Geometri PDF
Geometri PDF
with an Introduction to
Cosmic Topology
2018 Edition
Michael P. Hitchman
Michael P. Hitchman
Department of Mathematics
Lineld College
McMinnville, OR 97128
http://mphitchman.com
2018 Edition
ISBN-13: 978-1717134813
ISBN-10: 1717134815
Cover image: Five geodesic paths between two points in a two-holed torus with
constant curvature.
Preface
iii
iv
provide possible shapes of the universe. Sections 8.2 and 8.3 present two
research programs in cosmic topology: cosmic crystallography and circles-in-
the-sky. Measurements taken and analyzed over the last twenty years have
greatly altered the way many cosmologists view the universe, and the text ends
with a discussion of our present understanding of the state of the universe.
Compass and ruler constructions play a visible role in the text, primarily
because inversions are emphasized as the basic building blocks of transfor-
mations. Constructions are used in some proofs (such as the Fundamental
Theorem of Möbius Transformations) and as a guide to denitions (such as
the arc-length dierential in the hyperbolic plane). We encourage readers to
practice constructions as they read along, either with compass and ruler on
paper, or with software such as The Geometer's Sketchpad or Geogebra. Some
Geometer's Sketchpad templates and activites related to the text can be found
at the text's website.
Changes from the previously published version For those familiar with
the oringial version of the text published by Jones & Bartlett, we note a few
changes in the current edition. First, the numbering scheme has changed, so
Example and Theorem and Figure numbers will not match the old hard copy.
Of course the numbering schemes on the website and the new print options
of the text do agree. Second, several exercises have been added. In sections
with additional exercises, the new ones typically appear at the end of the
section. Finally, Chapter 7 has been reorganized in an eort to place more
emphasis on the family (Xk , Gk ), and the key theorems common to all these
geometries. This family now receives its own section, Section 7.4. The previous
Section 7.4 (Observing Curvature in a Universe) has been folded into Section
7.3. Finally, the essays in Chapter 8 on cosmic topology and our understanding
of the universe have been updated to include research done since the original
publication of this text, some of which is due to sharper measurements of the
temperature of the cosmic microwave background radiation obtained with the
launch of the Planck satellite in 2009.
Acknowledgements
Many people helped with the development of this text. Je Weeks' text The
Shape of Space inspired me to rst teach a geometry course motivated by
cosmic topology. Colleagues at The College of Idaho encouraged me to teach
the course repeatedly as the manuscript developed, and students there provided
valuable feedback on how this story can be better told.
I would like to thank Rob Beezer, David Farmer, and all my colleagues
at the UTMOST 2017 Textbook Workshop and in the PreTeXt Community
for helping me convert my dusty tex code to a workable PreTeXt document
in order to make the text freely available online, both as a webpage and as a
printable document. I also thank Jennifer Nordstrom for encouraging me to
use PreTeXt in the rst place.
I owe my colleagues at The College of Idaho and the University of Oregon
a debt of gratitude for helping to facilitate a sabbatical during which an earlier
incarnation of this text developed and was subsequently published in 2009 with
Jones & Bartlett. Richard Koch discussed some of its content with me, and Jim
Dull patiently discussed astrophysics and cosmology with me at a moment's
notice.
The pursuit of detecting the shape of the universe has rapidly evolved over
the last twenty years, and Marcelo Rebouças kindly answered my questions
regarding the content of Chapter 8 back in 2008, and pointed me to the
latest papers. I am also grateful to the authors of the many cosmology papers
accessible to the amateur enthusiast and written with the intent to inform.
Finally, I would like to thank my family for their support during this
endeavor, for letting this project permeate our home, and for encouraging
its completion when we might have done something else, again.
v
vi
To my son, Jasper
Contents
Preface iii
Acknowledgements v
1 An Invitation to Geometry 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 A Brief History of Geometry . . . . . . . . . . . . . . . . . . . 3
1.3 Geometry on Surfaces: A First Look . . . . . . . . . . . . . . . 7
3 Transformations 25
3.1 Basic Transformations of C . . . . . . . . . . . . . . . . . . . . 25
3.2 Inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3 The Extended Plane . . . . . . . . . . . . . . . . . . . . . . . . 44
3.4 Möbius Transformations . . . . . . . . . . . . . . . . . . . . . . 47
3.5 Möbius Transformations: A Closer Look . . . . . . . . . . . . . 54
4 Geometry 63
4.1 The Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Möbius Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5 Hyperbolic Geometry 72
5.1 The Poincaré Disk Model . . . . . . . . . . . . . . . . . . . . . 72
5.2 Figures of Hyperbolic Geometry . . . . . . . . . . . . . . . . . 77
5.3 Measurement in Hyperbolic Geometry . . . . . . . . . . . . . . 81
5.4 Area and Triangle Trigonometry . . . . . . . . . . . . . . . . . 90
5.5 The Upper Half-Plane Model . . . . . . . . . . . . . . . . . . . 103
ix
x Contents
Albert Einstein
János Bolyai
1.1 Introduction
a c
The innite plane model of the two-dimensional universe works well enough
for most purposes, but cosmologists and mathematicians, who notice that
everything within the universe is nite, consider the possibility that the
universe itself is nite. Would a nite universe have a boundary? Can it
have an edge, a point beyond which one cannot travel? This possibility is
1
2 1. An Invitation to Geometry
Thus, the top edge of the rectangle has been identied, point by point,
with the bottom edge. In three dimensions one can physically achieve this
identication, or gluing, of the edges. In particular, one can bend the rectangle
to produce a cylinder, being careful to join only the top and bottom edges
together, and not any other points. The left and right edges of the rectangle
have now become the left and right circles of the cylinder, which themselves
get identied, point by point. Bend the cylinder to achieve this second gluing,
and one obtains a donut, also called a torus.
Of course your two-dimensional self would not be able to see this torus
surface in 3-space, but you could understand the space perfectly well in its
rectangle-with-edges-identied form. It is clearly a nite area universe without
any edge.
A sphere, like the surface of a beach ball, is another nite area two-
dimensional surface without any edge. A bug cruising around on the surface
of a sphere will observe that locally the world looks like a at plane, and that
the surface has no edges.
Consideration of a nite-area universe leads to questions about the type
of geometry that applies to the universe. Let's look at a sphere. On small
scales Euclidean geometry works well enough: small triangles have angle sum
◦
essentially equal to 180 , which is a dening feature of Euclidean geometry.
But on a larger scale, things go awry. A very large triangle drawn on the
◦
surface of the sphere has an angle sum far exceeding 180 . (By triangle, we
1.2. A Brief History of Geometry 3
mean three points on the surface, together with three paths of shortest distance
between the points. We'll discuss this more carefully later.)
Consider the triangle formed by the north pole and two points on the
◦
equator in Figure 1.1.4. The angle at each point on the equator is 90 , so the
◦
total angle sum of the triangle exceeds 180 by the amount of the angle at the
north pole. We conclude that a non-Euclidean geometry applies to the sphere
on a global scale.
1. One can draw a straight line from any point to any point.
5. If a straight line falling on two straight lines makes the interior angles
on the same side less than two right angles, the two straight lines, if
produced indenitely, meet on that side on which the angles are less
than two right angles.
4 1. An Invitation to Geometry
Does one postulate not look like the others? The rst four postulates are
short, simple, and intuitive. Well, the second might seem a bit odd, but all
Euclid is saying here is that you can produce a line segment to any length you
want. However, the 5th one, called the parallel postulate , is not short or
simple; it sounds more like something you would try to prove than something
you would take as given.
Indeed, the parallel postulate immediately gave philosophers and other
thinkers ts, and many tried to prove that the fth postulate followed from
the rst four, to no avail. Euclid himself may have been bothered at some level
by the parallel postulate since he avoids using it until the proof of the 29th
proposition in his text.
In trying to make sense of the parallel postulate, many equivalent
statements emerged. The two equivalent statements most relevant to our study
are these:
0
5 . Given a line and a point not on the line, there is exactly one
line through the point that does not intersect the given line.
For God's sake, I beseech you, give it up. Fear it no less than the
sensual passions because it too may take all your time and deprive
you of your health, peace of mind and happiness in life.
1
But János continued to work on the problem, as did the Russian mathematician
Nikolai Lobachevsky (1792-1856). They independently discovered that a well-
dened geometry is possible in which the rst four postulates hold, but the
fth doesn't. In particular, they demonstrated that the fth postulate is not
a necessary consequence of the rst four.
In this text we will study two types of non-Euclidean geometry. The
rst type is called hyperbolic geometry, and is the geometry that Bolyai and
Lobachevsky discovered. (The great Carl Friedrich Gauss (1777-1855) had
also discovered this geometry; however, he did not publish his work because
he feared it would be too controversial for the establishment.) In hyperbolic
geometry, Euclid's fth postulate is replaced by this:
5H. Given a line and a point not on the line, there are at least two
lines through the point that do not intersect the given line.
◦
In hyperbolic geometry, the sum of the angles of any triangle is less than 180 ,
a fact we prove in Chapter 5.
The second type of non-Euclidean geometry in this text is called elliptic
geometry, which models geometry on the sphere. In this geometry, Euclid's
fth postulate is replaced by this:
1 See for instance, Martin Gardner's book The Colossal Book of Mathematics , W.W.
Norton & Company (2001), page 176.
1.2. A Brief History of Geometry 5
5E. Given a line and a point not on the line, there are zero lines
through the point that do not intersect the given line.
◦
In elliptic geometry, the sum of the angles of any triangle is greater than 180 ,
a fact we prove in Chapter 6.
2 See, for instance, The Brothers Karamazov, Fyodor Dostoevsky (a new translation by
Richard Pevear and Larissa Volokhonsky), North Point Press (1990), page 235.
6 1. An Invitation to Geometry
K C
A L B
D M E
Even galactic triangles determined by the positions of three nearby stars have
◦
angle sum indistinguishable from 180 .
However, on a larger scale, things might be dierent.
Maybe we live in a universe that looks at (i.e., Euclidean) on smallish
scales but is curved globally. This is not so hard to believe. A bug living in
a eld on the surface of the Earth might reasonably conclude he is living on
an innite plane. The bug cannot sense the fact that his at, visible world
is just a small patch of a curved surface (Earth) living in three-dimensional
space. Likewise, our apparently Euclidean three-dimensional universe might
3 Relativity: The Special and General Theory , Crown Publications Inc (1961), p. 148.
1.3. Geometry on Surfaces: A First Look 7
be curving in some unseen fourth dimension so that the global geometry of the
universe might be non-Euclidean.
Under reasonable assumptions about space, hyperbolic, elliptic, and
Euclidean geometry are the only three possibilities for the global geometry of
our universe. Researchers have spent signicant time poring over cosmological
data in hopes of deciding which geometry is ours. Deducing the geometry of
the universe can tell us much about the shape of the universe and perhaps
whether it is nite. If the universe is elliptic, then it must be nite in volume.
If it is Euclidean or hyperbolic, then it can be either nite or innite. Moreover,
each geometry type corresponds to a class of possible shapes. And, if that isn't
exciting enough, the overall geometry of the universe may be fundamentally
connected to the fate of the universe. Clearly there is no more grand application
of geometry than to the fate of the universe!
Exercises
1. Use Euclid's parallel postulate to prove the alternate interior angles
theorem. That is, in Figure 1.2.3(a), assume the line BD is parallel to the line
AC . Prove that ∠BAC = ∠ABD.
2. Use Euclid's parallel postulate and the previous problem to prove that
◦
the sum of the angles of any triangle is 180 . You may nd Figure 1.2.3(b)
helpful, where segment CD is parallel to segment AB .
D B B D
A C A C
(a) (b)
Think for a minute about the space we live in. Think about objects that live
in our space. Do the features of objects change when they move around in our
space? If I pick up this paper and move it across the room, will it shrink? Will
it become a broom?
If you draw a triangle on this page, the angles of the triangle will add
◦
to 180 . In fact, any triangle drawn anywhere on the page has this property.
Euclidean geometry on this at page (a portion of the plane) is homogeneous :
the local geometry of the plane is the same at all points. Our three-dimensional
space appears to be homogeneous as well. This is nice, for it means that if we
3 3
buy a 5 ft freezer at the appliance store, it doesn't shrink to 0.5 ft when we
get it home. A sphere is another example of a homogenous surface. A two-
dimensional bug living on the surface of a sphere could not tell the dierence
(geometrically) between any two points on the sphere.
The surface of a donut in three-dimensional space (see Figure 1.3.1) is not
homogeneous, and a two-dimensional bug living on this surface could tell the
dierence between various points. One approach to discovering dierences in
geometry involves triangles.
8 1. An Invitation to Geometry
In the Euclidean plane, geodesics are Euclidean lines. One way to determine
a geodesic on a surface physically is to pin some string at A and draw the string
tight on the surface to a point B. The taut string will follow the geodesic from
A to B. In Figure 1.3.3 we have drawn geodesic triangles on three dierent
surfaces.
Figure 1.3.3: Depending on the shape of the surface, geodesic triangles can
◦
have angle sum greater than, less than, or equal to 180 .
She'd go home, write up the result, emphasizing the fact that a triangle in the
◦
rst convex region will have angle sum greater than 180 , while a triangle
◦
in the saddle-shaped region will have angle sum less than 180 . This happy
bug will conclude her donut surface is not homogeneous. She will then sit back
and watch the accolades pour in. Perhaps even a Nobel prize. Thus, small
triangles and their angles can help a two-dimensional bug distinguish points
on a surface.
The donut surface is not homogeneous, so let's build one that is.
B
A θ
C C
don't get the tip of the cone on the inside of the triangle, adjust the
points accordingly.
c. With your protractor, carefully measure the angle θ subtended
by the circular sector. To emphasize θ's role in the shape of the cone,
we let S(θ) denote the cone surface determined by θ.
d. With your protractor, carefully measure the three angles of your
triangle. The angle at point C is the sum of the angles formed by the
triangle legs and the radial segments. Let ∆ denote the sum of these
three angles.
e. State a conjecture about the relationship between the angle θ
and ∆, the sum of the angles of the triangle. Your conjecture can be
in the form of an equation. Then prove your conjecture. Hint: if you
draw a segment connecting the 2 copies of point C, what is the angle
sum of the quadrilateral ABCC ?
b H c
c b
a c
c
a c b
b c a
It turns out every surface can be given one of three types of homogeneous
geometry: Euclidean, hyperbolic, or elliptic. We will return to the geometry of
surfaces (and of our universe) after we develop hyperbolic and elliptic geometry.
12 1. An Invitation to Geometry
If it doesn't make a whole lot of sense right now, don't sweat it, but please use
these facts as motivation for learning about these non-Euclidean geometries.
Exercises
1. Work through the Coneland Example 1.3.5.
Within this framework, i = (0, 1), meaning that any complex number (x, y)
can be expressed as x + yi as suggested here:
The expression x+yi is called the Cartesian form of the complex number.
This form can be helpful when doing arithmetic of complex numbers, but it
can also be a bit gangly. We often let a single letter such as z or w represent a
complex number. So, z = x + yi means that the complex number we're calling
z corresponds to the point (x, y) in the plane.
It is sometimes helpful to view a complex number as a vector, and complex
addition corresponds to vector addition in the plane. The same holds for scalar
multiplication. For instance, in Figure 2.1.1 we have represented z = 2 + i,
w = −1 + 1.5i, as well as z + w = 1 + 2.5i, as vectors from the origin to these
13
14 2. The Complex Plane
z+w
w w z−w
i z z
Note that |z| gives the Euclidean distance of z to the point (0,0).
The conjugate of z = x + yi, denoted z, is
z = x − yi.
In the exercises the reader is asked to prove various useful properties of the
modulus and conjugate.
z · w = (3 − 4i)(2 + 7i)
= 6 + 28 − 8i + 21i
= 34 + 13i.
4z = 12 − 16i
p
|z| = 32 + (−4)2 = 5
zw = 34 − 13i.
Exercises
1. z + w, sz , |z|, and z · w.
In each case, determine
a. z = 5 + 2i, s = −4, w = −1 + 2i.
b. z = 3i, s = 1/2, w = −3 + 2i.
c. z = 1 + i, s = 0.6, w = 1 − i.
2. 2
Show that z · z = |z| , where z is the conjugate of z .
2.2. Polar Form of a Complex Number 15
A point (x, y) in the plane can be represented in polar form (r, θ) according to
the relationships in Figure 2.2.1.
(x, y)
r x = r cos(θ)
θ y = r sin(θ)
x + yi = r cos(θ) + r sin(θ)i
= r(cos(θ) + i sin(θ)).
This leads us to make the following denition. For any real number θ, we
dene
eiθ = cos(θ) + i sin(θ).
For instance, eiπ/2 = cos(π/2) + i sin(π/2) = 0 + i · 1 = i.
i0
Similarly, e = cos(0) + i sin(0) = 1, and it's a quick check to see that
eiπ = −1, which leads to a simple equation involving the most famous numbers
in mathematics (except 8), truly an all-star equation:
eiπ + 1 = 0.
z = x + yi and (x, y) has polar form (r, θ) then z = reiθ is called
If the
polar form of z. The non-negative scalar |r| is the modulus of z, and the
angle θ is called the argument of z , denoted arg(z ).
16 2. The Complex Plane
z −3 + 4i
4 r
v
α θ
3
u
−1
−3 + 4i = 5ei(π−tan (4/3))
≈ 5e2.21i .
Theorem 2.2.3. The product of two complex numbers in polar form is given
by
reiθ · seiβ = (rs)ei(θ+β) .
Proof. We use the denition of the complex exponential and some
trigonometric identities.
where the equation is taken modulo 2π . That is, depending on our choices
for the arguments, we have arg(vw) = arg(v) + arg(w) + 2πk for some integer
k.
Example 2.2.4: Polar form with r ≥ 0.
When representing a complex number z in polar form as z = reiθ , we
may assume that r is non-negative. If r < 0, then
reiθ = −|r|eiθ
= (eiπ ) · |r|eiθ since − 1 = eiπ
= |r|ei(θ+π) , by Theorem 2.2.3.
2.3. Division and Angle Measure 17
Exercises
1. Convert the following points to polar form and plot them: 3 + i, −1 − 2i,
3 − 4i, 7, 002, 001, and −4i.
2. Express the following points in Cartesian form and plot them: z = 2eiπ/3 ,
w = −2eiπ/4 , u = 4ei5π/3 , and z · u.
3. Modify the all-star equation to involve 8. In particular, write an
expression involving e, i, π, 1, and 8, that equals 0. You may use no other
numbers, and certainly not 3.
2+i 2 + i 3 − 2i
= ·
3 + 2i 3 + 2i 3 − 2i
(6 + 2) + (−4 + 3)i
=
9+4
8−i
=
13
8 1
= − i.
13 13
z 1
=z·
w w
1
= reiθ · e−iβ
s
r i(θ−β)
= e .
s
18 2. The Complex Plane
So,
z
arg = arg(z) − arg(w)
w
where equality is taken modulo 2π .
Thus, when dividing by complex numbers, we can rst convert to
polar form if it is convenient. For instance,
√ iπ/4
1+i 2e 1 1
=√ = e−iπ/2 = − i.
