You are on page 1of 11

Euler's Polyhedral Formula

1. Introduction
It's coming to the end of the calendar year and a lot of people are producing lists. What were the 10
largest box-office blockbusters? What were the 10 best movies of the year? Who are the 10 best
dressed men and 10 worst dressed women? One can also construct more grandiose lists. Who were
the 10 best pitchers of all time or what were the 10 greatest movies? What appears on a list
constructed by the same person can change dramatically with slight wording changes. Thus, the list
of my 10 favorite movies might not coincide with my list of the 10 greatest movies ever made.
Does it make sense to construct lists related to mathematics? What about a list of the 10 greatest
mathematicians? 10 greatest women mathematicians? The 10 most influential theorems? The 10
niftiest theorems?
On the one hand constructing lists is perhaps silly. How can one make a list of the 10
greatest composers of classical music? Must I leave Tchaikovsky out to include Mahler or
Handel? Yet, from another perspective constructing lists of this kind makes one think
about a wide variety of value-laden issues. What makes a composer great? Should a
composer of a few great pieces be put on a short list of greats while another composer
who perhaps composed nothing that rose to the heights of the first person, yet composed
100 times as many pieces at a very high level of inspiration, is omitted? This being the first
of my last two columns as solo editor of the Feature Column, perhaps readers will indulge
me if I write two columns about a theorem which would make both my list of 10 favorite
theorems and my list of 10 most influential theorems. This theorem involves Euler's
polyhedral formula (sometimes called Euler's formula).
Today we would state this result as: The number of vertices V, faces F, and edges E in a
convex 3-dimensional polyhedron, satisfy V + F - E = 2. Aspects of this theorem illustrate
many of the themes that I have tried to touch on in my columns.
2. Basic ideas
Polyhedra drew the attention of mathematicians and scientists even in ancient times. The
Egyptians built pyramids and the Greeks studied "regular polyhedra," today sometimes
(Leonard
referred to as the Platonic Solids . What is a polyhedron? This question is remarkably hard
Euler )
to answer simply! Part of the problem is that as scholars have studied "polyhedral" objects
in greater detail, the idea of what is a "legal" polyhedron has changed and evolved. "Traditional
polyhedra" consist of flat faces, straight edges, and vertices, but exactly what rules do these
individual parts have to obey, and what additional global conditions should be imposed? Which of the
following objects below should be allowed to qualify as polyhedra?

a. A cube with a triangular tunnel bored through it.


(Problem: The "faces" that lie in planes are not always polygons.)
b. The portion of the surface of three pairwise intersecting vertical planes (e.g. "triangular cylinder").
(Problem: This surface does not have any vertices.)
c. The surface formed based on three rays which meet at a point as shown below.

(Problem: The "faces" are not polygons but unbounded portions of planes.)
Almost certainly, in the early days of the study of polyhedra, the word referred to convex polyhedra. A
set is convex if the line segment joining any two points in the set is also in the set. Among the nice
properties of convex sets is the fact that the set of points in common to a collection of convex sets is
convex. Loosely speaking, non-convex sets in two dimensions have either notches or holes. The
diagram below shows a non-convex polygon and a convex polygon.

The planar set below is not convex, but note that it does not satisfy the usual definition of a polygon,
even through it is bounded by sections of straight lines.

The faces of a convex polyhedron consist of convex polygons. However, this approach to defining
polyhedra rules out a "polyhedron" which goes off to infinity, such as the surface below:

This polyhedron has three rays (which, if extended, should meet at a point) and three line segments
as edges of the polyhedron, rather than having edges which are line segments. One can also have
examples where there are only rays (see earlier diagram).
Traditionally, what is essential for a polyhedron is that it consist of pieces of flat surfaces, but as time
has gone on, the definition of "legal" polyhedra has been broadened. Typically this has resulted in a
dramatic improvement in the insights that have been obtained concerning the original "narrower"
class of polyhedra and the newer class of polyhedra under the broader definition.
For example, in Euclid's Elements there is a "proof" that there are 5 regular polyhedra, based on the
implicit assumption that the faces of the polyhedra are convex polygons. A regular polyhedron is one
in which all faces are congruent regular (convex) polygons and all vertices are "alike." This means
that there are the same number of regular polygons at every vertex. Using this definition one finds
there are 5 regular polyhedra.
If one has a supply of regular pentagonal polygons such as the one below,

one can assemble 12 of them, three at each vertex, to form the solid known as the regular
dodecahedron.
However, the Greeks knew about the polygon that today is called the pentagram:

This polygon has a good claim to be called a regular polygon because all its sides have equal length
and the angle between two consecutive sides of the polygon is always the same. However, this
polygon, when drawn in the plane, does not define a convex set and the sides of the polygon
intersect each other. Nonetheless, when mathematicians considered the idea of "non-convex"
regular polyhedra where such "star polygons" were permitted, they discovered some new examples
of "regular polyhedra," now known as the Kepler-Poinsot polyhedra . Later H.S.M. Coxeter and
Branko Grnbaum broadened the rules for polyhedra and discovered other "regular polyhedra."
The systematic study of polyhedra began relatively early in the history of mathematics as shown by
the place they hold in Euclid's Elements . Yet it remained until the 18th century before there was a
statement of the result that
V + F - E = 2.
3. Euler's contribution
It appears that Leonard Euler (1707-1783) was the first person to notice the fact that for convex 3dimensional polyhedra V + F - E = 2.

Euler mentioned his result in a letter to Christian Goldbach (of


Goldbach's Conjecture fame) in 1750. He later published two papers in
which he described what he had done in more detail and attempted to
give a proof of his new discovery.
It is sometimes claimed that Descartes (1596-1650) discovered Euler's
polyhedral formula earlier than Euler. Though Descartes did discover
facts about 3-dimensional polyhedra that would have enabled him to
deduce Euler's formula, he did not take this extra step. With hindsight it
is often difficult to see how a talented mathematician of an earlier era
did not make a step forward that with today's insights seems natural,
however, it often happens. If one is wandering through a maze of tunnels in a cave, perhaps one
comes across a large hall of impressive formations with many tunnels. You may explore some of
these without successfully discovering any impressive new formations and go back home contented.
However, it might have turned out that one of the many tunnels you did not have time to explore
would have led you to an even more spectacular chamber. Descartes' Theorem is a very lovely result
in its own right, and in 3 dimensions it is equivalent to Euler's polyhedral formula.
It is tempting to speculate about why all the able mathematicians, artists, and scholars who
investigated polyhedra in the years before Euler did not notice the polyhedral formula. There
certainly are results in Euclid's Elements and in the work of later Greek geometers that appear more
complex than Euler's polyhedral formula. Presumably a major factor, in addition to the lack of
attention paid to counting problems in general up to relatively recent times, was that people who
thought about polyhedra did not see them as structures with vertices, edges, and faces. It appears
that Euler did not view them this way either! He seems to have adopted a fairly conventional view of
polyhedra. In his view the "vertex" of a polyhedron is a solid angle or a part of a "polyhedral cone"
that starts at the vertex. Given that today Euler is credited with being the "father" of graph theory for
having solved the Knigsberg bridge problem using graph theory ideas, one might have expected
him to see the graph associated with a polyhedron. However, he does not
appear to have thought of polyhedra as graphs, a step not exploited until
rather later by Cauchy (1789-1856).
Because he did not look at the polyhedra he was studying as graphs,
Euler attempted to give a proof of the formula based on decomposing a
polyhedron into simpler pieces. This attempt does not meet modern
standards for a proof. His argument was not correct. However, results
proven later make it possible to use Euler's technique to prove the
polyhedral formula.
Although Euler did not give the first correct proof of his formula, one can
not prove conjectures that have not been made. It appears to have been the French mathematician
Adrian Marie Legendre (1752-1833) who gave the first proof, though he did not use combinatorial
methods. Ironically, this quintessentially combinatorial theorem was given a metric proof by

