )I ....4;E--..
• a type-1 edge in a current graph, 199
347
348 Subject Index
Coi1,226 Cycle:
Coloring: rank, 135
of graph, 37 space, 54
of map, 37 structure of permutation, 83
Commutator of group elements, 29, 275
Complete bipartite graph, 14 Deficiency:
genus of, 210ff of branch point, 175
maximum genus of, 148 of graph, 143
Complete graph, 10 of spanning tree, 143
crosscap number of, 223tT Dihedral group, 285
genus of, 223ff presentation for, 285
maximum genus of, 146 Disk sum of two surfaces, 131
Complex: Distance between vertices, 36
cell, 99 Dodecahedron, 19
planar 2-complex, 117 Double-cover conjecture, 46
simplicial, 96 Dual:
abstract, 97 graph, 31,116
Composition of graph, 210 imbedding, 31, 116
triangular imbedding for, 210 plus-direction of edge, 202
Connected, 8 voltage graph, 204tT
component, 18 Duality, 31, 116
k-connected, 40 current- voltage, 202tT
k-edge-connected, 145
Co- spectral graphs, 7 Edge:
Covering, 23, 72 coloring of graph, 156
branched, 182tT saturated color, 156
folded, 175 contraction, 39
irregular, 72 tT -deletion surgery, 124
projection, 23, 72 direction (plus, minus), 4
regular, 23, 75, 187 initial, terminal vertices of, 4
space, 23, 75 loop, 4
wrapped, 207 proper, 4
Crosscap, 26 sliding, 125
number of graph, 13 3 tT -twisting surgery, 134
relation to genus, 137tT Euclidean:
range, 13 3 tT space group, 276tT
Cube ( n- cube), 19 triangle group, 280
crosscap number of, 159 Euler:
genus of, 159 characteristic:
Current assignment, 194 of cellular imbedding (relative), 121 tT
Current graph (ordinary), 194ff of surface, 2 7, 130
derived graph of, 194 equation, 27
derived imbedding of, 194 formula, 27
dual, 206 Eulerian:
derived graph of, 206 circuit, 34
derived imbedding of, 206 graph, 34
natural projection for, 206
nonorientable, 198tT Face, 100
edge types of, 199 Tracing Algorithm, 115
reversed comer of, 199 Factor ( n- factor), 35
orientable, 194 Factorable (n-factorable), 35
Current group, 194 Five-color theorem, 219
Cutpoint, 40 Folded covering, 175
Cycle (n-cycle), 8 Four-Color theorem, 38, 215
Subject Index 349
Frucht's theorem, 70 Gustin's current graphs, 193
Fundamental:
group of graph, 92 Hadwiger conjecture, 40
semigroup of graph, 92 Hajos conjecture, 41
Hamiltonian cycle, 35
Genus: Hamiltonian graph, 35
of graph, 29, 132 Handle decomposition, 109
average, 147 Hauptvermutung, 98
distribution, 14 7 Heawood:
maximum, 132 inequality, 215 ff
minimum, 132 number, 215fT
nonorientable ( crosscap number), 29, 133 problem:
range, 132 orientable, 215
of group, 93, 249fT regular cases, 224fT
abelian, 249fT irregular cases, 224fT
infinite, 249 nonorientable, 215
strongly symmetric, 264 regular cases, 224fT
symmetric, 264 irregular cases, 224fT
of surface, 29 Heffier, Lothar, 191
nonorientable ( crosscap number), 29 Hill-climb, 149
Girth, 29 Hurwitz's theorem, 296
Glide, 276 refined, 298
n-Gon, 27 Hypergraph, 97
Graph, l
adjacency matrix, 4 Icosahedron, 41
amalgamation, 20 Imbedding:
automorphism, 6 adjacent, 13 3
group, 6 cellular (2-cell), 100
bipartite, 14 distribution, 13 2
cartesian product, 19 crosscap range, 132
edge-complement, 19 Euler characteristic range, 148
edge set, l genus range, 132
homeomorphic, 18 dual, 31, 116
incidence matrix, 3 equivalent, lO l
isomorphism, 5 face (region) of, 26, 100
map, 5 boundary walk of, l 0 l
regular ( n- regular), 13 side of, 26, lO l
simplicial (simple), l size of, 26, 101
subdivision, 18 quadrilateral, 27
subgraph, 16 triangular, 2 7
induced, 17 type (0, l) of edge, II 0
spanning, 16 type (0, l) of walk, 110
suspension (join), 20 weakly equivalent (congruent), l 02
topological representation, l 7 Interpolation theorem:
vertex set, l nonorientable, 135
Graphical regular representation for group, orientable, 134
72
Group: Jordan curve theorem, 100
abelian, 254
Cayley graph for, lOfT Kempe chain, 220
crystallographic, 2 77 Kirchoff current law ( KCL), 196
current, l 94 ff Kirchoff voltage law (KVL), 166
of small genus, 283fT Klein bottle, 26, 103
voltage, 57fT K-metacyclic group, 318