−3 + 3i 18e i3π/4 3 3
Angle Measure Given two rays L1 and L2 having common initial point, we
let∠(L1 , L2 ) denote the angle between rays L1 and L2 , measured from L1
to L2 . We may rotate ray L1 onto ray L2 in either a counterclockwise direction
or a clockwise direction. We adopt the convention that angles measured
counterclockwise are positive, and angles measured clockwise are negative, and
admit that angles are only well-dened up to multiples of 2π . Notice that
∠(L1 , L2 ) = −∠(L2 , L1 ).
To compute ∠(L1 , L2 ) where z0 is the common initial point of the rays, let
z1 be any point on L1 , and z2 any point on L2 . Then
z2 − z0
∠(L1 , L2 ) = arg
z1 − z0
= arg(z2 − z0 ) − arg(z1 − z0 ).
L1
3 + 3i
2 + 2i θ
4+i L2
w−v
∠uvw = θ = arg .
u−v
2.4. Complex Expressions 19
θ
u
v
Exercises
1. Express
1
x+yi in the form a + bi.
2. Express these fractions in Cartesian form or polar form, whichever seems
more convenient.
1 1 4+i 2
, , , .
2i 1 + i 1 − 2i 3 + i
u
w
0 1
b 0 a
In this section we look at some equations and inequalities that will come up
throughout the text.
20 2. The Complex Plane
αz + αz + d = 0 (equation of a line)
We may also view any line in C as the collection of points equidistant from
two given points.
Theorem 2.4.2. Any line in C can be expressed by the equation |z−γ| = |z−β|
for suitably chosen points γ and β in C, and the set of all points (Euclidean)
equidistant from distinct points γ and β forms a line.
Proof. Given two points γ and β in C, z is equidistant from both if and
only if |z − γ|2 = |z − β|2 . Expanding this equation, we obtain
(z − γ)(z − γ) = (z − β)(z − β)
|z| − γz − γz + |γ|2 = |z|2 − βz − βz + |β|2
2
√
± r0 eiθ0 /2 .
√
For instance, z2 = i has two solutions. Since i = 1eiπ/2 , i =
√ √ √
±e iπ/4
. In Cartesian form, i = ±( 22 + 22 i).
More generally, the complex quadratic equation αz 2 + βz + γ = 0
where α, β, γ are complex constants, will have one or two solutions.
This marks an important dierence from the real case, where a
quadratic equation might not have any real solutions. In both cases we
may use the quadratic formula to hunt for roots, and in the complex
case we have solutions
p
−β ± β 2 − 4αγ
z= .
2α
For instance, z 2 + 2z + 4 = 0 has two solutions:
√
−2 ± −12 √
z= = −1 ± 3i
2
√
since −1 = i.
z 2 − (3 + 3i)z − (2 − 3i) = 0.
p
(3 + 3i)2 + 4(2 − 3i)
3 + 3i ±
z=
√ 2
3 + 3i ± 8 + 6i
z= .
2
To determine the solutions in Cartesian form, we need to evaluate
√
8 + 6i. We oer two approaches.
√ The rst approach considers the
following task: Set x + yi = 8 + 6i and solve for x and y directly by
squaring both sides to obtain a system of equations.
√
x + yi = 8 + 6i
2
(x + yi) = 8 + 6i
x − y 2 + 2xyi = 8 + 6i.
2
x2 − y 2 = 8 (1)
2xy = 6. (2)
22 2. The Complex Plane
x2 + y 2 = 10. (3)
8 + 6i
10
√ √
We know 8 + 6i = ± 10eiθ/2 , so we want to nd θ/2. Well, we
can determine tan(θ/2) easily enough using the half-angle formula
sin(θ)
tan(θ/2) = .
1 + cos(θ)
The right triangle in the diagram shows us that sin(θ) = 3/5 and
cos(θ) = 4/5, so tan(θ/2) = 1/3. This means that any point reiθ/2 lives
on the line through the origin having slope 1/3, and can be described
√
by k(3
√ + i) for some scalar k . Since 8 + 6i √ has this form,
√ it follows
that
√ + i) for some k . Since | 8 + 6i|
8 + 6i = k(3 √ = 10 , it follows
that |k(3 + i)| = 10, so k = ±1. In other words, 8 + 6i = ±(3 + i).
Now let's return to the solution of the original quadratic equation
in this example:
√
3 + 3i ± 8 + 6i
z=
2
3 + 3i ± (3 + i)
z= .
2
Thus, z = 3 + 2i or z = i.
|z − z0 | = r (equation of a circle)
|z − z0 | = |(x − h) + (y − k)i|
p
= (x − h)2 + (y − k)2 .
2.4. Complex Expressions 23
1. |1/z| > 2.
Taking the reciprocal of both sides, we have |z| < 1/2, which is
the interior of the circle centered at 0 with radius 1/2.
2. Im(z) < Re(z).
3. Im(z) = |z|.
p
Setting z = x + yi, this equation is equivalent to y = y 2 + x2 .
Squaring both sides
p we obtain 0 = x2 , so that x = 0. It
follows that y = y 2 = |y| so the equation describes the points
(0, y) with y ≥ 0. These points determine a ray on the positive
imaginary axis.
Im(z) = |z|
1
2
Im(z) <Re(z)
Moving forward, lines and circles will be especially important objects for
us, so we end the section with a summary of their descriptions in the complex
plane.
|z − z0 | = r.
24 2. The Complex Plane
Exercises
1. Use a complex variable to describe the equation of the line y = mx + b.
Assume m 6= 0. In particular, show that this line is described by the equation
(m + i)z + (m − i)z + 2b = 0.
2. In each case, sketch the set of complex numbers z satisfying the given
condition.
a. |z + i| = 3.
b. |z + i| = |z − i|.
c. Re(z) = 1.
d. |z/10 + 1 − i| < 5.
e. Im(z) > Re(z).
f. Re(z) = |z − 2|.
3. Suppose u, v, w are three complex numbers not all on the same line. Prove
that any point z in C is uniquely determined by its distances from these three
points. Hint: Suppose β and γ are complex numbers such that |u−β| = |u−γ|,
|v − β| = |v − γ| and |w − β| = |w − γ|. Argue that β and γ must in fact be
equal complex numbers.
Transformations will be the focus of this chapter. They are functions rst
and foremost, often used to push objects from one place in a space to a more
convenient place, but transformations do much more. They will be used to
dene dierent geometries, and we will think of a transformation in terms of
the sorts of objects (and functions) that are unaected by it.
f g
A B A B
Basic Transformations of C.
◦ Translation by b: Tb (z) = z + b
25
26 3. Transformations
Tb (v)
b
Tb (w)
v
For instance, the origin gets moved to the point b, (i.e., Tb (0) = b),
and every other point in the plane gets moved the same amount and in
the same direction. It follows that two dierent points, such as v and
w in the diagram, cannot get moved to the same image point (thus,
the function is one-to-one). Also, any point in the plane is the image
of some other point (just follow the vector −b to nd this pre-image
point), so the function is onto as well.
We now oer a formal argument that the translation Tb is a
transformation. Recall, b is a xed complex number.
That Tb is onto :
To show that Tb w denote an arbitrary element of C.
is onto, let
We must nd a complex number z such that Tb (z) = w. Let z = w − b.
Then Tb (z) = z + b = (w − b) + b = w . Thus, T is onto.
That Tb is one-to-one :
To show that Tb is 1-1 we must show that if z1 6= z2 then
Tb (z1 ) 6= Tb (z2 ). We do so by proving the contrapositive. Recall,
the contrapositive of a statement of the form If P is true then Q is
true is If Q is false then P is false. These statements are logically
equivalent, which means we may prove one by proving the other. So, in
the present case, the contrapositive of If z1 6= z2 then Tb (z1 ) 6= Tb (z2 )
is If Tb (z1 ) = Tb (z2 ), then z1 = z2 . We now prove this statement.
Suppose z1 and z2 are two complex numbers such that Tb (z1 ) =
Tb (z2 ). Then z1 + b = z2 + b. Subtracting b from both sides we see that
z1 = z2 , and this completes the proof.
3.1. Basic Transformations of C 27
Rθ (z)
θ z
T−z R Tz
0
z 7−−−→ θ
z − z0 7−−→ eiθ (z − z0 ) 7−−→
0
eiθ (z − z0 ) + z0 .
In other words, the desired rotation R is the composition Tz0 ◦Rθ ◦T−z0
and
R(z) = eiθ (z − z0 ) + z0 .
That the composition of these three transformations is itself a
transformation follows from the next theorem.
Theorem 3.1.5. If T and S are two transformations of the set A, then the
composition S ◦ T is also a transformation of the set A.
Proof. We must prove that S◦T :A→A is 1-1 and onto.
That S ◦ T is onto :
Suppose c is in A. We must nd an element a in A such that S ◦ T (a) = c.
Since S is onto, there exists some element b in A such that S(b) = c.
Since T is onto, there exists some element a in A such that T (a) = b.
Then S ◦ T (a) = S(b) = c, and we have demonstrated that S ◦ T is onto.
That S ◦ T is 1-1 :
Again, we prove the contrapositive. In particular, we show that if
S ◦ T (a1 ) = S ◦ T (a2 ) then a1 = a2 .
If S(T (a1 )) = S(T (a2 )) then T (a1 ) = T (a2 ) since S is 1-1.
And T (a1 ) = T (a2 ) implies that a1 = a2 since T is 1-1.
Therefore, S ◦ T is 1-1.
28 3. Transformations
αz + αz + d = 0
3.1. Basic Transformations of C 29
w−b w−b
α +α +d=0
a a
which can be rewritten
α α αb αb
w+ w+d− − = 0.
a a a a
Now, for any complex number β the sum β + β is a real number, so in the
αb αb
above expression, d − (
a + a ) is a real number. Therefore, all w in T (L)
satisfy a line equation. That is, T (L) is a line.
b. The proof of this part is left as an exercise.
4i
4 + 3i
2i
D T (D)
2 4
L2
z2
L1
z0 z1
Then,
z2 − z0
∠(L1 , L2 ) = arg .
z1 − z0
Since general linear transformations preserve lines, T (Li ) is the line through
T (z0 ) and T (zi ) for i = 1, 2 and it follows that
T (z2 ) − T (z0 )
∠(T (L1 ), T (L2 )) = arg
T (z1 ) − T (z0 )
az2 + b − az0 − b
= arg
az1 + b − az0 − b
z2 − z0
= arg
z1 − z0
= ∠(L1 , L2 ).
z∗
z 7→ z − 5i
π
7→ e− 4 i (z − 5i)
π π
7→ e− 4 i (z − 5i) = e 4 i (z + 5i)
π π π
7→ e 4 i · e 4 i (z + 5i) = e 2 i (z + 5i)
π
7→ e 2 i (z + 5i) + 5i.
π
Simplifying (and noting that e 2 i = i), the reection about the line
L:y =x+5 has formula
rL (z) = iz − 5 + 5i.
rL (z) = eiθ z + b
L2 be the line parallel to L1 through the midpoint of segment 0b. Also let ri
denote reection about line Li for i = 1, 2.
Now, given any z in C, let L be the line through z that is parallel to vector b
(and hence perpendicular to L1 and L2 ). The image of z under the composition
r2 ◦ r1 will be on this line. To nd the exact location, let z1 be the intersection
of L1 and L, and z2 the intersection of L2 and L, (see the gure). To reect
z about L1 we need to translate it along L twice by the vector z1 − z . Thus
r1 (z) = z + 2(z1 − z) = 2z1 − z .
Next, to reect r1 (z) about L2 , we need to translate it along L twice by
the vector z2 − r1 (z). Thus,
Notice from Figure 3.1.18(a) that z2 − z1 is equal to b/2. Thus r2 (r1 (z)) =
z+b is translation by b.
Rotation about the point z0 by angle θ can be achieved by two reections.
The rst reection is about the line L1 through z0 parallel to the real axis,
and the second reection is about the line L2 that intersects L1 at z0 at an
angle of θ/2, as in Figure 3.1.18(b). In the exercises you will prove that this
composition of reections does indeed give the desired rotation.
L1 L2 z
r2 ◦ r1 (z) L2
θ
r1 (z) r2 ◦ r1 (z) L θ
z1 z2 2 L1
z z0
b
0 b
2 r1 (z)
(a) (b)
Exercises
1. Is T (z) = −z a translation, dilation, rotation, or none of the above?
3.2. Inversion 33
4. Suppose T is a rotation by 30
◦
about the point 2, and S is a rotation by
◦
45 about the point 4. What is T ◦ S? Can you describe this transformation
geometrically?
7. x = k.
Find a formula for reection about the vertical line
11. S(z) = kz is a dilation about the origin. Find an equation for a dilation
of C by factor k about an arbitrary point z0 in C.
3.2 Inversion
Inversion oers a way to reect points across a circle. This transformation plays
a central role in visualizing the transformations of non-Euclidean geometry, and
this section is the foundation of much of what follows.
Suppose C r and center z0 . Inversion in the circle
is a circle with radius
C sends a point z 6= z0 to the point z ∗ dened as follows: First, construct the
ray from z0 through z . Then, let z ∗ be the unique point on this ray that
satises the equation
|z − z0 | · |z ∗ − z0 | = r2 .
The point z∗ is called the symmetric point to z with respect to C .
r z∗
z
z0
versa. The closer z gets to the center of the circle, the further iC (z) gets from
the circle.
|z| · |z ∗ | = 1.
r2
iC (z) = + z0 .
(z − z0 )
iC (z)
iC (w)
w
z iC (v)
iC (p) v
p
z0
u iC (u)
As Example 3.2.2 suggests, the distinction between lines and circles gets
muddied a bit by inversion. A line can get mapped to a circle and vice versa.
In what follows, it will be helpful to view reection in a line and inversion in a
circle as special cases of the same general map. To arrive at this view we rst
make lines and circles special cases of the same general type of gure.
3.2. Inversion 35
czz + αz + αz + d = 0
where z = x + yi is a complex variable, α is a complex constant, and c, d are
real numbers. If c = 0 the equation describes a line, and if c 6= 0 and |α|2 > cd
the equation describes a circle.
The word cline (pronounced `Klein') might seem a bit forced, but it
represents the shift in thinking we aim to achieve. We need to start thinking of
lines and circles as dierent manifestations of the same general class of objects.
What class? The class of clines.
Letting α = a + bi and z = x + yi, the cline equation czz + αz + αz + d = 0
can be written as
czz + αz + αz + d = 0
r
|α|2 − cd
Re(α) Im(α)
z0 = − , and r= ,
c c c2
From now on, if you read the phrase inversion in a cline, know that this
means inversion in a circle or reection about a line, and if someone hands you
a cline C, you might say, Thanks! By the way, is this a line or a circle?
We note here the construction of a cline through three points in C. This
construction is used often in later chapters to generate gures in non-Euclidean
geometry.
36 3. Transformations
Theorem 3.2.4. There exists a unique cline through any three distinct points
in C.
Proof. Suppose u, v , and w are distinct complex numbers. If v is on the line
through u and w then this line is the unique cline through the three points.
Otherwise, the three points do not lie on a single line, and we may build a
circle through these three points as demonstrated in Figure 3.2.5. Construct
the perpendicular bisector to segment uv , and the perpendicular bisector to
segment vw. These bisectors will intersect because the three points are not
collinear. If we call the point of intersection z0 , then the circle centered at z0
through w is the unique cline through the three points.
z0 w
Figure 3.2.5: Constructing the unique circle through three points not on a
single line.
czz + αz + αz + d = 0,
where c, d ∈ R, α ∈ C.
We want to show that the image of this cline under inversion in the unit
circle, iS1 (C), is also a cline. Well, iS1 (C) consists of all points w = 1/z , where
z satises the cline equation for C. We show that all such w live on a cline.
If z 6= 0 then we may multiply each side of the cline equation by 1/(z · z)
to obtain
1 1 11
c+α +α +d = 0.
z z zz
But since w = 1/z and w = 1/z , this equation reduces to
c + α · w + α · w + dww = 0,
or
dww + α · w + α · w + c = 0.
Thus, the image points w form a cline equation. If d=0 then the original
cline C passed through the origin, and the image cline is a line. If d 6= 0 then
C did not pass through the origin, and the image cline is a circle. (In fact,
we must also check that |α|2 > dc. This is the case because the original cline
2
equation ensures |α| > cd.)
3.2. Inversion 37
We will call two clines orthogonal if they intersect at right angles. For
instance, a line is orthogonal to a circle if and only if it goes through the
center of the circle. One very important feature of inversion in C is that
clines orthogonal to C get inverted to themselves. To prove this fact, we rst
prove the following result, which can be found in Euclid's Elements (Book III,
Proposition 36).
Proof. Suppose the line throughp does not pass through the center of C , as
in the diagram below. Let q mn, and let d = |q −o|
be the midpoint of segment
as in the diagram. Note also that the line through q and o is the perpendicular
bisector of segment mn. In particular, |m − q| = |q − n|.
n
q
m r
d
p s o
|p − q|2 + d2 = s2 , (1)
|q − n|2 + d2 = r2 . (2)
|p − q|2 − |q − n|2 = s2 − r2 ,
which factors as
We note that the quantity s2 − r2 in the previous lemma is often called the
power of the point p with respect to the circle C . That is, if circle C has radius
r and a point p is a distance s from the center of C then the quantity s2 − r2
is called the power of the point p.
z
C z0 o
s
r k
D
t
k
C L1
m1
C2
z
z∗
m2
C1
L2
The advantage to describing ∠(L1 , L2 ) with these circles is that the image
of the angle, ∠(iC (L1 ), iC (L2 )), is also described by these two circles, at
∗
their other intersection point z . Notice that these angles will have opposite
signs. For instance, in Figure 3.2.11, our initial angle is negative, described
by sweeping arc C1 clockwise onto C2 , but in the image, we sweep iC (C1 )
counterclockwise onto iC (C2 ). We leave it as an exercise for the reader to
check that the angle of intersection of C1 and C2 at z∗ is the same magnitude
as the angle between C1 and C2 at z.
Now we show that inversion preserves angle magnitudes for angles that
occur on the circle C (i.e., z is on C ). C0
be a concentric circle to C . Then
Let
iC (z) = S◦iC 0 where S is a dilation of C whose xed point is the common center
0 0
of circles C and C (see Exercise 3.2.12). Since our angle is not on circle C , iC 0
preserves the magnitude of the angle by reason of the preceding argument. The
dilation S preserves angles according to Theorem 3.1.12. Thus iC preserves
angle magnitudes as well. We leave the case of the angle occurring at the
origin to the next section. Bearing that exception in mind, this completes the
proof.
F E
C
p
p∗
E∗ q
∗
q
F ∗ D∗
D
z∗
1
q
p q∗
|p − z ∗ | |p − q ∗ |
= ,
|p − q| |p − z|
3.2. Inversion 41
|z − q|
|z ∗ − q ∗ | = |z ∗ − p| ·
|p − q|
|z − q| 1
= [|z ∗ − p| · |z − p|] · ·
|z − p| |p − q|
1 1
=1· · .
k |p − q|
Thus, the set D of all points z satisfying |z − p| = k|z − q| has image iC (D)
under this inversion consisting of all points z ∗ on a circle centered at q ∗ with
radius (k|p − q|)−1 . Since inversion preserves clines and p is not on iC (D), it
follows that D itself is a circle.
C1
p z1 q zo L1
C2
t
C
Figure 3.2.15: Finding two points symmetric with respect to a line and circle.