Legendre. Despite using metrical methods the proof is very insightful and clever.
While Euler first formulated the polyhedral formula as a theorem about polyhedra, today it is often
treated in the more general context of connected graphs (e.g. structures consisting of dots and line
segments joining them that are in one piece). Cauchy seems to have been the first person to make
this important connection. He in essence noticed that the graph of a convex polyhedron (or more
generally what today we would call a polyhedron which is topologically homeomorphic to a sphere)
has a planar connected graph. A plane graph is one which has been drawn in the plane in such a
way that the edges meet (intersect) only at a vertex. A graph which has the potential to be drawn in
this way is known as a planar graph. A typical plane graph is shown in the diagram below. The
number of edges at a vertex in such a connected graph (e.g. one piece) is called the degree or
valence of the vertex. For example, in the diagram below the vertices on the vertical mirror line of the
drawing have valence 3, 4, 5, 4, from top to bottom. Do you think there is some convex 3dimensional polyhedron whose graph is isomorphic to this graph? (You will be able to answer this
question using Steinitz's Theorem, which will be treated in the continuation of this column next
month.)

An intuitive way to see that convex polyhedra give rise to plane graphs (it's the converse that is the
difficult part of Steinitz's Theorem) follows. Cut out one of the polyhedron's faces along the edges
that bound the face, and transform the remaining faces into stretchable rubber. Now stretch the
polyhedron with one face removed to lay it flat in the plane so that the removed face becomes the
unbounded region associated with the resulting plane graph.
To help you understand this process look at the box-like polyhedron below, which from a
combinatorial point of view is a "cube." The lines which are red are those edges of the box which are
"hidden" when the box is viewed from the front.

In the diagram below we have highlighted in turquoise the face of the polyhedron that we remove
from this box, and now pretend that the remaining structure is made of rubber.

Figure 1
We now stretch and flatten out the remaining surface so that the turquoise face becomes the
unbounded region of the resulting graph that can be drawn in the plane (see Figure 2 below). The
position of the red edges is shown to help you visualize what is happening. Of course the turquoise
region should be shown as going off to "infinity," even though it appears only as a bounded face in
the original polyhedron. The back face of the box is now shown as having a vertical and horizontal
red edge. The plane that contains this back face can be thought of as the plane into which we stretch
and flatten out the surface of the box with the turquoise face removed. It's fairly clear that the faces
of the original polyhedron are preserved by this process. It is also clear where these faces appear in
the plane graph representation of the polyhedron, except perhaps for the turquoise face in Figure 1.
It is common for people to forget to count this face when they are counting the faces of a graph
drawn in the plane. Plane graphs always have at least one face, the "infinite face."

Figure 2
Although there are other ways of seeing that the graphs of convex 3-dimensional polyhedra or
graphs on the sphere have planar graphs (i.e. at some stage use stereographic projection), this
argument hopefully makes clear that the graphs that can be drawn on a sphere (polyhedra whose
graphs are homeomorphic to a sphere) where edges meet only at vertices are the same as the
graphs that can be drawn in a plane where edges cross only at vertices.
4. Proofs of the polyhedral formula
There are many proofs of the Euler polyhedral formula, and, perhaps, one indication of the
importance of the result is that David Eppstein has been able to collect 17 different proofs . In a
sense the most straightforward proofs are ones using mathematical induction. One can prove the
result by doing induction on either the number of edges, faces or vertices of the graph.
The following proof is especially attractive. It follows the development given by H. Rademacher and
O. Toeplitz based on an approach of Von Staudt (1798-1867). Suppose that G is a connected graph
that has been embedded in the plane. Think of the infinite face of the graph as being the ocean, the
edges of the graphs being dikes made of earth, and the faces of the graph other than the infinite face
as being dry fields. A typical example of such a connected plane graph is shown below.

The blue in this diagram does not go off to infinity but you should think of it as doing that. In the
diagram below an edge of one circuit in the graph which bounds a face is shown in red, and this
edge is adjacent to the ocean. When the dike represented by this red edge is "breached," the field on
the other side of the dike is opened up to the ocean and becomes flooded, as indicated by the
turquoise (for illustrative purposes - the color of the water really should be the same as the ocean!).