350 Subject Index
Konigsberg bridge problem, 34 Petersen graph, 21, 58, 72, 86, 181
Kuratowski graphs, 28 generalized, 59, 3 16
for higher genus, 53 Petersen's theorem, 36
Kuratowski's theorem, 28, 42fT Piecewise linear, 282
Planarity, 42fT
Ladder: algorithms, 51 ff
circular, 59 Fary's theorem, 48
Mobius, 59 Kuratowski's theorem, 42fT
Lift: Mac Lane's theorem, 42fT
offace, 163 Tutte' s theorem, 48
of rotation, 162 Whitney's Duality theorem, 54
of walk, 62 Whitney's Uniqueness theorem, 49
Linear map of simplexes, 282 Projective plane, 26, 102
Line graph, 3 3 Proulx's classification of groups of genus one,
Link of vertex, 107 300fT
Local genus minimum, 149 Pseudofree, 182
Local voltage group, 87, 169 Pseudosurface, 104
Log of circuits, 253 two-fold triple systems, 105
Longitude, 120
Loop, 3 Quatemion group, 299
Quotient:
Manifold (n-manifold), 95 graph, 22
boundary of, 95 natural group action, 270
closed, 96 space, 186
Maschke's classification of genus zero groups, surface, 186, 270
290, 314
Meridian, 120 Rank of group, 250
Mobius band, 25 Reflection, 269
dividing circles of, 269
Octahedron (cocktail-party) graph, 20, 206 halves of, 269
crossing number for, 214 Regular covering, 23, 186
Ordinary voltages, 57 graph, 13
Orientability: Relator, 306
algorithm for surface, Ill reduced, 306
of surface, 6, 10 Riemann- Hurwitz equation, 179, 284
test for derived surface, 170 Riemann surface, 174
type of edge or walk, II 0 Ringel, Gerhard, 30, 191
Orientation: Ringel- Youngs Theorem, 30, 215fT
of imbedding, l 06 Rotation:
of surface, l 06 system, 113
of triangulation, l 06 as permutation, 161
projection of, 113
Patchwork, 155 pure, 113
disjoint, 15 5 vertex form, 114
even, 156 at vertex, 112
quadrilateral, 156
residual faces of, 156 Schemes, 192
Path (n-path), 8 Schoenfliess theorem, l 00
Permutation voltage graph, 81 Schreier color graph, 73
base graph, 81 Schreier (coset) graph, 7 3
derived graph, 81 alternative, 73
imbedded, 166 Self- adjacent, 4
derived imbedding, 166 Self-dual imbeddings:
face lifting, 16 7 of Cayley graphs, 208
natural projection, 81 of complete graphs, 208, 213
Subject Index 351
Simplex (geometric k-simplex), 96ff Twofold triple system, 105
barycenter of, 99 Twisting an edge, 134
face of, 96 Type-0, -1:
Simplicial complex, 96 of edge, 110
abstract, 97 of walk, 110
isomorphism of, 97
carrier of, 96 Valence, 3
m- complex, 96 average, 216
r- skeleton, 96 sequence, 7
subdivision, 97 Voltage assignment, 57ff
first barycentric, 99, 107 natural, 251
Slide of one edge along another, 125 net:
Space group, 277 on cycle, 61
Euclidean, 277 on path, 61
classification, 2 77 ordinary, 57ff
hyperbolic, 279 permutation, 81 ff
spherical, 279 relative, 74ff
Star of vertex, 24, 106 Voltage graph:
Surface, 24ff, 96ff imbedded, 162ff
classification, 119ff derived imbedding, 163
closed, 24, 96 derived rotation system, 163
crosscap number of, 29 dual to current graph, 204
disk sum, 131 face lifting, 16 3
Euler characteristic of, 130 natural action of voltage group, 186
genus of, 29 natural projection, 177
nonorientable, 25 orientability test, 170
orientable, 25 type-0 local group, 169
Suspension (join), 20 ordinary, 57ff
Symmetric, 13 base graph, 57
genus, 264ff derived graph, 57
strong, 264ff fiber, 60
group on n symbols, 13 natural (left) action of voltage group,
imbedding, 264ff 66
strong, 264ff natural projection, 60
weak, 264ff permutation, 81
Symmetry group, 276 base graph, 81
derived graph, 81
Torus, 30, 276 fiber, 81
Tree, 8 Voltage group, 57
T- potential for, 8 9 local, 87
T- voltage for, 90 natural action of, 66
Triangle group, 279ff
Euclidean, 279 Walk, 8
full (p, q, r), 280 boundary, 101
hybrid (m, m, n), 290 closed, 8
hyperbolic, 280 length of, 8
ordinary~ q, r), 281 open, 8
presentation of, 281 Wheel graph, 23
proper quotient of, 298 White-Pisanski imbedding, 155ff
spherical, 280 Word:
Triangulation: in group, 14, 306
orientation for, 106 problem, 14
of surface, 98 Wrapped covering, 207
of topological space, 98
Tubes, 157 Youngs, J. W. T., 30