Theorem 3.2.16. Suppose we have two clines that do not intersect, and at
least one of them is a circle. Then there exist two points, p and q , that are
symmetric with respect to both clines.
Proof. L, and the other is a circle C centered
First, assume one cline is a line
at the point z0 L1 be the line through z0 that
as pictured in Figure 3.2.15. Let
is perpendicular to L, and let z1 be the point of intersection of L and L1 . Next,
construct the circle C1 having the diameter z0 z1 . Circle C1 intersects circle C
at some point, which we call t. Notice that ∠z0 tz1 is right, and so the circle
C2 centered at z1 through t is orthogonal to C . Furthermore, the center of C2 ,
z1 , lies on line L, so C2 is orthogonal to L. Let p and q be the two points at
which C2 intersects L1 . By construction, and by using Theorem 3.2.8, p and q
are symmetric to both C and L.
Now, suppose C1 and C2 are circles that do not intersect. We may rst
∗
perform an inversion in a circle C that maps C1 to a line C1 , and C2 to another
∗
circle, C2 , as suggested in Figure 3.2.17 (any circle C centered on a point of C1
works). Then by reason of the preceeding argument, there exist two points p
∗ ∗
and q that are symmetric with respect to C1 and C2 . Since inversion preserves
42 3. Transformations
symmetry points, iC (p) and iC (q) are symmetric with respect to both iC (C1∗ )
∗ ∗ ∗
and iC (C2 ). But iC (C1 ) = C1 and iC (C2 ) = C2 so we're found two points
symmetric to both C1 and C2 . (In fact, we have one exception. If C1 and C2
are concentric circles, this strategy will produce points iC (p) and iC (q), one of
which is the center of C, and we have not yet extended the notion of inversion
to include the center. We do so in the next section in such a way that the
theorem applies to this exceptional case as well.)
C1∗
C1 C2
C2∗
Exercises
1. Prove the general formula for inversion in a circle C centered at z0 with
radius r. In particular, show in this case that
r2
iC (z) = + z0 .
(z − z0 )
(1)
(4)
z∗ (1)
t t z
(3) z z∗
z0
z0 (2)
C C (2)
(a) (b)
Figure 3.2.18: Constructing the symmetric point (a) if z is inside the circle
of inversion; (b) if z is outside the circle of inversion.
7. Prove that inversion in the unit circle maps the circle (x − a)2 + (y − b)2 =
2
r to the circle
2 2 2
a b r
x− + y− =
d d d
where d = a2 + b2 − r2 , provided that d 6= 0.
8. Determine in standard form the image of the circle C given by
(x − 1)2 + y 2 = 4 under inversion in the unit circle. Give a careful plot of
the unit circle, the circle C, and the image of C under the inversion.
10. Suppose C and D are orthogonal circles. Corollary 3.2.9 tells us that
inversion in C maps D to itself. Prove that this inversion also takes the interior
of D to itself.
11. Finish the proof of Theorem 3.2.10 by showing that the angle of
intersection at z∗ equals the angle of intersection at z in Figure 3.2.11.
12. Suppose C is the circle |z −z0 | = r and C is the circle |z −z0 | = r0 . Find
0
13. Complete the proof of Lemma 3.2.7 by proving the case in which the
line through p passes through the center of C.
iC (z) = ∞ if z = z0 ;
z
0 if z = ∞.
iz + (3i + 1)
T (z) = ,
2iz + 1
then T (i/2) = ∞ and T (∞) = 1/2.
We emphasize that the following key results of the previous section extend to
C+ as well:
There exists a unique cline through any three distinct points in C+ . (If
one of the given points in Theorem 3.2.4 is ∞, the unique cline is the
line through the other two points.)
S2 = {(a, b, c) ∈ R3 | a2 + b2 + c2 = 1}.
We will usually refer to the unit 2-sphere as simply the sphere. Stereo-
graphic projection of the sphere onto the extended plane is dened as follows.
Let N = (0, 0, 1) denote the north pole on the sphere. For any point P 6= N
−−→
on the sphere, φ(P ) is the point on the ray N P that lives in the xy -plane. See
Figure 3.3.4 for the image of a typical point P of the sphere.
46 3. Transformations
z
N
P = (a, b, c)
x
φ(P )
y
This line intersects the xy -plane when its z coordinate is zero. This occurs
1 a b
when t=1−c , which corresponds to the point (
1−c , 1−c , 0).
Thus, for a point (a, b, c) on the sphere with c 6= 1, stereographic projection
φ: S2 → C+ is given by
a b
φ((a, b, c)) = + i.
1−c 1−c
Where does φ send the north pole? To ∞, of course. A sequence of points
on S2 that approaches N will have image points in C with magnitudes that
approach ∞.
Exercises
1. In each case nd T (∞) and the input z0 such that T (z0 ) = ∞.
a. T (z) = (3 − z)/(2z + i).
b. T (z) = (z + 1)/eiπ/4 .
c. T (z) = (az + b)/(cz + d).
3.4. Möbius Transformations 47
x2 + y 2 − 1
−1 2x 2y
φ (x, y) = , , .
x2 + y 2 + 1 x2 + y 2 + 1 x2 + y 2 + 1
az + b
=w
cz + d
for z, which is possible so long as a and c are not both 0 (causing the z terms
to vanish). Since ad − bc 6= 0, we can be assured that this is the case, and
solving for z we obtain
−dw + b
z= .
cw − a
Thus T is onto, and T is a transformation.
To prove the converse we show the contrapositive. We suppose ad − bc = 0
and show T (z) = (az + b)/(cz + d) is not a transformation by tackling two
cases.
Case 1 : ad = 0. In this case, bc = 0 as well, so a or d is zero, and b
or c is zero. In all four scenarios, one can check immediately that T is not a
48 3. Transformations
transformation of C+ .
For instance, if a = c = 0 then T (z) = b/d is neither
+
1-1 nor onto C .
Case 2 : ad 6= 0. In this case, all four constants are non-zero, and a/c = b/d.
Since T (0) = b/d and T (∞) = a/c, T is not 1-1, and hence not a transformation
+
of C .
az + b a (bc − ad)/c
T (z) = = + ,
cz + d c cz + d
which can be viewed as the composition T3 ◦ T2 ◦ T1 (z), where T1 (z) = cz + d,
bc−ad a
T2 (z) = 1/z and T3 (z) = c z + c . Note that T1 and T3 are general linear
transformations, and
1 1
T2 (z) = =
z z
3.4. Möbius Transformations 49
is inversion in the unit circle followed by reection about the real axis. Thus,
each Ti is the composition of an even number of inversions, and the general
Möbius transformation T is as well.
To prove the other direction, we show that if T is the composition of two
inversions then it is a Möbius transformation. Then, if T is the composition of
any even number of inversions, it is the composition of half as many Möbius
transformations and is itself a Möbius transformation by Theorem 3.4.3.
Case 1 : T is the composition of two circle inversions. Suppose T = iC1 ◦iC2
where C1 is the circle |z −z1 | = r1 and C2 is the circle |z −z2 | = r2 . For i = 1, 2
the inversion may be described by
ri2
iCi = + zi ,
z − zi
and if we compose these two inversions we do in fact obtain a Möbius
transformation. We leave the details of this computation to the reader but
note that the determinant of the resulting Möbius transformation is r12 r22 .
Case 2 : T is the composition of one circle inversion and one line reection.
Reection in the line can be given by rL (z) = eiθ z + b and inversion
in the
2
r
circle C is given by iC = + zo where z0 and r are the center and
radius
z−zo
of the circle, as usual. Work out the composition and you'll see that we have a
iθ 2
Möbius transformation with determinant e r (which is non-zero). Its inverse,
the composition iC ◦ rL , is also a Möbius transformation.
Case 3 : T is the composition of two reections. Either the two lines of
reection are parallel, in which case the composition gives a translation, or
they intersect, in which case we have a rotation about the point of intersection
(Theorem 3.1.17). In either case we have a Möbius transformation.
It follows that the composition of any even number of inversions yields a
Möbius transformation.
az + b
= z,
cz + d
for z, which gives the quadratic equation
cz 2 + (d − a)z − b = 0. (1)
With this xed point theorem in hand, we can now prove the Fundamental
Theorem of Möbius Transformations, which says that if we want to induce a
one-to-one and onto motion of the entire extended plane that sends my favorite
three points (z1 , z2 , z3 ) to your favorite three points (w1 , w2 , w3 ), as dramatized
below, then there is a Möbius transformation that will do the trick, and there's
only one.
z2
z1
w3
w1 w2
z3
S −1 ◦ T (z1 ) = S −1 (1) = w1
S −1 ◦ T (z2 ) = S −1 (0) = w2
S −1 ◦ T (z3 ) = S −1 (∞) = w3 .
1
0
∞
T S
z1 w1
z2 w2
z3 S −1 ◦ T w3
z−u w−v
(z, w; u, v) = · .
z−v w−u
If z is a variable, and w, u, and v are distinct complex constants, then
T (z) = (z, w; u, v) is the (unique!) Möbius transformation that sends w →
7 1,
u 7→ 0, and v 7→ ∞.
Example 3.4.11: Building a Möbius transformation.
Find the unique Möbius transformation that sends 1 7→ 3, i 7→ 0, and
2 7→ −1.
One approach: Find T (z) = (z, 1; i, 2) and S(w) = (w, 3; 0, −1). In
this case, the transformation we want is S −1 ◦ T .
To nd this transformation, we set the cross ratios equal:
−z + i 4w
= .
(1 − i)z − 2 + 2i 3w + 3
−3z + 3i
V (z) = .
(7 − 4i)z + (−8 + 5i)
It's quite easy to check our answer here. Since there is exactly one
Möbius transformation that does the trick, all we need to do is check
whether V (1) = 3, V (i) = 0 and V (2) = −1. Ok... yes... yes... yep!
We've got our map!
1+1 i+i
(1, i; −1, −i) = ·
1+i i+1
2 2i
=
1+i1+i
4i
=
(1 − 1) + 2i
3.4. Möbius Transformations 53
4i
=
2i
= 2. (Yep!)
Theorem 3.4.15. Given any two clines C1 and C2 , there exists a Möbius
transformation T that maps C1 onto C2 . That is, T (C1 ) = C2 .
Proof. Let p1 C1 and q1 and q1∗ symmetric with respect to
be a point on
C1 . Similarly, let p2 be a point on C2 and q2 and q2∗ be symmetric with respect
to C2 . Build the Möbius transformation that sends p1 7→ p2 , q1 7→ q2 and
q1∗ 7→ q2∗ . Then T (C1 ) = C2 .
Exercises
1. Find a transformation of C+ that rotates points about 2i by an angle
π/4. Show that this transformation has the form of a Möbius transformation.
9. Prove that the cross ratio of four distinct complex numbers is a real
number if and only if the four points lie on the same cline. Hint: Use the
previous exercise and the invariance of the cross ratio.
15. Find a non-trivial Möbius transformation that xes the points -1 and
1, and call this transformation T. Then, let C be the imaginary axis. What is
the image of C under this map. That is, what cline is T (C)?
16. Suppose z1 , z2 , z3 are distinct points in
+
C . Show that by an even
number of inversions we can map z1 7→ 1, z2 7→ 0, and z3 7→ ∞ in the case
that z3 = ∞.
p q
Figure 3.5.1: Type I clines (solid) and Type II clines (dashed) of p and q.
By Theorem 3.2.8, any type II cline of p and q intersects any type I cline of
p and q at right angles. Furthermore, because Möbius transformations preserve
clines and symmetry points, we can be assured that Möbius transformations
preserve type I clines as well as type II clines. In particular, if C is a type I
cline of p and q, then T (C) is a type I cline of T (p) and T (q). Similarly, if C
is a type II cline of p and q , then T (C) is a type II cline of T (p) and T (q). We
can use this to our advantage.
For instance, the type I clines of the points 0 and ∞ are, precisely, lines
through the origin, while the type II clines of 0 and ∞ are circles centered
at the origin. (Remember, inversion in a circle takes the center of the circle
to ∞.) The type I clines in this case are clearly perpendicular to the type
II clines, and they combine to create a coordinate system of the plane (polar
coordinates), as pictured in Figure 3.5.2(a). We can move this system of clines
3.5. Möbius Transformations: A Closer Look 55
p q
(a) (b)
Figure 3.5.2: (a) Type I clines (solid) and type II clines (dashed) of points 0
and ∞ (solid); (b) A Möbius transformation sending 0 7→ p and ∞ 7→ q sends
type I and II clines of 0 and ∞ to type I and II clines of p and q , respectively.
U = S ◦ T ◦ S −1 (1)
Notice
U (0) = S ◦ T ◦ S −1 (0) = S ◦ T (p) = S(p) = 0,
and
U (∞) = S ◦ T ◦ S −1 (∞) = S ◦ T (q) = S(q) = ∞.
That is, is a Möbius transformation that xes 0 and ∞.
U So, by
Example 3.5.3, U is a rotation, a dilation, or some combination of those, and
U looks like U (z) = reiθ z .
In any event, focusing on T again and using equation (1), which can be
rewritten as S ◦ T = U ◦ S , we arrive at the following equation, called the
normal form of the Möbius transformation in this case.
T (z) − p z−p
= reiθ ·
T (z) − q z−q
S U (S(z))
T (z)
z
p q S(z)
S −1
First, z gets sent via S to S(z), which is at the intersection of a line through
the origin and a circle centered at the origin. Second, U (which has the form
U (z) = reiθ z ), sends S(z) along this line through the origin (by dilation factor
r), and then around a new circle centered at the origin (by rotation factor θ)
−1
to the point U (S(z)). Third, S sends U (S(z)) back to the intersection of a
3.5. Möbius Transformations: A Closer Look 57
type I cline of p and q and a type II cline of p and q . This point of intersection
is S −1 (U (S(z))) and is equivalent to T (z).
Though fatigued, our well-traveled point realizes there's a shortcut. Why
go through this complicated wash? We can understand T as follows: T will
push points along type I clines of p and q (according to the dilation factor r)
and along type II clines of p and q (according to the rotation factor θ ).
iθ
We emphasize two special cases of this normal form. If |re | = 1 there is
no dilation, and points simply get rotated about type II clines of p and q as
in Figure 3.5.5. Such a Möbius transformation is called an elliptic Möbius
transformation .
q
p
The second special case occurs when θ = 0. Here we have a dilation factor
r, but no rotation. All points move along type I clines of p and q, as in
Figure 3.5.6. A Möbius transformation of this variety is called a hyperbolic
Möbius transformation . A hyperbolic Möbius transformation xing p and
q either sends all points away from p and toward q or vice versa, depending on
the value of r.
q
p
z∞ − p
= λ · 1.
z∞ − q
Plug z=∞ into the normal form to see
w∞ − p
1=λ .
w∞ − q
Next solve each equation for λ, set them equal, cross multiply, and simplify
as follows to get the result:
z∞ − p w∞ − q
=
z∞ − q w∞ − p
(z∞ − p)(w∞ − p) = (w∞ − q)(z∞ − q)
p2 − pw∞ − pz∞ = q 2 − qw∞ − qz∞
p2 − q 2 = p(z∞ + w∞ ) − q(z∞ + w∞ )
(p − q)(p + q) = (p − q)(z∞ + w∞ )
p + q = z∞ + w∞ . (since p 6= q )
Solve this expression for T (z) and reduce it using the fact that p+q =
z∞ + w∞ to get the expression for T that appears in the statement of the
theorem. The details are left to the reader.
(6 + 3i)z + (2 − 3i)
T (z) = .
z+3
First we nd the xed points and the normal form of T. To nd
the xed points we solve T (z) = z for z.
(6 + 3i)z + (2 − 3i)
=z
z+3
(6 + 3i)z + (2 − 3i) = z 2 + 3z
z 2 − (3 + 3i)z − (2 − 3i) = 0.
Hey! Wait a moment! This looks familiar. Let's see ... yes! We
showed in Example 2.4.4 that this quadratic equation has solutions
z=i and z = 3 + 2i.
So the map has these two xed points, and the normal form of T is
T (z) − i z−i
=λ .
T (z) − (3 + 2i) z − (3 + 2i)
To nd the value of λ, plug into the normal form a convenient value
of z. For instance, T (−3) = ∞, so
−3 − i
1=λ .
−3 − (3 + 2i)
4i T (4i)
w∞
3 + 2i
z∞ 0 1 T (1)
T (0)
T (z) − i z−i
=λ .
T (z) − 0 z−0
60 3. Transformations
2−i
= λ(1 − i).
2
Solving for λ we have
3 1
λ= + i,
4 4
and the map is loxodromic.
√
Expressing λ in polar form, λ = reiθ , gives r = 410 and θ =
arctan(1/3). So T pushes points along type I clines of i and 0 according
to the scale factor r and along type II clines of i and 0 according to
the angle θ .
Now we consider Möbius transformations that x just one point. One such
Möbius transformation comes to mind immediately. For any complex number
d, the translation T (z) = z + d xes just ∞. In the exercises, you prove
that translations are the only Möbius transformations that x ∞ and no other
point.
1
Now suppose T xes p 6= ∞ S(z) = z−p
(and no other point). Let be
−1
a Möbius transformation taking p to ∞, and let U = S ◦ T ◦ S . Then
U (∞) = S(T (S −1 (∞))) = S(T (p)) = S(p) = ∞, and U xes no other point.
Thus, U (z) = z + d for some complex constant d.
The composition equation S ◦ T = U ◦ S gives the following equation called
the normal form of a Möbius transformation T xing p 6= ∞ (and no
other point):
1 1
= +d
T (z) − p z−p
Observe that U (z) = z + d pushes points along lines parallel to one another
in the direction of d (as in the right of Figure 3.5.12). All of these parallel
−1
lines meet at ∞ and are mutually tangent at this point. The map S takes
this system of clines to a system of clines that meet just at p, and are tangent
to one another at p, as pictured. The slope of the single line in this system
depends on the value of the constant d. In fact, the single line in the system
of clines is the line through p and T (∞) (see Exercise 3.5.12 for details).
S
d
p
0
S −1
Figure 3.5.12: A parabolic map xing p pushes points along clines that are
mutually tangent at p.
3.5. Möbius Transformations: A Closer Look 61
A map that xes just p will push points along such a system of clines that
are mutually tangent at p. Such a map is called parabolic . In a sense, a
parabolic map sends points both toward and away from p along these clines,
just as any translation pushes points along a line toward ∞ and also away from
∞.
Example 3.5.13: Normal form, one xed point..
Consider T (z) = (7z − 12)/(3z − 5). To nd its normal form we start
by nding its xed points.
z = T (z)
z(3z − 5) = 7z − 12
2
3z − 12z + 12 = 0
z 2 − 4z + 4 = 0
(z − 2)2 = 0
z = 2.
1 1
= + d.
T (z) − 2 z−2
1 1
= +d
0.4 −2
so that d = 3. The normal form is then
1 1
= + 3.
T (z) − 2 z−2
Exercises
1. Complete the proof of Theorem 3.5.8.
1 1
= + d.
T (z) − p z−p
1
Prove that the line through p and p+ d gets sent to itself by T.
13. Analyze T (z) = [(1 + 3i)z − 9i]/[iz + (1 − 3i)] by nding the xed
points, nding the normal form, and sketching the appropriate system of clines
indicating the motion of the transformation.
14. Find a parabolic transformation with xed point 2+i for which
T (∞) = 8.