If a connected graph has circuits that bound faces, one can eliminate one edge from each of these
circuits in a way that preserves the property that the graph remains connected. There is a one-to-one
correspondence between the edges removed to get to a connected graph without circuits (because
of the Jordan curve theorem) and the faces of the graph which are not the infinite face. The edges
necessary to do this are shown in red below, and the fact that the fields flood is shown in turquoise.

The resulting graph of black edges forms a spanning tree of the original graph. A spanning tree of a
connected graph is a subgraph which is a tree (e.g. a graph which is connected and has no circuit)
and includes all the vertices of the original graph. Thus, a spanning tree of a connected graph has
the same number of vertices as the original graph, but fewer edges (unless the original graph is a
tree). Note that for any tree, we have:

Now, let us count the edges of the original graph by adding the number of edges removed to get a
tree, the red edges above, to the number of black edges above. Thus,

Now, we know that ERED is equal to one less than the number of faces of the original graph. This is
true because we needed one red edge to flood each dry field (e.g. the non-infinite faces), and the
graph had one more face, the infinite face. The number of EBLACK edges is given by one less than the
number of vertices in the graph, because the black edges are a spanning tree of the original graph.
Hence we have:

Rearranging the terms in this equation gives:

Euler's formula does not hold for any graph embedded on a surface. It holds for graphs embedded
so that edges meet only at vertices on a sphere (or in the plane), but not for graphs embedded on
the torus, a one-holed donut. The fundamental theorem that is used directly or indirectly in a proof of
the Euler polyhedra formula for graphs depends on the Jordan Curve theorem , which states that any
simple closed curve divides the plane into three sets, those points on the curve, those inside the
curve, and those outside the curve.
As interest in Euler's polyhedral formula evolved in the 19th century, many attempts were made to
generalize the formula. This research laid the foundation for the development of topology and
algebraic topology and for a theory of surfaces. It was soon discovered that some surfaces could be
"oriented" in a consistent way only locally but not globally. The famous Mbius (1790-1868) Band
serves as an illustration of a surface of this kind. (This surface was independently discovered
somewhat earlier by Johann Listing (1808-1882).) This research led to a classification of "orientable
surfaces" in terms of what is known as the Euler Characteristic . This concept involves the notion of
the genus of a graph: the smallest number of handles g that must be added to the surface of a
sphere so that a graph can be embedded on the augmented surface in such a way that edges meet
only at vertices. It turns out that every orientable surface in Euclidean space can be thought of as a
sphere with a certain number of handles. To construct a handle on a sphere so as to eliminate an
edge crossing for a graph on the surface, one can cut out two circles from the surface and join these
circles by a cylindrical tube, the handle, as shown below.

If two edges in a graph drawn on a surface cross, one can eliminate the crossing by having one of
the edges go over a small handle to avoid the crossing. It turns out that if a graph can be embedded
without edges meeting anywhere other than at vertices on a surface with g handles, but on no
surface with a smaller number of handles, then:

Since the sphere has no handles, g = 0 for the sphere, and the formula above reduces to Euler's
formula. The connection between Euler's polyhedral formula and the mathematics that led to a
theory of surfaces, both the orientable and unorientable surfaces, is still being pursued to this day. In
addition, motivated by many problems involving the design of computer chips (integrated circuits),
there has been an explosion of research about crossing number problems for graphs in the plane.
This involves finding the minimum number of crossings when an abstractly defined graph is drawn in
the plane. It turns out that the crossing number depends on whether or not the edges are allowed to
be straight or curved. Many of these questions involve the use of Euler's formula to get estimates for
the smallest number of crossings. Furthermore, Euler's formula is intimately connected with coloring
problems of the faces of a graph. These coloring problems include not only the famous 4-color
conjecture but also problems about coloring the faces of graphs on surfaces such as spheres with
many handles.

You might also like