15. Given distinct points p, q , and z in C, prove there exists a type II cline
of p and q that goes through z.
4
Geometry
Recall the two paragraphs from Section 1.2 that we intended to spend time
making sense of and working through:
63
64 4. Geometry
T = {Tb | b ∈ C}.
C
i
B E F
C
G H
L
A
For instance, suppose D is the set of all lines in C. Let f be the function
that takes a line to its slope. In translational geometry, (C, T ), the set D of
all lines is an invariant set because if A is any line, then so is its image, T (A),
under any translation T in T. Furthermore, f is an invariant function because
any translation of any line preserves the slope of that line.
Of course, two gures in an invariant set need not be congruent. For
instance, in translational geometry the set D of all lines is an invariant set,
although if lines A and B in D have dierent slopes then they are not congruent.
This feature of the set D makes it seem too big, in some sense. Can invariant
sets be more exclusive, containing only members that are congruent to one
another? You bet they can.
A = {T (A) | T ∈ G}.
Notice that for any T ∈ G, T (A) is in the set D since D is invariant. This
means that A is a subset of D.
Furthermore, A itself is an invariant set, thanks to the group nature of G.
In particular, if C is any member of A, then C = T0 (A) for some particular T0
in G. Thus, applying any transformation T to C ,
|z − r| ≤ |z| + r.
Notice that
Exercises
1. Find a particular translation to prove that in Figure 4.1.7 H ∼
= L in
translational geometry.
2. Let A be the set of all circles in C centered at the origin, and let G be
the set of all inversions about circles in A. That is,
G = {iC | C ∈ A}
11. Prove that (C, E) is isotropic. That is, show the group E contains all
rotations about all points in C.
12. Which gures from Figure 4.1.7 are congruent in (C, E)?
13. Let's create a brand new geometry, using the set of integers Z =
{. . . , −2, −1, 0, 1, 2, . . .}. For each integer n, we dene the transformation
Tn : C → C by Tn (z) = z + ni. Let G denote the set of all transformations Tn
for all integers n. That is, G = {Tn | n ∈ Z}.
a. Prove that (C, G) is a geometry.
b. Consider the set of gures D consisting of all lines in the plane with
slope 4. Is D an invariant set of (C, G)? Is it minimally invariant? Explain.
c. My favorite line, for clear and personal reasons, is y = x + 8. Please
describe a minimally invariant set of gures containing this line.
d. Determine the set of points in C congruent to i in this geometry. Is C
homogeneous?
While we're at it, let's restate three other facts about Möbius transforma-
tions:
Exercises
1. Which gures in Figure 4.2.2 are congruent in (C+ , M).
2. Describe a minimally invariant set in
+
(C , M) containing the triangle
comprised of the three vertices 0, 1, and i and the three Euclidean line segments
connecting them. Be as specic as possible about the members of this set.
4. Repeat the previous exercise for the set G consisting of all hyperbolic
Möbius transformations that x p and q.
4.2. Möbius Geometry 71
A B C D
E F G H
0
C
S1∞
Figure 5.1.2: Inversion about a cline orthogonal to the unit circle takes D to
D.
1 See Arthur Miller's chapter in [13] for a discussion of Poincaré's diverse interests.
72
5.1. The Poincaré Disk Model 73
C1
z0∗
0 z0
C
S1∞
times, once, or twice. In Figure 5.1.6 we illustrate these three cases. In each
case, we build a transformation T in H by inverting about the solid clines in
the gure (rst about L1 , then about L2 ). The gure also tracks the journey
of a point z under these inversions, rst to z0 by inverting about L1 , then
0
onto T (z) by inverting z about L2 . The dashed clines in the gure represent
some of the clines of motion, the clines along which points are moved by the
transformation. Notice these clines of motion are orthogonal to both clines of
inversion. Let's work through the three cases in some detail. If the two clines
L1 L1
0
z z
p z
0
z
p
T (z)
T (z)
L2
L2 (b)
(a) p (b)
L1
z
L2
z0
T (z)
q
(c)
of inversion, L1
L2 , intersect inside D, say at the point p, then they also
and
intersect outside D at the point p∗ symmetric to p with respect to the unit circle
1
since both clines are orthogonal to S∞ . This scene is shown in Figure 5.1.6(a).
5.1. The Poincaré Disk Model 75
q
T 2 (z)
T (z)
L1
T 2 (z) p
T (z)
L2
L2 z z L1
p
(a) (b)
Figure 5.1.7: Moving an `M' about the hyperbolic plane, by (a) rotation
about p; and (b) translation xing ideal points p and q.
z − z0
T (z) = eiθ
1 − z0z
where θ is some angle, and z0 is the point inside D that gets sent to 0.
z1 − z0∗ z − z0
T (z) = (z, z1 ; z0 , z0∗ ) = · .
z1 − z0 z − z0∗
But z0∗ = 1
z0 , so
z1 − 1/z0 z − z0 z0 z1 − 1 z − z0
T (z) = · = · .
z1 − z0 z − 1/z0 z1 − z0 z0 z − 1
Since |z1 | = 1 and, in general |β| = |β| we see that this expression
has modulus 1, and can be expressed as eiθ for some θ. Thus, if T is a
transformation in H it may be expressed as
z − z0
T (z) = eiθ
1 − z0z
where θ is some angle, and z0 is the point inside D that gets sent to 0.
Is the converse true? Is every transformation in the form above actually a
member of H? The answer is yes, and the reader is asked to work through the
details in the exercises.
Exercises
1. Prove that (D, H) is homogeneous.
5.2. Figures of Hyperbolic Geometry 77
z − z0
T (z) = eiθ ,
1 − z0z
where z0 is in D. Prove that T maps D to D by showing that if |z| < 1 then
|T (z)| < 1. Hint: It is easier to prove that |z|2 < 1 implies |T (z)|2 < 1.
3. Construct the xed points of the hyperbolic translation dened by the
inversion of two nonintersecting clines that intersect S1∞ at right angles, as
shown in the following diagram.
S1∞
Theorem 5.2.3. There exists a unique hyperbolic line through any two distinct
points in the hyperbolic plane.
78 5. Hyperbolic Geometry
Theorem 5.2.4. Any two hyperbolic lines are congruent in hyperbolic geom-
etry.
Proof. We rst show that any given hyperbolic line L is congruent to the
hyperbolic line on the real axis. Suppose p
L, and v is one of its
is a point on
ideal points. By Lemma 5.1.5 there is a transformation T in H that maps p to
∗
0, v to 1, and p to ∞. Thus T (L) is the portion of the real axis inside D, and
L is congruent to the hyperbolic line on the real axis. Since any hyperbolic
line is congruent to the hyperbolic line on the real axis, the group nature of H
ensures that any two hyperbolic lines are congruent.
L
p
1
0 T (L)
Figure 5.2.5: Any hyperbolic line is congruent to the hyperbolic line on the
real axis.
0
v
Figure 5.2.7: Through a point not on a given hyperbolic line L there exist
two hyperbolic lines parallel to L.
q
r
The next section develops a distance function for the hyperbolic plane. As
in Euclidean geometry, we want to be able to compute the distance between
two points, the length of a path, the area of a region, and so on. Moreover, the
distance function should be an invariant; the distance between points should
not change under a transformation in H. With this in mind, consider again a
hyperbolic rotation about a point p, as in Figure 5.1.6(a). It xes the point p
and moves points around type II clines of p and p∗ . If the distance between
points is unchanged under transformations in H, then all points on a given
type II cline of p and p∗ will be the same distance away from p. This leads us
to dene a hyperbolic circle as follows.
p p∗
Exercises
1. Suppose C is a hyperbolic circle centered at z0 through point p.
Show that there exists a hyperbolic line L tangent to C at p, and that L
is perpendicular to hyperbolic segment z0 p.
2. Constructing a hyperbolic line through two given points.
a. Given a point p in D, construct the point p∗ symmetric to p with respect
to the unit circle (see Figure 3.2.18).
b. Suppose q D. Construct the cline through p, q , and
is a second point in
p∗ . Call this cline C. C intersects the unit circle at right angles.
Explain why
c. Consider the portion of cline C you constructed in part (b) that lies in
D. This is the unique hyperbolic line through p and q . Mark the ideal points
of this hyperbolic line.
6. Explain why Theorem 5.2.6 applies in the general case, when z0 is not at
the origin.
7. Given a point and a hyperbolic line not passing through it, prove that
there is a hyperbolic line through the point that is perpendicular to the given
line. Is this perpendicular unique?
Perhaps the least obvious of the features listed is the last one. One theme
of this text is that locally, on small scales, non-Euclidean geometry behaves
much like Euclidean geometry. A small segment in the hyperbolic plane is
approximated to the rst order by a Euclidean segment. Small hyperbolic
triangles look like Euclidean triangles and hyperbolic angles correspond to
Euclidean angles; the hyperbolic distance formula will t with this theme.
To nd the distance function, start with a point's distance from the origin.
Given a point z in D, rotate about 0 so that z gets sent to the point x = |z|
on the positive real axis.
We may nd a hyperbolic line L about which x gets reected to the origin.
Such a hyperbolic line is constructed in the proof of Theorem 5.1.3. Recall, the
line L is on the circle centered at x∗ (the point symmetric to x with respect
1 1
to to S∞ ) that goes through the points at which S∞ intersects the circle with
∗
diameter 0x . Let x+h be a point near x on the positive real axis, and suppose
x + h gets inverted to the point w, as depicted in Figure 5.3.1. One can show
(in Exercise 5.3.1) that
−h
w= .
1 − x2 − hx
82 5. Hyperbolic Geometry
w 0 x x+h
Also, 0, x, and x + h are all on the same hyperbolic line (the real axis), so
assuming h>0
d(w)
d0 (x) = lim
h→0 h
d(w) |w|
= lim
h→0 |w| h
h
= lim 2 ·
h→0 (1 − x2 − hx)h
2
= .
1 − x2
5.3. Measurement in Hyperbolic Geometry 83
d(x) we integrate:
To get back to the distance function
Z Z
2 1 1
dx = + dx (partial fractions)
1 − x2 1−x 1+x
= − ln(1 − x) + ln(1 + x)
1+x
d(x) = ln ,
1−x
1 + |z|
dH (0, z) = ln .
1 − |z|
Notice that if z inches its way in D out toward the circle at innity (i.e.,
|z| → 1), the hyperbolic distance from 0 to z approaches ∞. This is a good
thing. Thinking of Euclid's postulates, this notion of distance satises one of
our fundamental requirements: One can produce a hyperbolic segment to any
nite length.
To arrive at a general distance formula dH (p, q), observe something curious.
The hyperbolic line through 0 and x has ideal points -1 and 1. Furthermore,
the expression (1 + x)/(1 − x) corresponds to the cross ratio of the points 0,
x, 1, and -1. In particular,
= dH (0, x)
= ln((0, x; 1, −1))
= ln((p, q; u, v)), (invariance of cross ratio)
where u and v are the ideal points of the hyperbolic line through p and q.
To be precise, u is the ideal point you would head toward as you went from p
to q, and v is the ideal point you would head toward as you went from q to p.
84 5. Hyperbolic Geometry
v q
p T
−1 0 x 1
Figure 5.3.2: To nd the distance between p and q, we may rst transform
p to 0 and q to the positive real axis.
z−p
T (z) = .
1 − pz
The map T sends q to some other point, T (q), in D. Assuming again that
T preserves distance, it follows that dH (p, q) = dH (0, T (q)), and
1 + |T (q)|
dH (p, q) = ln .
1 − |T (q)|
q−p
Making the substitution T (q) = provides us with the following
1 − pq
working formula for the hyperbolic distance between two points.
p q
z
1.49
4.64
w
r : [a, b] → C
r(t + ∆t)
r(t) ∆s
r(a)
r(b)
Z 2π
= | − a sin(t) + ia cos(t)| dt
0
Z 2π q
= a2 sin2 (t) + a2 cos2 (t) dt
0
Z 2π
= a dt
0
= 2πa.
In the hyperbolic plane, we may deduce the arc-length dierential by a
similar argument. Suppose r is a smooth curve in D given by r(t) = x(t)+iy(t),
for a ≤ t ≤ b. One may approximate the length of a tiny portion of the curve,
say from r(t) to r(t+∆t), by the hyperbolic distance between these two points,
dH (r(t), r(t + ∆t)). To compute this distance, we rst send the point r(t) to
0 by the transformation
z − r(t)
T (z) = ,
1 − r(t)z
so that
We are now in a position to argue that in the hyperbolic plane, the shortest
path (geodesic) connecting two points p and q is along the hyperbolic line
through them.
Theorem 5.3.9. Hyperbolic lines are geodesics; that is, the shortest path
between two points in (D, H) is along the hyperbolic segment between them.
Proof Sketch : We rst argue that the geodesic from 0 to a point c on the
positive real axis is the real axis itself.
Suppose r(t) = x(t) + iy(t) for a ≤ t ≤ b, is an arbitrary smooth curve
from 0 to c (so r(a) = 0 and r(b) = c).
Suppose further that x(t) is nondecreasing (if our path backtracks in the x
direction, we claim the path cannot possibly be a geodesic). Then
Z b
2 p
L(r) = (x0 (t))2 + (y 0 (t))2 dt.
a 1 − [(x(t))2 + (y(t))2 ]
The hyperbolic line segment from 0 to c can be parameterized by r 0 (t) =
x(t) + 0i for a ≤ t ≤ b, which has length
Z b
2 p
L(r 0 ) = 2
(x0 (t))2 dt.
a 1 − [x(t)]
The curve r0 is essentially the shadow of r on the real axis.
One can compare the integrands directly to see that L(r) ≥ L(r 0 ).
Since transformations in H preserve arc-length and hyperbolic lines, it
follows that the shortest path between any two points in D is along the
hyperbolic line through them.
r
p q
L q
Lpq
q
Lqr p
Lpr r
Exercises
1. Suppose 0 < x < 1 and L is a hyperbolic line about which x gets inverted
to the origin. (Such an inversion was constructed in Theorem 5.1.3.) For a
real number h, let w be the image of x+h under this inversion. Prove that
−h
w= 1−x2 −hx .
2. Determine a point in D whose hyperbolic distance from the origin is
2,003,007.4 units.
4. Determine the hyperbolic distance from the point p = 0.5 to the point
q = 0.25 + 0.5i.
5. Prove Theorem 5.3.7.
90 5. Hyperbolic Geometry
0 a x
ea − 1
x= .
ea + 1
This circular region may be described in polar coordinates by
0 ≤ θ ≤ 2π and 0 ≤ r ≤ x. The area of the region is then given
by the following integral, which we compute with the u-substitution
u = 1 − r2 :
Z 2π Z x
4r
drdθ
0 0 (1 − r2 )2
5.4. Area and Triangle Trigonometry 91
Z 2π Z 1−x2
−2
= dθ du
0 1 u2
1−x2
2
= 2π
u
1
2
= 2π − 2
1 − x2
x2
= 4π .
1 − x2
Replace x in terms of a to obtain
This last expression can be rewritten using the hyperbolic sine function,
evaluated at a/2. We investigate the hyperbolic sine and cosine functions in
the exercises but note their denitions here.
Denition 5.4.3. The hyperbolic sine function , denoted sinh(x), and the
hyperbolic cosine function , denoted cosh(x), are functions of real numbers
dened by
ex − e−x ex + e−x
sinh(x) = and cosh(x) = .
2 2
The area derivation in Example 5.4.2 may then be summarized as follows.
Theorem 5.4.5. The area of a 32 -ideal triangle having interior angle α is equal
to π − α.
The proof of this theorem is given in the following section. The proof there
makes use of a dierent model for hyperbolic geometry, the so-called upper
half-plane model.
92 5. Hyperbolic Geometry
α
p
Figure 5.4.6: A
2
3 -ideal triangle having interior angle α has area equal to
π − α.
An ideal triangle consists of three ideal points and the three hyperbolic
lines connecting them. It turns out that all ideal triangles are congruent (a
fact proved in the exercises); the set of all ideal triangles is minimally invariant
in (D, H).
−1 1
It is a remarkable fact that π is an upper bound for the area of any triangle
in (D, H). No triangle in (D, H) can have area as large as π , even though side
lengths can be arbitrarily large!
q
R1
β
p α R3
γ
R2
r
t
b
γ α
a
c
β
α w
z c
In this setting, ∆zwu is a 31 -ideal triangle, and the second hyperbolic law
of cosines applies with γ = 0 and β = π/2 to yield the following result.
p
u
z w
L
q
a q
xi
a p
a
0 a x
Figure 5.4.16: A journey that would trace a square in the Euclidean plane
does not get you home in (D, H).
96 5. Hyperbolic Geometry
L2 L1
z
C1
−1 1
−a a
C2
z
While squares don't exist in the hyperbolic plane, we may build right-angled
regular polygons with more than four sides using hyperbolic line segments. In
5.4. Area and Triangle Trigonometry 97
fact, for each triple of positive real numbers (a, b, c) we may build a right-angled
hexagon in the hyperbolic plane with alternate side lengths a, b, and c. We
encourage the reader to work carefully through the construction of this hexagon
in the proof of Theorem 5.4.19. We use all our hyperbolic constructions to get
there.
Theorem 5.4.19. For any triple (a, b, c) of positive real numbers there exists a
right-angled hexagon in (D, H) with alternate side lengths a, b, and c. Moreover,
all right-angled hexagons with alternate side lengths a, b, and c are congruent.
Proof. We prove the existence of a right-angled hexagon with vertices
v0 , v1 , · · · , v5 such that dH (v0 , v1 ) = a, dH (v2 , v3 ) = b, and dH (v4 , v5 ) = c.
C
v4
v5
D c v3
b
v2
B
a
v0 v1
Figure 5.4.20: Building a right hexagon in the hyperbolic plane that has
alternate side lengths (a, b, c).
First, let v0 be the origin in the hyperbolic plane, and place v1 on the
positive real axis so that dH (v0 , v1 ) = a. Note that
ea − 1
v1 = .
ea + 1
Next, construct the hyperbolic line A perpendicular to the real axis at the
point v1 . This line is part of the cline that has diameter v1 v1∗ .
Pick any point v2 on the line A. For the sake of argument, assume that v2
lies above the real axis, as in Figure 5.4.20.
Next, construct the hyperbolic line B perpendicular to A at the point v2 .
This line is part of the cline through v2 and v2∗ with center on the line tangent
to A at v2 .
Next, construct the point v3 on line B that is a distance b away from v2 .
This point is found by intersecting B with the hyperbolic circle centered at v2
with radius b. (To construct this circle, we rst nd the scalar k so that the
hyperbolic distance between kv2 and v2 is b.)
Next, draw the perpendicular C to line B at v3 .
Then construct the common perpendicular of C and the imaginary axis,
call this perpendicular D. We construct this common perpendicular as follows.
First nd the two points p and q symmetric to both C and the imaginary axis.
∗ 1
Then nd p , the point symmetric to p with respect to S∞ . The cline through
∗
p, q , and p is perpendicular to C , the imaginary axis and S1∞ , so this gives us
our line D .
If C and the imaginary axis intersect, no such perpendicular exists (think
triangle angles), so drag v2 toward v1 until these lines do not intersect. Then
98 5. Hyperbolic Geometry
−1 + 2i
C
i
−1 + i 1+i
p q
L1 c L2
C1 −1 C2
1
0
1 + |q|
dH (0, q) = ln
1 − |q|
5.4. Area and Triangle Trigonometry 99
√1
1 +
5
= ln
1− √1
5
√
5+1
= ln √
5−1
√
( 5 + 1)2
= ln √ √
( 5 − 1)( 5 + 1)
√
( 5 + 1)2
= ln
4
= ln(ϕ2 )
= 2 ln(ϕ).
Exercises
1. Properties of sinh(x) = (ex − e−x )/2 and cosh(x) = (ex + e−x )/2.
a. Verify that sinh(0) = 0 and cosh(0) = 1.
d d
b. Verify that
dx [sinh(x)] = cosh(x) and dx [cosh(x)] = sinh(x).
2 2
c. Verify that cosh (x) − sinh (x) = 1.
d. Verify that the power series expansions for cosh(x) and sinh(x) are
x2 x4
cosh(x) = 1 + + + ···
2 4!
x3 x5
sinh(x) = x + + + ··· .
3! 5!
2. Prove that the circumference of a hyperbolic circle having hyperbolic
radius r is C = 2π sinh(r).
3. The hyperbolic plane looks Euclidean on small scales.
a. Prove
4π sinh2 (r/2)
lim+ = 1.
r→0 πr2
Thus, for small r, the Euclidean formula for the area of a circle is a good
approximation to the true area of a circle in the hyperbolic plane.
b. Prove
2π sinh(r)
lim+ = 1.
r→0 2πr
Thus, for small r, the Euclidean formula for the circumference of a circle
is a good approximation to the true circumference of a circle in the hyperbolic
plane.
4. Prove that all ideal triangles are congruent in hyperbolic geometry. Hint:
Prove any ideal triangle is congruent to the one whose ideal points are 1, i,
and -1 (see Figure 5.4.7).
6. 1 1 1
2 , and 2 + 2 i.
Consider the hyperbolic triangle with vertices at 0,
Calculate the area of this triangle by determining the angle at each vertex.
Hint: To determine the angle at a vertex it may be convenient to move it to
the origin via an appropriate transformation in H.
100 5. Hyperbolic Geometry
7. Recall the block constructed in Example 5.4.17. Prove that all four angles
are 90◦ , and that for any choice of z, opposite sides have equal hyperbolic
length.
1+r 2
|b − v0 | 2r −r
tan(π/8) = = 1+r 2
.
|z0 − b| 2r · tan(π/8)
d. Solve the equation in (c) for r to obtain r = (1/2)(1/4) .
v2
v1
v3 z0
v4
v0 b vo∗
v5 v7
v6
11. Prove the rst hyperbolic law of cosines by completing the following
steps.
a. Show that for any positive real numbers x and y,
2 2
x +y x2 − y 2
cosh(ln(x/y)) = and sinh(ln(x/y)) = .
2xy 2xy
b. Given two points p and q in D, let c = dH (p, q). Use the hyperbolic
distance formula from Theorem 5.3.3 and part (a) to show
c. Now suppose our triangle has one vertex at the origin, and one point on
the positive real axis. In particular, suppose p = r (0 < r < 1) and q = keiγ
(0 < k < 1), with angles α, β, γ and hyperbolic side lengths a, b, and c as in
Figure 5.4.23.
α c
b
γ
β
0 a p 1
Figure 5.4.23: A hyperbolic triangle with one corner at the origin and one
leg on the positive real axis.
Show
1 + r2 2r
cosh(a) = ; sinh(a) = ;
1 − r2 1 − r2
1 + k2 2k
cosh(b) = ; sinh(b) = .
1 − k2 1 − k2
d. Show that for the triangle in part (c),
q∗
mq
α z0
q
R
β
α
b R
c
γ β
0 a p mp p∗
13. Prove the second law of hyperbolic cosines. Hint: This result follows
from repeated applications of the rst law and judicial use of the two identities
cos2 (x) + sin2 (x) = 1 and cosh2 (x) − sinh2 (x) = 1.
14. Recall the journey in Example 5.4.17 in which a bug travels a path that
would trace a square in the Euclidean plane. For convenience, we assume the
starting point p is such that the rst right turn of 90◦ occurs at the point x on
the positive real axis and the second turn occurs at the origin. (This means
that x = (ea − 1)/(ea + 1).) The third corner must then occur at xi. We
have reproduced the journey with some more detail in Figure 5.4.25. In this
exercise we make use of hyperbolic triangle trig to measure some features of
this journey from p to q in terms of the length a of each leg.
a. Determine the hyperbolic distance between p and the origin (corner two
of the journey). In particular, show that dH (0, p) = cosh2 (a). Note that if
this journey had been done in the Euclidean plane, the corresponding distance
√ √
would be 2a. Is cosh2 (a) close to 2a for small positive values of a?
1
b. Let θ = ∠x0p. Show that tan(θ) =
cosh(a) . What is the corresponding
angle if this journey is done in the Euclidean plane? What does θ approach as
a → 0+ ?
c. Show that ∠x0p = ∠0px.
4
d. Show that dH (p, q) = cosh (a)[1 − sin(2θ)] + sin(2θ).
e. Let α = ∠qp0, b = dH (0, p) and c = dH (p, q) Show that
sinh(b)
sin(α) = cos(2θ).
sinh(c)
f. Determine the area of the pentagon enclosed by the journey if, after
reaching q we return to p
along the geodesic. In particular, show that the area
3π
of this pentagon equals
2 − 2(θ + α). What is the corresponding area if the
journey had been done in the Euclidean plane?
g. Would any of these measurements change if we began at a dierent point
in the hyperbolic plane and/or headed o in a dierent direction initially than
the ones in Figure 5.4.25?
5.5. The Upper Half-Plane Model 103
a q
xi
c
a α p
b
a
0
θ
a x
Figure 5.4.25: A journey that would trace a square in the Euclidean plane
does not get you home in (D, H). But how close will the nish point q be to
the starting point p?
The Poincaré disk model is one way to represent hyperbolic geometry, and for
most purposes it serves us very well. However, another model, called the upper
half-plane model, makes some computations easier, including the calculation
of the area of a triangle.
C
i
−1 1
Figure 5.5.2: Inversion in C maps the unit disk to the upper-half plane.
Notice that inversion about the circle C xes -1 and 1, and it takes i to
∞. Since reection across the real axis leaves these image points xed, the
composition of the two inversions is a Möbius transformation that takes the
unit circle to the real axis. The map also sends the interior of the disk into
104 5. Hyperbolic Geometry
the upper half plane. Notice further that the Möbius transformation takes ∞
to −i; therefore, by Theorem 3.5.8, the map can be written as
−iz + 1
V (z) = .
z−i
This Möbius transformation is the key to transferring the disk model of
the hyperbolic plane to the upper half-plane model. In fact, when treading
back and forth between these models it is convenient to adopt the following
convention for this section: Let z denote a point in D, and w denote a point
in the upper half-plane U, as in Figure 5.5.3. We record the transformations
linking the spaces below.
−iz + 1 iw + 1
w = V (z) = and z = V −1 (w) = .
z−i w+i
V
w
z
real axis
S1∞
V −1
Dene the hyperbolic distance between two points w1 , w2 in the upper half-
plane model, denoted dU (w1 , w2 ), to be the hyperbolic distance between their
pre-images in the disk model.
Suppose w1 and w2 are two points in V whose pre-images in the unit disk
are z1 and z2 , respectively. Then,
where u and v are the ideal points of the hyperbolic line through z1 and z2 .
But, since the cross ratio is preserved under Möbius transformations,
where p, q are the ideal points of the hyperbolic line in the upper half-plane
through w1 and w2 . In particular, going from w1 to w2 we're heading toward
ideal point p.
5.5. The Upper Half-Plane Model 105
We now derive the hyperbolic arc-length dierential for the upper half-
plane model working once again through the disk model. Recall the arc-length
dierential in the disk model is
2|dz|
ds = .
1 − |z|2
2|dz|
ds =
1 − |z|2
iw+1
2|d w+i |
iw+1
= (z = w+i )
iw+1 2
1 − w+i
|iw + 1|2
2|i(w + i)dw − (iw + 1)dw|
= 1− (chain rule)
|w + i|2 |w + i|2
4|dw|
=
|w + i| − |iw + 1|2
2
106 5. Hyperbolic Geometry
4|dw|
=
(w + i)(w − i) − (iw + 1)(−iw + 1)
4|dw|
=
2i(w − w)
|dw|
= .
Im(w)
This leads us to the following denition:
b
|r 0 (t)|
Z
L(r) = dt
a Im(r(t))
b
b−a
Z
1
L(r) = dt = .
a k k
dw
From the arc-length dierential ds = Im(w) comes the area dierential:
Denition 5.5.8. In the upper half-plane model (U, U) of hyperbolic geome-
try, the area of a region R described in cartesian coordinates, denoted A(R),
is given by ZZ
1
A(R) = dxdy.
R y2
α
w
π−α
Z 1
1
= √ dx.
cos(π−α) 1 − x2
√
With the trig substituion cos(θ) = x, so that 1 − x2 = sin(θ) and
− sin(θ)dθ = dx, the integral becomes
0
− sin(θ)
Z
= dθ
π−α sin(θ)
= π − α.
2
It turns out that any
3 -ideal triangle is congruent to one of the form 1w∞
where w is on the upper half of the unit circle (Exercise 5.5.3), and since our
transformations preserve angles and area, we have proved the area formula for
2
a
3 -ideal triangle.
Exercises
1. What becomes of horocycles when we transfer the disk model of
hyperbolic geometry to the upper half-plane model?
2. What do hyperbolic rotations in the disk model look like over in the
upper half-plane model? What about hyperbolic translations?
4. Determine the area of the triangular region pictured below. What is the
image of this triangle under V −1 in the disk model of hyperbolic geometry?
Why doesn't this result contradict Theorem 5.4.9?
1 + 2i
i 1+i
5. Another type of block. Consider the four-sided gure pqst in (D, H) shown
in the following diagram. This gure is determined by two horocycles C1 and
C2 , and two hyperbolic lines L1 and L2 all sharing the same ideal point. Note
that the lines are orthogonal to the horocycles, so that each angle in the four-
◦
sided gure is 90 .
a. By rotation about the origin if necessary, assume the common ideal point
is i and use the map V to transfer the gure to the upper half-plane. What
does the transferred gure look like in U? Answer parts (b)-(d) by using this
transferred version of the gure.
pq and st are equal.
b. Prove that the hyperbolic lengths of sides
c equal the hyperbolic length of the leg pt along the larger radius
c. Let
horocycle C1 , and let d equal the hyperbolic length of the leg sq on C2 . Show
x
that c = e d where x is the common length found in part (b).
d. Prove that the area of the four-sided gure is c − d.
108 5. Hyperbolic Geometry
6
Elliptic Geometry
Recall, the unit 2-sphere S2 consists of all points (a, b, c) in R3 for which
2 2 2 2
a + b + c = 1, and S may be mapped onto the extended plane by the
stereographic projection map φ : S2 → C+ dened by
(
a
+ b i if c 6= 1;
φ(a, b, c) = 1−c 1−c
∞ if c = 1.
−1/z
if z 6= 0, ∞;
za = ∞ if z = 0;
z = ∞.
0 if
109
110 6. Elliptic Geometry
Lemma 6.1.2. Given two diametrically opposed points on the unit sphere,
their image points under stereographic projection are antipodal points in C+ .
Proof. First note that the north pole N = (0, 0, 1) and the south pole
S = (0, 0, −1) φ to ∞
are diametrically opposed points and they get sent by
C+ ; so the lemma holds in this case.
and 0, respectively, in
Now suppose P = (a, b, c) is a point on the sphere with |c| 6= 1, and
Q = (−a, −b, −c) is diametrically opposed to P . The images of these two
points under stereographic projection are
a b −a b
φ(P ) = + i and φ(Q) = − i.
1−c 1−c 1+c 1+c
If we expand the following product
a b −a b
φ(P ) · φ(Q) = + i · + i ,
1−c 1−c 1+c 1+c
we obtain
a2 + b2
φ(P ) · φ(Q) = −
1 − c2
which reduces to −1 since a2 + b2 + c2 = 1.
Thus, diametrically opposed points on the sphere get mapped via stereo-
graphic projection to antipodal points in the extended plane.
za
z 0 1
wa
pa
0
p
q
a b a b
φ(P ) = + i and φ(P ∗ ) = + i
1−c 1−c 1+c 1+c
112 6. Elliptic Geometry
are on the same ray beginning at the origin. Indeed, one is the positive
scalar multiple of the other:
1−c
φ(P ∗ ) = φ(P ).
1+c
Moreover,
1−c
|φ(P )| · |φ(P ∗ )| = · |φ(P )|2
1+c
1 − c a2 + b2
= ·
1 + c (1 − c)2
a2 + b2
=
1 − c2
= 1.
iS1 ◦ φ = φ ◦ R.
We end the section with one more feature of the stereographic projection
map. The proof can be found in [10].
(u2 + v 2 ) + 4u + 1 = 0.
(u + 2)2 + v 2 = 3
√
having center (−2, 0) and radius 3.
−2 · 3x − 2 · 2y + (1 − 13 + 4)z + (1 + 13 − 4) = 0
or
−3x − 2y − 4z + 5 = 0.
Exercises
1. Constructing an antipodal point. Suppose z is a point inside the unit
circle. Prove that the following construction, which is depicted in Figure 6.1.11,
gives za , the point antipodal to z: (1) Draw the line through z and the origin;
(2) draw the line through the origin perpendicular to line (1), and let T be on
line (2) and the unit circle; (3) construct the segment zT ; (4) construct the
perpendicular to segment (3) at point T. Line (4) intersects line (1) at the
point za . Use similar triangles to prove that za = − |z|1 2 z.
(2)
(1)
0
(3)
(4)
za
T
Figure 6.1.11: Constructing the antipodal point to z.
5. Determine the image under φ of the circle z = 1/2 on the unit sphere.
n
p
s r
m
o
b. Prove the intersecting chords theorem: If mn and ab are any two chords
of C passing through a given interior point p, then
|m − p| · |a − p| = |m − p| · |b − p|.
e−iθ
R(z) · R(za ) = −eiθ z · = −1.
z
Thus, the image points R(z) and R(za ) are still antipodal points.
Rotations about the origin belong to the group S.
−1
T (z) = . (1)
T (−1/z̄)
6.2. Elliptic Geometry 115
−a/z̄ + b
T (−1/z̄) =
−c/z̄ + d
−ā/z + b̄
=
−c̄/z + d¯
b̄z − ā
= ¯ .
dz − c̄
Substituting into equation (1) of this derivation yields
az + b ¯ + c̄
−dz
= .
cz + d b̄z − ā
The transformation on the right also has determinant one, so these two
transformations are identical up to sign. We can assume that d = −ā and that
c = b̄, so that T may be expressed as follows:
az + b
T (z) = .
b̄z − ā
We can make this general form look a lot like the general form for a
transformation in H. To do so, rst multiply each term by −1/ā, assuming
a 6= 0. (If a = 0, what would the transformation look like?)
− āa z − b
ā
T (z) =
− āb̄ z + 1
− āa (z + ab )
=
− āb̄ z + 1
z − z0
= eiθ · ,
z0z + 1
where eiθ = −a/ā and z0 = −b/a. Thus, we have derived the following
algebraic description of transformations in S:
Transformations in S .
Any transformation T in the group S has the form
z − z0
T (z) = eiθ
1 + z0z
One can show that the following converse holds: Any transformation having
the form above preserves antipodal points. It follows that for any point
z0 ∈ C+ , there exists a transformation T in S such that T (z0 ) = 0, and
since rotations about the origin live in S we can prove the following useful
result (see Exercise 6.2.1).
ppa − λppa
T (0) = .
pa − λp
On the other hand, setting z=∞ and solving for T (∞), one checks that
p − λpa
T (∞) = .
1−λ
If we expand this expression and solve it for λλ (using the fact that
p · pa = −1, and p 6= pa ), we obtain λλ = 1, from which it follows that
|λ| = 1. Thus T is an elliptic Möbius transformation.
In Exercise 3.5.8 we showed that any elliptic Möbius transformation is the
composition of two inversions about clines that intersect at the two xed points.
In the case of a Möbius transformation that preserves antipodal points, these
two xed points must be antipodal to each other. It follows by Lemma 6.1.4
that the two clines of inversion are great circles in C+ . Thus, we may view each
transformation of S as the composition of two inversions about great circles.
This is reassuring. By Theorem 6.1.6, these inversions correspond to reections
of the sphere about great circles, and composing two of these reections of
the sphere yields a rotation of the sphere. The transformation group S,
then, consists of Möbius transformations that correspond via stereographic
projection to rotations of the sphere. We summarize these facts below.
We can think of this space as the closed unit disk with its two edges (top-half
circle and bottom-half circle) identied according to the arrows in Figure 6.2.7.
Notice the pleasant journey a bug has taken from p to q in this gure. From
p she heads o toward point c, which appears on the boundary of our model.
When she arrives there she simply keeps walking, though in our model we see
118 6. Elliptic Geometry
her leave the screen and reappear at the antipodal point ca . She has her
sights set on point d and saunters down there, continues on (reappearing at
da ), and heads on to q, hungry but content.
da
ca
q
p
c
d
Denition 6.2.8. The disk model for elliptic geometry , (P2 , S), is the
geometry whose space is P2 and whose group of transformations S consists of
all Möbius transformations that preserve antipodal points.
a
c
b d
Theorem 6.2.11. There is a unique elliptic line connecting two points p and
q in P2 .
Proof. Suppose p and q are distinct points in P2 . This means q 6= pa
+
as points in C . Construct the antipodal point pa , which gives us three
distinct points in C+ : p, q and pa . There exists a unique cline through these
three points. Since this cline goes through p and pa , it is an elliptic line by
Lemma 6.1.4.
Note that elliptic lines through the origin are Euclidean lines, just as was the
case in the Poincaré model of hyperbolic geometry. As a result, to prove facts
about elliptic geometry, it can be convenient to transform a general picture to
the special case where the origin is involved.
6.2. Elliptic Geometry 119
Exercises
1. The transformation group in elliptic geometry.
a. Prove that S is a group of transformations.
b. For each θ ∈ R and z0 ∈ C, prove that the following Möbius
transformation is in S :
z − z0
T (z) = eiθ .
1 + z0 z
c. For each θ ∈ R, prove that T (z) = eiθ z1 is in S.
d. Use (b) and (c) to prove that for any distinct points p, q ∈ C+ there
exists a transformation in S that sends p to 0 and q to a point on the positive
real axis, thus proving Lemma 6.2.3.
Rather than derive the arc-length formula here as we did for hyperbolic
geometry, we state the following denition and note the single sign dierence
from the hyperbolic case. This sign dierence is consistent with the sign
dierence in the algebraic descriptions of the transformations in the respective
geometries.
ZZ
4r
A(R) = drdθ.
R (1 + r2 )2
(1)
p
(2)
Figure 6.3.3: There is a single elliptic line joining points p and q, but two
elliptic line segments. The distance from p to q is the shorter of these two
segments.
Proof. We rst determine the elliptic distance between the origin and a
point x (with 0 < x ≤ 1) on the positive real axis.
The elliptic line through 0 and x lives on the real axis, and we may
parameterize the eastbound segment connecting 0 to x by r(t) = t for
0 ≤ t ≤ x. (The westbound segment from 0 to x is clearly not shorter
than the eastbound segment.) The length of this segment is
x x
2|r 0 (t)|
Z Z
2|1|
dt = dt
0 1 + |r(t)|2 0 1 + |t|
2
Z x
1
=2 dt
0 1 + t2
= 2 arctan(x).
z−p
T (z) = eiθ
1 + pz
will do the trick. Now, T is a Möbius transformation of the entire extended
plane, and it may take q outside the unit disk, in which case qa will be
mapped to a point on the real axis inside the unit disk. So, either |T (q)|
or |T (qa )| will be a real number between 0 and 1. If we call this number x,
then dS (p, q) = dS (0, x), since transformations in S preserve distance between
points.
Now,
q−p
|T (q)| = ,
1 + pq
and the reader can check, using the fact that qa = −1/q , that
1 + pq
|T (qa )| =
.
q−p
It follows that
q−p 1 + pq
dS (p, q) = min 2 arctan
, 2 arctan
.
1 + pq q−p
This completes the proof.
With the sphere as our model, we can check our formulas against
measurements on the sphere. For instance, dS (0, 1) = 2 arctan(1) = π/2. The
elliptic segment from 0 to 1 corresponds via stereographic projection to one-
quarter of a great circle on the unit sphere. Any great circle on the unit sphere
has circumference 2π , π/2 on the
so one-quarter of a great circle has length
sphere. We also note that the distance formula dS (0, x) = 2 arctan(x) applies
+
to spherical geometry (C , S) for all positive real numbers x, and this distance
matches the corresponding distances of the points on the unit 2-sphere, see
Exercise 6.3.12.
We emphasize that π/2 is an upper bound for the distance between two
points in (P2 , S). However, there is no upper bound on how long a journey
along an elliptic line can be. If Bormit the bug wants to head out from point
p and travel r units along any line, Bormit can do it, without obstruction, for
any r > 0. Of course, if r is large enough, Bormit will do laps on this journey.
We say a path is a geodesic path if it follows along an elliptic line.
122 6. Elliptic Geometry
Denition 6.3.5. In(P2 , S), the elliptic circle centered at z0 with radius r
consists of all points z ∈ P2 such that there exists a geodesic path of length r
from z0 to z.
Each transformation T in S is an elliptic Möbius transformation by
Theorem 6.2.4 that xes two antipodal points, say p and pa . So T pushes
points along type II clines of p and pa , and since transformations preserve
distance between points, these type II clines of p and pa determine elliptic
circles; all points on these type II clines are equidistant from p.
Now, suppose p and q are any distinct points in P2 . There exists a type II
cline of p and pa that goes through q. If this cline lives entirely inside the closed
unit disk it represents the elliptic circle centered at p through q. Of course,
this cline may not live entirely inside the disk, as is the case in Figure 6.3.6.
But each point on the type II cline of p and pa through q has antipodal point
on the type II cline of p and pa through qa . So, in P2 we may represent the
elliptic circle centered at p through q by the portions of these two type II clines
of p and pa that live in the closed unit disk.
q
p
pa
qa
The area of a triangle We now turn our attention to nding a formula for
the area of a triangle in elliptic geometry. We begin with lunes. A lune is
the region in P2 bounded between two elliptic lines. How do two elliptic lines
bound a region? Two lines trap a region because we are identifying antipodal
points on the unit circle. A bug living in the shaded region of P2 pictured
in Figure 6.3.8 would be able to visit all shaded points without crossing the
boundary walls determined by the two elliptic lines. The shaded region is a
single, connected region bounded by two lines. So what is the area of this
region?
6.3. Measurement in Elliptic Geometry 123
Lemma 6.3.9 The area of a lune. In (P2 , S), the area of a lune with angle
α is 2α.
Proof. To compute the area of a lune, rst move the vertex of the lune
to the origin in such a way that one leg of the lune lies on the real axis,
as in Figure 6.3.10. Then half of the lunar region can be described in polar
coordinates by 0≤r≤1 and 0 ≤ θ ≤ α.
Z α Z 1
4r
A=2 drdθ (Let u = 1 + r2 )
0 0 (1 + r2 )2
Z α 1
−2
=2 dθ
0 1 + r 2 0
Z α
=2 1 dθ
0
= 2α.
p
α
β q
r γ
Theorem 6.3.12. In elliptic geometry (P2 , S), the area of a triangle with
angles α, β, γ is
A = (α + β + γ) − π.
From this theorem it follows that the angles of any triangle in elliptic
◦
geometry sum to more than 180 .
We close this section with a discussion of trigonometry in elliptic geometry.
We derive formulas analogous to those in Theorem 5.4.12 for hyperbolic
triangles. We assume here that the triangle determined by distinct points p, q
and z in (P2 , S) is formed by considering the shortest paths connecting these
three points. So triangle side lenghths will not exceed π/2 in what follows.
q
α c
b
γ β
r
0 a
Figure 6.3.13: An elliptic triangle with side lengths and angles marked.
Proof. a. Position our triangle conveniently, with one corner at the origin,
one on the positive real axis at the point r (0 < r ≤ 1), and one at the point
q = keiγ (with 0 < k ≤ 1) as in Figure 6.3.13. Then a = dS (0, r) = 2 arctan(r),
so that
by the cosine double angle formula. If we set θ = arctan(r) we may use the
following right triangle to rewrite the above description of cos(a) as follows:
1 r2 2
r
cos(a) = − √ 1+
1 + r2 1 + r2 r
2
1−r
= .
1 + r2 θ
1
1 − k2 2k
cos(b) = and sin(b) = .
1 + k2 1 + k2
And the third side? Theorem 6.3.4 tells us
iγ
iγ
ke − r
c = dS (r, ke ) = 2 arctan
.
1 + rkeiγ
1 + r2 k 2 − k 2 − r2 + 2rk(eiγ + e−iγ )
cos(c) = .
1 + r2 + k2 + r2 k2
q
α
R
1
mr 0
γ β
-
r
α r
β R
p mq
qa
Let mr be the midpoint of the segment connecting r and −1/r, and let mq
be the midpoint of the segment connecting q and qa . So,
1 + r2
1 1
|mr − r| = r+ = ,
2 r 2r
and
1 iγ 1 + k 2
1
|mq − q| = k +
e = .
2 k 2k
∆pmr r is right, and ∠mr rp = π/2−β so that ∠rpmr = β .
Note further that
From this right triangle we seesin(β) = (1 + r2 )/(2rR).
Similarly, ∆pmq q is right, and we have ∠mq pq = α, and sin(α) =
(1 + k 2 )/(2kR).
Comparing ratios,
sin(a) 2r 2kR
= ·
sin(α) 1 + r2 1 + k2
2k 2rR
= ·
1 + k 1 + r2
2
sin(b)
= .
sin(β)
To see that the ratio sin(c)/ sin(γ) must match the preceding common ratio,
note that transformations in elliptic geometry preserve distances and angles,
so we may transform our triangle above to one in which the length c is now
on the real axis with one end at the origin. The argument above ensures that
the ratio sin(c)/ sin(γ) then matches one of the other ratios, and so all three
agree.
6.3. Measurement in Elliptic Geometry 127
Exercises
1. Prove Theorem 6.3.7. Namely, prove that the circumference of a circle in
elliptic geometry is C = 2π sin(r), where r < π/2 is the elliptic radius. Hint:
Assume your circle is centered at the origin.
2. Prove
2π sin(r)
lim = 1.
r→0+ 2πr
Thus, for small r, the Euclidean formula for the circumference of a circle is
a good approximation to the true circumference of a circle in elliptic geometry.
4. Let C be an elliptic circle with center z0 and elliptic radius r > 0. For
what value(s) of r is C an elliptic line? For what value(s) of r is C a single
point?
A = 4π sin2 (r/2).
7. Prove
4π sin2 (r/2)
lim = 1.
r→0+ πr2
Thus, for small r, the Euclidean formula for the area of a circle is a good
approximation to the true area of a circle in elliptic geometry.
9. Prove that the area of P2 is 2π . Thus, unlike the hyperbolic case, the
space in elliptic geometry has nite area.
10. An intrepid tax collector lives in a country in the elliptic space P2 . For
collection purposes, the country is divided into triangular grids. The collector
observes that the angles of the triangle she collects in are 92◦ , 62◦ , and 27◦ .
What is the area of her triangle? (Be sure to convert the angles to radians.)
Can the entire space P2 be subdivided into a nite number of triangles?
11. Find a formula for the area of an n-gon in elliptic geometry (P2 , S),
given that its n angles are α1 , α2 , · · · , αn .
12. In this exercise we show the distance between any two points p and q in
spherical geometry (C+ , S) is
q−p
dS (p, q) = 2 arctan ,
1 + pq
128 6. Elliptic Geometry
and that dS (p, q) corresponds to the distance on the sphere between φ−1 (p)
−1
and φ (q).
a. The denition of arc-length in spherical geometry (C+ , S) is the same
2
as the one for (P , S). Using this denition, follow the proof of Theorem 6.3.4
to show that for any positive real number C+ , dS (0, x) = 2 arctan(x).
x in
+
b. Use invariance of arc-length to explain why, for arbitrary p and q in C ,
q−p
dS (p, q) = 2 arctan .
1 + pq
c. Suppose x>0 is a real number. Determine φ−1 (x), the point on the
unit sphere corresponding to x via stereographic projection. What is φ−1 (0)?
−1 −1
d. Determine the distance between φ (0) and φ (x) on the sphere. In
particular, show that this distance equals arccos((1 − x2 )/(1 + x2 )). Hint: the
distance between these points will equal the angle between the vectors to these
points, and this angle can be found using the formula cos(θ) = ~v · w
~ for two
unit vectors.
e. Show that for x > 0, arccos((1 − x2 )/(1 + x2 )) = 2 arctan(x). Hint: You
may nd the following half-angle formula useful: tan(θ/2) = tan(θ)/(sec(θ) +
1).
13. Prove that (P2 , S) is isotropic.
Without much fanfare, we have shown that the geometry (P2 , S) satises the
rst four of Euclid's postulates, but fails to satisfy the fth. This is also the
case with hyperbolic geometry (D, H). Moreover, the elliptic version of the
fth postulate diers from the hyperbolic version. It is the purpose of this
section to provide the proper fanfare for these facts.
Recall Euclid's ve postulates:
1. One can draw a straight line from any point to any point.
5. Given a line and a point not on the line, there is exactly one line through
the point that does not intersect the given line.
long distance from any point, we can describe a circle of any radius about the
point.
The fourth postulate follows since Möbius transformations preserve angles
and the maps in S are special Möbius transformations.
The fth postulate fails because any two elliptic lines intersect (Theo-
rem 6.2.13). Thus, given a line and a point not on the line, there is not a
single line through the point that does not intersect the given line.
Recall that in our model of hyperbolic geometry, (D, H), we proved that
given a line and a point not on the line, there are two lines through the point
that do not intersect the given line.
So we have three dierent, equally valid geometries that share Euclid's
rst four postulates, but each has its own parallel postulate. Furthermore,
on a small scale, the three geometries all behave similarly. A tiny bug living
on the surface of a sphere might reasonably suspect Euclid's fth postulate
holds, given his limited perspective. A tiny bug in the hyperbolic plane would
reasonably conclude the same. Small triangles have angles adding up nearly
◦
to 180 , and small circles have areas and circumferences that are accurately
described by the Euclidean formulas πr2 and 2πr. We explore geometry on
surfaces in more detail in the next chapter.
7
Geometry on Surfaces
In hyperbolic geometry (D, H) and elliptic geometry (P2 , S), the area of a
triangle is determined by the sum of its angles. This is a signicant dierence
from Euclidean geometry, in which a triangle with three given angles can be
built to have any desired area. Does this mean that if a bug lives in a world
adhering to elliptic geometry, it can never stumble upon a triangle with three
right angles having area 3π ? Yes and no. In the elliptic geometry as dened in
Chapter 6, no such triangle exists because a triangle with 3 right angles must
have area
π
( π2 + π π
2 + 2 ) − π = 2 . So the answer appears to be yes. However,
2
the elliptic geometry (P , S) models the geometry of the unit
sphere, and this
choice of sphere radius is somewhat arbitrary. What if the radius of the sphere
changes? Imagine a triangle with three right angles having one vertex at the
north pole and two vertices on the equator. If the sphere uniformly expands,
the angles of the triangle will stay the same, but the area of the triangle will
increase. So, if a bug is convinced she lives in a world with elliptic geometry,
but is also convinced she has found a triangle with three right angles and area
3π , the bug might be drawn to conclude she lives in a world modeled on a
larger sphere than the unit 2-sphere.
The key geometrical property of a space dictating the relationship between
the angles of a triangle and its area is called curvature. Curvature also dictates
the relationship between the circumference of a circle and its radius.
7.1 Curvature
Consider the smooth curve in Figure 7.1.1. The curvature of the curve at a
point is a measure of how drastically the curve bends away from its tangent
line, and this curvature is often studied in a multivariable calculus course.
The radius of curvature at a point corresponds to the radius of the circle that
best approximates the curve at this point. The radius r of this circle is the
reciprocal of the curvature k of the curve at the point: k = 1/r.
130
7.1. Curvature 131
its tangent plane at the point. There are three fundamental types of curvature.
A surface has positive curvature at a point if the surface lives entirely on one
side of the tangent plane, at least near the point of interest. The surface
has negative curvature at a point if it is saddle-shaped, in the sense that the
tangent plane cuts through the surface. Between these two cases is the case
of zero curvature. In this case the surface has a line along which the surface
agrees with the tangent plane. For instance, a cylinder has zero curvature, as
suggested in Figure 7.1.2(c).
1.2
(a)
c/(2πr)
1
(b)
ratio
(c)
0.8
Figure 7.1.3: Plotting the ratio c/2πr against r in (a) hyperbolic geometry,
(b) Euclidean geometry, and (c) elliptic geometry.
N
r
x
θ s
1 the term `space' is intentionally vague here. Our space needs to have a well-dened
metric, so that it makes sense to talk about radius and circumference. The space might
be the Euclidean plane, the hyperbolic plane or the sphere. Other spaces are discussed in
Section 7.5.
7.1. Curvature 133
r
c = 2πs sin .
s
d2 2πs sin(r/s)
k = −3 lim .
r→0+ dr 2 2πr
Cancelling the 2π terms and replacing sin(r/s) with its power series
expansion, we have
r3 r5
d2 s( rs − 6s 3 + 120s5 − · · · )
k = −3 lim+ 2
r→0 dr r
2 2
r4
d r
= −3 lim+ 2 1 − 2 + − ···
r→0 dr 6s 120s4
12r2
−1
= −3 lim+ + − ··· .
r→0 3s2 120s4
Note that all the terms of the second derivative after the rst have
powers of r in the numerator, so these terms go to 0 as r → 0+ , and
the curvature of the sphere at the north pole is 1/s2 . In fact because
the sphere is homogeneous, the curvature at any point is
1
k= .
s2
r3 r5
sinh(r) = r + + + ··· .
3! 5!
d2 2π sinh(r)
k = −3 lim+ 2
r→0 dr 2πr
2
r2 r4
d
= −3 lim+ 2 1 + + + ···
r→0 dr 3! 5!
1 12r2
= −3 lim + + ··· .
r→0+ 3 5!
Again, each term of the second derivative after the rst has a power
of r in its numerator, so in the limit as r → 0+ , each of these terms
vanishes. Thus, the curvature of the hyperbolic plane in (D, H) is
k = −1.
134 7. Geometry on Surfaces
Exercises
1. Use our working denition to show that the curvature of the projective
plane in elliptic geometry is 1. Recall, c = 2π sin(r) in this geometry.
2. Use our working denition to explain why the curvature of the Euclidean
plane is k = 0.
1 1
if za = − then T (za ) = − .
kz kT (z)
b
2|r 0 (t)|
Z
L(r) = dt.
a 1 + k|r(t)|2
As before, arc-length is an invariant, and the shortest path between two
points is along the elliptic line through them. In the exercises we derive a
formula for the distance between points in this geometry. The greatest possible
√
distance between two points in (P2k , Sk ) turns out to be π/(2 k).
The area of a region R given in polar form is computed by the formula
ZZ
4r
A(R) = drdθ.
R (1 + kr2 )2
To compute the area of a triangle, proceed as in Chapter 6. First, tackle
the area of a lune , a 2-gon whose sides are elliptic lines in (P2k , Sk ).
Lemma 7.2.1. Assume k > 0. A lune in (P2k , Sk ) with interior angle α has
area 2α/k.
Proof. Without loss of generality, we may consider the vertex of our lune
to be the origin. As before, elliptic lines through the origin must also pass
through ∞, so our two lines forming the lune are Euclidean lines. After a
convenient rotation, we may further assume one of these lines is the real axis,
so that the lune resembles the one in Figure 6.3.10. To compute the area of
the lune, compute the integral
√
Z α Z 1/ k
4r
A=2 drdθ.
0 0 (1 + kr2 )2
2
Letting
√u = 1 + kr so that du = 2krdr, the bounds of integration change
from [0, 1/ k] to [1, 2]. Then,
Z α Z 2 Z α
2 du 4 1 2α
A=2 2
dθ = dθ = .
0 k 1 u k 0 2 k
Thus, the angle of a lune with interior angle α is 2α/k .
We remark that the lune with angle
√ π actually covers the entire disk of
radius 1/ k . Thus, the area of the entire space
√ P2k is 2π/k , which matches
√
half the surface area of a sphere of radius 1/ k . We often call s = 1/ k the
radius of curvature for the geometry; it is the radius of the disk on which
we model the geometry.
Also, the integral computation in the proof of Lemma 7.2.1 reveals the
following useful antiderivative:
−2
Z
4r
2 2
dr = + C.
(1 + kr ) k(1 + kr2 )
This fact may speed up future integral computations.
1
A= (α + β + γ − π).
k
Can you nd the area of the triangle formed by Paris, New York, and
Rio? Use a globe, a protractor, and some string. The string follows a
geodesic between two points when it is pulled taut.
Exercises
1. Prove that for k > 0, any transformation in Sk has the form
z − z0
T (z) = eiθ ,
1 + kz0 z
where θ is any real number and z0 is a point in P2k . Hint: Follow the derivation
of the transformations in S found in Chapter 6.
dk (0, x) = 2s arctan(x/s).
b. The circumference of the circle centered at the origin with elliptic radius
√
r < π/(2 k) is C = 2πs sin(r/s).
c. The area of the circle centered at the origin with elliptic radius
√
2 2 r
r < π/(2 k) is A = 4πs sin .
2s
4. In this exercise we investigate the idea that the elliptic formulas
in Exercise 7.2.3 for distance, circumference, and area approach Euclidean
formulas when k → 0+ .
a. Show that the elliptic distance dk (0, x) from 0 to x, where 0 < x ≤ s,
approaches 2x as k → 0+ (twice the usual notion of Euclidean distance).
b. Show that the elliptic circumference of a circle with elliptic radius r
approaches 2πr as k → 0+ .
c. Show that the elliptic area of this circle approaches πr2 as k → 0+ .
5. Triangle trigonometry in (P2k , Sk ).
Suppose we have a triangle in (P2k , Sk ) with side lengths a, b, c and angles
α, β, γ as pictured in Figure 6.3.13.
a. Prove the elliptic law of cosines in (P2k , Sk ):
√ √ √ √ √
cos( kc) = cos( ka) cos( kb) + sin( ka) sin( kb) cos(γ).
7.3. Hyperbolic Geometry with Curvature k<0 137
the origin in C. That is, Dk consists of all z in C such that |z| < 1/ |k| . In
p
this setting, the circle at innity is the boundary circle |z| = 1/ |k|.
The group Hk consists of all Möbius transformations that send Dk to itself.
The geometry (Dk , Hk ) with k < 0 is called hyperbolic geometry with
curvature k. Pushing analogy with the elliptic case, we may dene the group
of transformations to consist of all Möbius transformations with this property:
if z and z ∗ are symmetric with respect to the circle at innity then T (z) and
T (z ∗ ) are also symmetric with respect to the circle at innity. Noting that the
∗ 1 1
point symmetric to z with respect to this circle is z = = − , we draw
|k|z kz
the satisfying conclusion that T ∈ Hk if and only if the following holds:
1 1
if z∗ = − then T (z ∗ ) = − .
kz kT (z)
Thus, the group Hk in the hyperbolic case has been dened precisely as the
group Sk in the elliptic case. Furthermore, one can show that transformations
in Hk have the form
z − z0
T (z) = eiθ ,
1 + kz0 z
where z0 is a point in Dk .
Straight lines in this geometry are the clines in C+ orthogonal to the circle
at innity. By Theorem 3.2.8, a straight line in (Dk , Hk ) is precisely a cline
with the property that if it goes through z then it goes through its symmetric
−1
point .
kz
The arc-length and area formulas also get tweaked by the scale factor, and
now look identical to the formulas for elliptic geometry with curvature k.
The arc-length of a smooth curve r in Dk is
b
2|r 0 (t)|
Z
L(r) = dt.
a 1 + k|r(t)|2
As in Chapter 5 when k was xed at -1, the area formula is a bear to use,
and one may convert to an upper half-plane model to determine the area of
2
a
3 -ideal triangle in Dk . The ambitious reader might follow the methods of
138 7. Geometry on Surfaces
2
3 -ideal triangle in (Dk , Hk ) (k < 0) with
Section 5.5 to show that the area of a
1
interior angle α is − (π − α).
k
With this formula in hand, we can derive the area of any triangle in Dk in
terms of its angles, exactly as we did in Chapter 5.
v
θ w
z d
Proof. For this proof, let s = √1 . Note that s is the Euclidean radius of
|k|
the circle at innity in the disk model for hyperbolic geometry with curvature
k. Since angles and lines and distances are preserved, assume z is the origin
and L is orthogonal to the positive real axis, intersecting it at the point x (with
0 < x < s).
7.3. Hyperbolic Geometry with Curvature k<0 139
s
r
θ
x x∗
L
tan(θ)
tan(θ/2) = .
sec(θ) + 1
s+x
d = s ln
s−x
so that
ed/s − 1
x=s· .
ed/s + 1
Expressr in terms of d by rst expressing it in terms of x. Note that
segment xx∗ is a diameter of the circle containing L, where x∗ = −1 kx is the
point symmetric to x with respect to the circle at innity. Thus, r is half the
∗
distance from x to x :
1 + kx2
r=− .
2kx
Replacing k with −1/s2 , we have
s2 − x2
r= .
2x
One checks that after writing x in terms of d, r is given by
2sed/s
r= .
e2d/s − 1
Substitute this expression for r into the equation labeled (1) in this proof,
and after a dose of satisfactory simplifying one obtains the desired result:
tan(θ/2) = e−d/s .
p
Since s = 1/ |k| this completes the proof.
140 7. Geometry on Surfaces
Parallax If a star is relatively close to the Earth, then as the Earth moves
in its annual orbit around the Sun, the star will appear to move relative to
the backdrop of the more distant stars. In the idealized picture that follows,
e1 and e2 denote the Earth's position at opposite points of its orbit, and the
star s is orthogonal to the plane of the Earth's orbit. The angle p is called the
parallax , and in a Euclidean universe, p determines the star's distance from
the Sun, D, by the equation D = d/ tan(p), where d is the Earth's distance
from the Sun.
e1
d
p
Sun s
D
e2
e2
α
2d
2p u
e1 a s
The angle α = ∠e1 e2 s in Figure 7.3.5 is less than the angle of parallelism
θ = ∠e1 e2 u. Noting that tan(x) is an increasing function and applying
Lobatchevsky's formula it follows that
√
tan(α/2) < tan(θ/2) = e− |k|2d
.
2
ln(tan(α/2))
> |k|. x2 is decreasing for x<0
2d
To get a bound for |k| in terms of p, note that α ≈ π/2 − 2p (the triangles
◦
used in stellar parallax have no detectable angular deviation from 180 ), so
2
ln(tan(π/4 − p))
|k| < .
2d
We remark that for values of p near 0, the expression ln(tan(π/4 − p)) has
linear approximation equal to −2p, so a working bound for k , which appeared
2
in Schwarzschild's 1900 paper [27], is |k| < (p/d) .
Exercises
1. Prove that for k < 0, any transformation in Hk has the form
z − z0
T (z) = eiθ ,
1 + kz0 z
where θ is any real number and z0 is a point in Dk . Hint: Follow the derivation
of the transformations in H found in Chapter 5.
2.
p
Assume k < 0 and let s = 1/ |k|. Derive the following measurement
formulas in (Dk , Hk ).
a. The length of a line segment from 0 to x, where 0 < x < s is
s+x
dk (0, x) = s ln .
s−x
Denition 7.4.1. For each real number k the geometry (Xk , Gk ) has space
Dk
if k < 0;
Xk = C if k = 0;
2
Pk if k > 0,
unique line through p and q is the Euclidean line), and when k=0 the arc-
length is simply twice the usual Euclidean arc-length. So while we are scaling
distances in (X0 , G0 ), Euclidean geometry applies: triangles are Euclidean
◦
triangles and have angle sum equal to 180 . Triangles with a right angle are
Euclidean right triangles and satisfy the Pythagorean theorem.
Thus, we treat (Xk , Gk ) as one big family of geometries. The sign of k
dictates the type of geometry we have, and the magnitude of k dictates the
radius of the disk in which we model the geometry (unless k = 0 in which case
the space is C). Morevoer, Euclidean geometry (X0 , G0 ) marks the edge of the
knife from which we move into a hyperbolic world is k drops below 0, and into
an elliptic world if k rises above 0.
We now summarize some results established in the previous sections and
emphasize key features common to all (Xk , Gk ).
First and foremost, we note that arc-length is an invariant function of
(Xk , Gk ) and that the arc-length ensures that the shortest path from p to q
in (Xk , Gk ) is along the line between them. We have discussed these facts in
the cases k = −1, 0, 1, and the result holds for arbitrary k . So, the arc-length
formula provides a metric on (Xk , Gk ): Given p, q ∈ Xk , we dene dk (p, q) to
be the length of the shortest path from p to q . The circle in (Xk , Gk ) centered
at p through q consists of all points in Xk whose distance from p equals dk (p, q).
kA = (α + β + γ − π).
Proof of this tidy result has already appeared in pieces (see Exercise 1.2.2,
Lemma 7.2.2, and Lemma 7.3.1); we emphasize that this triangle area formula
reveals the locally Euclidean nature of all the geometries (Xk , Gk ): a small
144 7. Geometry on Surfaces
◦
triangle (one with area close to 0) will have an angle sum close to 180 . Observe
also that the closer |k| is to 0, the larger a triangle needs to be in order to detect
an angle sum dierent from 180◦ . Of course, if k = 0 the theorem tells us that
the angles of a Euclidean triangle sum to π radians.
∆1 ∆2
∆3
∆4
∆6 ∆5
By Theorem 7.4.3, the area Ai of the ith triangle ∆i is related to its angle
sum by
X
kAi = ( angles in ∆i ) − π.
Thus,
n−2
X
kA = kAi
i=1
n−2
X X
= ( angles in ∆i − π).
i=1
Now, the total angle sum of the n−2 triangles equals the interior angle
sum of the n-gon, so it follows that
n
X
kA = αi − (n − 2)π.
i=1
Evaluating this integral gives the result, and the details are left as an
exercise.
4πt2
A(c) = .
1 + kt2
x2 +y 2
Using the fact that t2 = 1+k2 x2 y 2 , one can now check by direct substitution
that
k
A(c) = A(a) + A(b) − A(a)A(b).
2π
While we have proved the theorem, it feels a bit like we have missed the
best part - discovery of the relationship. For more on this, we encourage the
reader to consult [20].
p
p
c c
c p a a
a θ θ
0 θ 0 0
a x a x a x
(a) (b) (c)
146 7. Geometry on Surfaces
k = −1
2
k=0
c
1.5
hypotenuse
k=1
0.5
s+x
s ln if k < 0;
s−x
dk (0, x) =
2x if k = 0;
2s arctan(x/s) if k > 0.
2πs sinh(r/s)
if k < 0;
C(r) = 2πr if k = 0;
2πs sin(r/s) if k > 0.
Exercises
1. Check that the measurement formulas in (Xk , Gk ) are correct when
k = 0. In particular, show that d0 (0, x) = 2x for any x > 0 on the real axis,
and that a circle with radius r as measured in (X0 , G0 ) has area A(r) = πr2
and circumference C(r) = 2πr.
2. Complete the proof of Lemma 7.4.6.
3. Use Denition 7.1.4 to prove that for all real numbers k , the curvature
of(Xk , Gk ) is indeed equal to k. Hint: Tackle three cases: k < 0, k = 0, and
k > 0.
4. Suppose an intrepid team of two-dimensional explorers sets out to
determine which 2-dimensional geometry is theirs. Their cosmologists have told
them there world is homogeneous, isotropic, and metric, so they believe that
the geometry of their universe is modeled by (Xk , Gk ) for some real number
k. They carefully measure the angles and area of a triangle. They nd the
◦ ◦ ◦ 2
angles to be 29.2438 , 73.4526 , and 77.2886 , and the area is 8.81 km . Which
geometry is theirs? What is the curvature of their universe?
7.5 Surfaces
On the surface of a donut there are loops one can draw that do not separate
the surface into disjoint pieces. The loop that goes around the donut like an
armband in Figure 7.5.1 is one such loop. Furthermore, this loop cannot be
continuously contracted to a point while staying on the surface. This suggests
the surface of a ball and the surface of a donut are topologically dierent
shapes.
The sphere is an example of a simply connected space because any loop
drawn on the surface can be contracted to a point. The torus is an example of
a multiconnected space because there exist loops in the torus that cannot
be contracted to a point.
Roughly speaking, two spaces are topologically equivalent, or homeomor-
phic , if one can be continuously deformed to look like the other. For instance,
a circle is topologically equivalent to a square. One can be mapped onto the
other via a homeomorphism : a continuous bijection between the objects that
has a continuous inverse. One homeomorphism is suggested in the following
example.
7.5. Surfaces 149
T (z)
z
zo
We are quickly heading into the realm of topology, and must resist the
temptation to dive headlong and formally into this rich subject. In this text
we restrict our focus to a whirlwind tour of topological tools that are used
to investigate possible shapes for two- and three-dimensional universes. We
encourage the reader interested in a more formal approach to the subject to
consult any number of good texts, including [9].
Our candidate universes are examples of manifolds. A topological n-
manifold is a space with the feature that each point in the space has a
neighborhood that looks like a patch of Rn . In cosmology, the spatial section of
space-time is assumed to be a 3-manifold, and when we ask about the shape of
the universe, we are asking about the shape of this 3-manifold. In this section
we focus on 2-manifolds.
n
A bit more formally, R consists of all n-tuples of real numbers p =
(x1 , x2 , . . . , xn ). The open n-ball centered at a point p in Rn with radius
r > 0, denoted Bn (p, r), is the set
Bn (p, r) = {x ∈ Rn | |x − p| < r},
where |x − p| is the Euclidean distance between the points x and p.
For instance, an open 1-ball is an open interval in R. The open 2-ball
B2 (z0 , r) consists of all points z in C such that |z − z0 | < r. The open 3-ball
B3 (p, r) consists of all points in R3 inside the sphere centered at p with radius
r.
r r
p
p−r p p+r z0
s
z
(a) (b)
q a u
a
b p b
q
p b u
a
Figure 7.5.7: The sphere S2 , the torus T2 , and the projective plane P2 .
Note that since Xi is a surface, each point in each space has a neighborhood
homeomorphic to an open 2-ball, so we may always achieve the connected sum
of two surfaces. The result is a new surface: it is still closed, bounded, and
152 7. Geometry on Surfaces
The surface Hg gets its name from the fact that it is topologically equivalent
to a sphere with g handles attached to it. For this reason, we set H0 = S2 . A
few handlebody surfaces are pictured in Figure 7.5.10.
7.5. Surfaces 153
H0 H2
H4
a2 a1
c c a2
a2 a1 a2 a1
(a) c
a1
a2 a1 (c)
c c
a2 a1
(b)
Polygonal Surfaces
Our representation of C2 is an example of a polygonal surface . To build
a polygonal surface, start with a nite number of polygons having an even
number of edges and then identify edges in pairs.
If the surface is built from a single polygon, the edge identications of
the polygon can be encoded in a boundary label . Each edge of the polygon
gets assigned a letter and an orientation. Edges that are identied have the
same letter, and we obtain a boundary label by traversing the boundary of our
polygon in the counter-clockwise direction (by convention) and recording the
letters we encounter, with an exponent of +1 if our walk is in the direction
of the oriented edge, and an exponent of -1 if our walk is in the opposite
direction of the oriented edge. For instance, a boundary label for the surface
C2 found in Figure 7.5.11 is a1 a1 a2 a2 . Furthermore, we may inductively show
that by repeating the connected sum operation of Figure 7.5.11 to Cg−1 #C1 ,
the surface Cg can be represented as a 2g -gon having boundary label
a1 a1 a2 a2 · · · ag ag .
It turns out that all surfaces can be constructed this way, an important and
useful result. We have seen that the torus can be constructed as a polygonal
surface: take a rectangle and identify the edges according to the boundary
label aba−1 b−1 . In Figure 7.5.13 we demonstrate that the 2-holed torus can be
represented as a regular octagon with boundary label
(a1 b1 a−1 −1 −1 −1
1 b1 )(a2 b2 a2 b2 ).
(a1 b1 a−1 −1 −1 −1 −1 −1
1 b1 )(a2 b2 a2 b2 ) · · · (ag bg ag bg ).
The remarkable theorem here is that the handlebody surfaces and the cross-
cap surfaces account for all possible surfaces, without redundancy.
b2 a1
a2 b1
a2 a2 b1 b1
c c b2 a1
c
b2 a1
(a) a2 b1
b2 a2 b1 a1
b2 a1
(c)
a2 c c b1
b2 a1
(b)
Characterizing a surface
If somebody throws a surface at us, how do we know which one we're catching?
One way to characterize a surface is to determine two particular topological
invariants: its orientability status and its Euler characteristic. Let's discuss
what these features of a surface are.
Orientability status
The Möbius strip is obtained from a rectangle by identifying the left and
right edges with a twist:
finis
h
start
look more like half balls). But the Klein bottle in the following example is a
non-orientable surface.
b b
Euler Characteristic
In addition to an orientability status, each surface has attached to it an
integer called the Euler characteristic. The Euler characteristic is a topological
invariant, and it can be calculated from a cell division of a surface, which is a
kind of tiling of the surface by planar faces. Let's make this more precise.
An n-dimensional cell , or n-cell , is a subset of a space whose interior is
homeomorphic to an open n-ball in Rn . For instance, a 1-cell, also called an
edge , is a set whose interior is homeomorphic to an open interval; a 2-cell, or
face , is a set whose interior is homeomorphic to an open 2-ball in the plane. We
call points in a space 0-cells. A 0-cell is also called a vertex (plural vertices ).
v=8,e=12,f=6
v=4, e=6, f=4 v=6, e=12,f=8
v=20,e=30,f=12
v=12,e=30,f=20
(a) (b)
(c) (d)
Cell division (a) has two vertices, six edges, and four faces. One
vertex is in the center of the rectangle, and the other vertex is in the
corner (remember, all four corners get identied to a single point).
As for the edges, four emanate from the center vertex, and we have
two others: the horizontal edge along the boundary of the rectangle
(appearing twice), and the vertical edge along the boundary (also
appearing twice). Thus, the edges of the rectangle are also edges of
the cell division, so the faces in this cell division are triangles, and
there are four of them.
Cell division (b) has six vertices, eight edges, and two faces. There
are four vertices in the interior of the rectangle, one vertex on the
horizontal boundary of the rectangle, and one vertex on the vertical
boundary. Notice that the corner point of the rectangle is not a vertex
of the cell division. To count the edges, observe that four edges form
the inner diamond, and one edge leaves each vertex of the diamond,
for a total of eight edges. The boundary edges of the rectangle do not
form edges in this cell division. Counting faces, we have one inside
7.5. Surfaces 159
the diamond and one outside the diamond. Convince yourself that the
region outside the diamond makes just one face.
You can check that cell division (c) has one vertex in the center of
the rectangle, two edges (one is horizontal, the other is vertical), and
one face.
The attempt (d) fails to be a cell division of the torus. Why is this?
At rst glance, we have four vertices, four edges, and two faces. The
trouble here is the face outside the inner square. It is not a 2-cell -
that is, its interior is not homeomorphic to an open 2-ball. To see this,
note that this region contains a loop that does not separate the face
into two pieces. Can you nd such a loop? Since no open 2-ball has
this feature, the region in question is not homeomorphic to an open
2-ball.
a2
b2 b1
5 2
4 7
a2 a1
3 8
6 1
b2 b1
a1
χ orientable non-orientable
2 H0
1 C1
0 H1 C2
-1 C3
-2 H2 C4
-3 C5
-4 H3 C6
. . .
. . .
. . .
This completes our brief, somewhat informal foray into the topology of
surfaces. Again, several good sources provide a rigorous development of these
ideas, including [9]. In the next section we turn to the task of attaching
geometry to a surface.
7.5. Surfaces 161
a a
b b
c b
d c
c a
c d
b a
(a) (b)
a d
e c
d b
d b
c a
c a
c b
e b
d a
(c) (d)
Exercises
1. Find the Euler characteristic of each surface in Figure 7.5.26. Then
determine whether each surface is orientable or non-orientable. Then classify
the surface.
a1 a2
1 3 s1
s1
b c d e
s2 s2
2 4
a1 a2
c c e e
a1 1 2 a1 a2 3 s1 4 a2
s1 s2 s2
b b d d
b e d c
a1 2 s2 4 a2 3 s1 1 a1
c d e b
If you've got a surface in your hand, you can nd a homeomorphic version
of the surface on which to construct hyperbolic geometry, elliptic geometry,
or Euclidean geometry. And the choice of geometry is unique: No surface
admits more than one of these geometries. As we shall see, of the innitely
many surfaces, all but four admit hyperbolic geometry (two admit Euclidean
geometry and two admit elliptic geometry). Thus, if you randomly generate a
constant curvature surface for a two-dimensional bug named Bormit, Bormit
will no doubt live in a world with hyperbolic geometry.
Chances are, too, that a constant curvature surface cannot be embedded in
three-dimensional space. Only the sphere has this nice feature. In fact, if X
is any surface that lives in R3 , like the handlebody surfaces in Figure 7.5.10,
then it must have at least one point with positive curvature. Why is this?
The surface in R3 is bounded, so there must be some sphere in R3 centered
at the origin that contains the entire surface. Shrink this sphere until it just
bumps into the surface at some point. The curvature of the surface at this
point matches the curvature of the sphere, which is positive.
We have seen that any surface is homeomorphic to a polygonal surface
representing Hg or Cg , for some g , and we now show that each of these may be
given a homogeneous, isotropic and metric geometry (so that it has constant
curvature). A polygonal surface can only run into trouble homogeneity-wise
where the corners come together. Any point in the interior of the polygon has
◦
a nice 360 patch of space about it, as does any point in the interior of an
7.6. Geometry of Surfaces 163
edge. However, if the angles of the corners that come together at a point do
◦
not add up to 360 , then the surface has either a cone point or a saddle-point,
◦
depending on whether the angle sum is less than or more than 360 . As we
saw in Chapter 1 (Exercises 1.3.5 and Checkpoint 1.3.7), in either case, we
will not have a homogeneous geometry; a two-dimensional bug would be able
to distinguish (with triangles or circles) a cone point from a at point from a
saddle point.
To smooth out such cone points or saddle points we change the angles at
the corners so that the angles do ◦
add up to 360 . We now have the means for
doing this. If we need to shrink the corner angles, we can put the polygon in
the hyperbolic plane. If we need to expand the angles, we can put the polygon
in the projective plane.
a2
a2
a3 a2 a3 a2
a3 a1 a3 a1
a1
a1
covers the entire projective plane. In fact, the surface of Example 1.3.9
is the projective plane, and it admits elliptic geometry.
Theorem 7.6.4. Each surface admits one homogeneous, isotropic, and metric
geometry. In particular, the sphere (H0 ) and projective plane (C1 ) admit
elliptic geometry. The torus (H1 ) and Klein bottle (C2 ) admit Euclidean
geometry. All handlebody surfaces Hg with g ≥ 2 and all cross-cap surfaces Cg
with g ≥ 3 admit hyperbolic geometry.
The following section oers a more formal discussion of how any surface
admits one of our three geometries but we present an intuitive argument here.
The sphere and projective plane admit elliptic geometry by construction:
The space in elliptic geometry is the projective plane, and via stereographic
projection, this is the geometry on S2 .
The torus and Klein bottle are built from regular 4-gons (squares) whose
edges are identied in such a way that all 4 corners come together at a point.
In each case, if we place the square in the Euclidean plane all corner angles are
π/2, so the sum of the angles is 2π , and our surfaces admit Euclidean geometry.
Each handlebody surfaces Hg for g ≥ 2 and each cross-cap surfaces Cg for
g ≥ 3 can be built from a regular n-gon where n ≥ 6. Again, all n corners
come together at a single point. A regular n-gon in the Euclidean plane has
interior angle (n − 2)π/n radians, so the corner angles sum to (n − 2)π radians.
This angle sum exceeds 2π radians since n ≥ 6. Placing a small version of this
n-gon in the hyperbolic plane, the corner angle sum will be very nearly equal
to (n − 2)π radians and will exceed 2π radians, but as we expand the n-gon
the corners approach ideal points and the corner angles sum will approach
0 radians. Thus, at some point the angle sum will equal 2π radians on the
nose, and the polygonal surface built from this precise n-gon admits hyperbolic
geometry. Exercise 7.6.4 works through how to construct this precise n-gon.
Of course, one need not build a surface from a regular polygon. For instance,
the torus can be built from any rectangle in the Euclidean plane and it will
inherit Euclidean geometry. So while the type of geometry our surface admits
is determined, we have some exibility where certain geometric measurements
are concerned. For instance, there is no restriction on the total area of the
torus, and the rectangle on which it is formed can have arbitrary length and
width dimensions. These dimensions would have a simple, tangible meaning
to a two-dimensional bug living in the surface (and might be experimentally
determined). Each dimension corresponds to the length of a geodesic path
that would return the bug to its starting point. A closed geodesic path in a
surface is a path that follows along a straight line (in the underlying geometry)
that starts and ends at the same point. Figure 7.6.5 shows three closed geodesic
paths, all starting and ending at a point near a bug's house. The length of one
path equals the width of the rectangle, the length of another equals the length
of the rectangle, and the third follows a path that is longer than the rst two.
7.6. Geometry of Surfaces 165
Figure 7.6.5: A at torus has many closed geodesic paths. The length of the
shortest closed geodesic path equals the length of the short side of the rectangle
on which the torus is modeled.
Even in hyperbolic surfaces, where the area of the surface is xed (for a
given curvature) by the Gauss-Bonnet formula, which we prove shortly, there is
freedom in determining the length of closed geodesic paths.
d1 c2 e1
d2 c1 e2
c2 c2
c1 c1
d2 d2 e2 e2
d1 d1 e1 e1
We label our cuts so that we can stitch up our surface later. Match
the ci , di , and ei edges to recover the two-holed torus. Any pair of pants
can be cut into two hexagons by cutting along the three vertical seams
in the pants. It follows that the two-holed torus can be constructed
from four hexagons, with edges identied in pairs as indicated below.
166 7. Geometry on Surfaces
c1 c2
c2
c1 a1 a3 a1 a3
a1 a3
a2
d2 d2 d1 d1 d2 d2
d1 d1 a2 a2
c1 c2
c2
c1 b1 b3 b1 b3
b1 b3
b2
e2 e2 e1 e1 e2 e2
e1 e1
b2 b2
A surface that admits one of our three geometries will have constant
curvature. The reader might have already noticed that the sign of the curvature
will equal the sign of the surface's Euler characteristic. Of course the magnitude
of the curvature (if k 6= 0) can vary if we place a polygonal surface in a scaled
version of P2 or D. That is, while the type of homogeneous geometry a surface
admits is determined by its Euler characteristic (which is determined by its
shape), the curvature scale can vary if k 6= 0. By changing the radius of a
sphere, we change its curvature (though it always remains positive). Similarly,
the surface C3 in Figure 7.6.2 has constant curvature -1 if it is placed in the
hyperbolic plane of Chapter 5. However, it can just as easily nd itself in the
hyperbolic plane with curvature k = −8. Recall, the hyperbolic plane with
curvature k < 0 is modeled
p on the open disk in C centered at the origin with
radius 1/ |k|. Placing the hexagonal representation of C3 into this space so
that its corner angle sum is still 2π produces a surface with constant curvature
k.
We have nally arrived at the elegant relationship between a surface's
curvature k, its area, and its Euler characteristic. This relationship crystalizes
the interaction between the topology and geometry of surfaces.
so in this case the Gauss-Bonnet formula reduces to the true statement that
0 = 0.
Any surface of constant negative curvature cam be represented by a regular
n-sided polygon with n ≥ 6. Furthermore, this polygon can be placed in the
hyperbolic plane with curvature k < 0, so that its interior angles sum to 2π
radians. According to Theorem 7.4.4,
n
kA = 2π − (n − 2)π = 2π 2 − ,
2
kA = 2π(2 − g) = 2πχ(Cg ).
Exercises
1. Suppose our intrepid team of two-dimensional explorers from Exer-
cise 7.4.4, after an extensive survey, estimates with high condence that the
2 2
area of their universe is between 800,000 km and 900,000 km . Assuming
their universe is homogeneous and isotropic, what shapes are possible for their
universe? Can they deduce from this information the orientability status of
their universe?
What must be the area of a constant curvature C3 surface so that they have
the same curvature?
We may bend a sheet of paper and join its left and right edges together to
obtain a cylinder. If we let I2 = {(x, y) ∈ R2 | 0 ≤ x ≤ 1, 0 ≤ y ≤ 1} represent
168 7. Geometry on Surfaces
[a] = {x ∈ A | x ∼ a}.
= Re(z) − Re(v).
Orbit Spaces
We may construct a natural quotient set from a geometry (X, G).
The group structure of G denes an equivalence relation ∼G on X as follows:
For x, y ∈ X , let
Indeed, for each x ∈ X , x ∼G x because the group G must contain the identity
transformation, so the relation is reexive. Next, if x ∼G y then T (x) = y for
−1 −1
some T in G. But the group contains inverses, so T is in G and T (y) = x.
Thus y ∼G x, and so ∼G is symmetric. Third, transitivity of the relation
follows from the fact that the composition of two maps in G is again in G.
Given geometry (X, G) we let X/G denote the quotient set determined by
the equivalence relation ∼G . In this setting we call the equivalence class of a
point x in X , the orbit of x. So, the orbit of x consists of all points in the
space X to which x can be mapped under transformations of the group G:
Put another way, the orbit of x is the set of points in X congruent to x in the
geometry (X, G).
Note that if the geometry G is homogeneous, then any two points in X
are congruent and, for any x ∈ X , the orbit of x is all of X . In this case the
quotient set X/G consists of a single point, which is not so interesting. We
typically want to consider orbit spaces X/G in which G is a small group of
transformations.
We say that a group of transformations G of X is a group of homeo-
morphisms of X if each transformation in G is continuous. In this case, we
call X/G an orbit space . If X has a metric, we say that a group of transfor-
mations of X is a group of isometries if each transformation of the group
preserves distance between points.
[p] = {p + n | n ∈ Z}.
strip. Furthermore, no two points in the interior of the strip are related.
By passing to the quotient, we are essentially rolling up the plane in
to an innitely tall cylinder. The rolling up process is described by the
map p : C → C/hT1 i given by p(z) = [z].
hR π2 i = {1, R π2 , Rπ , R 3π
2
}.
It turns out that every surface can be viewed as a quotient space of the
form M/G, where M is either the Euclidean plane C, the hyperbolic plane D,
or the sphere S2 , and G is a subgroup of the transformation group in Euclidean
geometry, hyperbolic geometry, or elliptic geometry, respectively. In topology
terminology, the space M is called a universal covering space of the orbit
space M/G.
Figure 7.7.9 depicts two points in the shaded fundamental domain, [u]
and [v]. The distance between them equals the Euclidean distance in C of
the shortest path between any point in equivalence class [u] and any point in
equivalence class [v]. There are many such nearest pairs, and one such pair
is labeled in Figure 7.7.9 where z is in [u] and w is in [v]. Also drawn in the
gure is a solid line (in two parts) that corresponds to the shortest path one
would take within the fundamental domain to proceed from [u] to [v]. This
path marks the shortest route a ship in the video game from Chapter 1 could
take to get from [u] to [v].
7.7. Quotient Spaces 173
[u]
w [v]
Figure 7.7.9: The distance between two points in the torus viewed as a
quotient of C.
c b
d
a
c m
b
d a
c b
d a
c b
d a
All surfaces Hg for g≥2 and Cg for g≥3 can be viewed as quotients of D
by following the procedure in the previous example.
Start with a perfectly sized polygon in D. The polygon must have corner
angle sum equal to 2π radians, and the edges that get identied must have equal
length so that an isometry can take one to the other. (In every example so
far, we have used regular polygons in which all sides have the same length, but
asymmetric polygons will also work.) Next, for each pair of oriented edges to be
identied, nd a hyperbolic isometry that maps one onto the other (respecting
the orientation of the edges). The group generated by these isometries creates
a quotient space homeomorphic to the space represented by the polygon, and it
inherits hyperbolic geometry. Note also that the initial polygon can be moved
by the isometries in the group to tile all of D without gaps or overlaps.
7.7. Quotient Spaces 175
Dirichlet Domain
We end this section with a discussion of the Dirichlet domain, which is an
important tool in the investigation of the shape of the universe. Suppose we
live in a surface described as a quotient M/G where M is either C , D, or S2 ,
and G is a xed-point free and properly discontinuous group of isometries of the
space. For each point x in M dene the Dirichlet domain with basepoint
x to consist of all points y in M such that
b b
i
c c
a 1 a 2
b b b b
c T (A) c T ◦ r2 (A) c
a a a a a a
c A c r2 (A) c r4 (A) c
a
a a a a a a
c T −1 ◦ r−1 (A) c T −1 ◦ r(A) c T −1 ◦ r3 (A) c
b b b b b b
Exercises
1. Show that the Dirichlet domain at any point of the torus in Example 7.7.8
is an a by b rectangle by completing the following parts.
a. Construct an a by b rectangle to be the fundamental domain, and place
eight copies of this rectangle around the fundamental domain as in Figure 7.7.9.
Then plot a point x in the fundamental domain, and plot its image in each of
the copies.
b. For each image x0 of x, construct the perpendicular bisector of the
segment xx0 . The eight perpendicular bisectors enclose the Dirichlet domain
based at x. Prove that the Dirichlet domain is also an a by b rectangle.
[w] [z]
a [w] [w]
[z] [z]
b
[z] [z]
b
[w] [z]
a
Recall that R3 is the set of ordered triples of real numbers (x, y, z), and a
3-manifold is a space with the feature that every point has a neighborhood
that is homeomorphic to an open 3-ball. We assume that the shape of the
universe at any xed time is a 3-manifold. Evidence points to a universe that
is isotropic and homogeneous on the largest scales. If this is the case, then just
as in the two-dimensional case, the universe admits one of three geometries:
hyperbolic, elliptic, or Euclidean. This section oers a brief introduction to the
three-dimensional versions of these geometries before turning to 3-manifolds.
The reader is encouraged to see [12] for a broader intuitive discussion of these
ideas, or [11] for a more rigorous approach.
178
8.1. Three-Dimensional Geometry and 3-Manifolds 179
The unit 2-sphere S2 bounds H3 , and is called the sphere at innity. Points
3
on the sphere at innity are called ideal points and are not points in H .
The group of transformations for three-dimensional hyperbolic geometry
is generated by inversions about spheres that are orthogonal to the sphere at
innity. Inversion about a sphere is dened analogously to inversion in a circle.
Suppose S is a sphere in R3 centered at v0 with radius r, and v is any point
in R 3
. Dene the point symmetric to v with respect to S to be the point v∗
−
v−
→
on the ray 0v such that
|v − v0 ||v ∗ − v0 | = r2 .
One may prove that inversion about a sphere S will send spheres orthogonal
to S to themselves. So, composing two inversions about spheres orthogonal to
S2∞ generates an orientation preserving transformation of hyperbolic three-
space H3 . The transformation group consists of all such compositions. Lines
in this geometry correspond to arcs of clines in H3 that are orthogonal to the
sphere at innity (see line L in Figure 8.1.1). These lines are geodesics in H3 .
3
As in the two-dimensional case, Euclidean lines through the origin of H are
also hyperbolic lines. A plane in this geometry corresponds to the portion of a
sphere or Euclidean plane inside H3 that meets the sphere at innity at right
angles, such as planes P1 and P2 in Figure 8.1.1. If we restrict our attention
to any plane in H3 , we recover the two-dimensional hyperbolic geometry of
Chapter 5.
P1
P2
L S2∞
Figure 8.1.1: One model of hyperbolic space H3 , the open unit 2-ball.
S3 = {(x, y, z, w) ∈ R4 | x2 + y 2 + z 2 + w2 = 1}.
I3 = {(x, y, z) | 0 ≤ x, y, z ≤ 1},
and all the images of this cube under the transformations in Γ tile R3 .
3
The 3-torus is a quotient of R generated by Euclidean isometries, and
it inherits Euclidean geometry.
The 3-torus is one of just ten compact and connected 3-manifolds admiting
Euclidean geometry. Of these ten, six are orientable, and four are non-
orientable. The 3-torus is orientable, and the other ve orientable Euclidean
manifolds are presented below in Example 8.1.7 and Example 8.1.9.
The elliptic 3-manifolds have also been classied. There are innitely many
dierent types, and it turns out they are all orientable. The 3-sphere is the
simplest elliptic 3-manifold, and Albert Einstein assumed the universe had this
shape when he rst solved his equations for general relativity. He found a static,
nite, simply connected universe without boundary appealing for aesthetic
reasons, and it cleared up some paradoxes in physics that arise in an innite
universe. However, the equations of general relativity only tell us about the
local nature of space, and do not x the global shape of the universe. In 1917,
the Dutch astronomer Willem De Sitter (1872-1934) noticed that Einstein's
solutions admitted a another, dierent global shape: namely three-dimensional
elliptic space obtained from the 3-sphere by identifying antipodal points, just
as we did in Chapter 6 in the two-dimensional case.
S3 = {(x, y, z, w) | x2 + y 2 + z 2 + w2 = 1}.
gure shows a path from point a to point b in S3 that goes through the
point p on the boundary of the left-hand solid ball before entering the
right-hand solid ball through the point on its boundary with which p
has been identied (also called p).
p p
Let c3
R denote real 3-space with a point at innity attached to it. To
construct the stereographic projection map, it is convenient to rst
express R3 in terms of pure quaternions. A pure quaternion is a
quaternion q whose scalar term is 0. That is, a pure quaternion has
form q = bi + cj + dk. The point (x, y, z) in R is identied with
3
the
the pure quaternion q = xi + y j + z k.
Though we are working in four-dimensional space, Figure 3.3.4
serves as a guide in the construction of the stereographic projection
map. Let N = (0, 0, 0, 1) be the north pole on S3 . If P = (a, b, c, d) is
3
any point on S other than N , construct the Euclidean line through N
and P. This line has parametric form
i + 1−d j + 1−d k
(
a b c
1−d if d 6= 1;
φ((a, b, c, d)) =
∞ if d = 1.
8.1. Three-Dimensional Geometry and 3-Manifolds 183
Extending the ideas of Section 7.7 to this case, one can show that the
Poincaré dodecahedral space space is a quotient of the 3-sphere S3 by a
group of isometries. The dodecahedron pictured here is a fundamental
domain of the space. Just as the polygons in Section 7.7 tiled the space
in which they lived, and the cubical fundamental domain of the 3-torus
tiles R3 , the dodecahedron tiles S3 . Although far from obvious, it turns
3
out that S is tiled by 120 copies of this dodecahedron!
1 1
v0 v4 v3
v1 v2
s
L(5, 2)
Let Ni denote the face in the northern hemisphere whose vertices are
vi , vi+1 , n. Let Si be the face in the southern hemisphere whose vertices
are vi , vi+1 , and s. The lens space L(p, q) is created by identifying face
Ni with face Si+q where the sum i + q is taken modulo p. For instance,
in L(5, 2), the shaded face N0 in the gure is identied with the shaded
face S2 (and the circle of points pictured in N0 gets sent to the circle
of points in S2 ). The faces N1 and S3 are identied, as are N2 with S4 ,
N3 with S0 and N4 with S1 .
A more elegant description of this manifold makes use of the orbit space
construction of Section 7.7. Consider a unit cube sitting in R3 . Let T1 be
186 8. Cosmic Topology
◦
the following Euclidean transformation: Rotate by 180 about segment 1 in
Figure 8.1.8(a) and then translate along the length of that segment. The
original cube and its image are pictured in Figure 8.1.8(b). Let T2 be dened
similarly using segment 2 of Figure 8.1.8(a). The original cube and its image
under T2 are pictured in Figure 8.1.8(c). The transformation T3 is dened
the same way using segment 3, and Figure 8.1.8(d) depicts its eect on the
original cube. Each of these transformations is a screw motion. In particular,
◦
each screw motion consists of rotation by 180 and translation by unit length.
The Hantzsche-Wendt manifold is then the quotient of R3 by the group of
Euclidean isometries generated by the three screw motions, and it inherits
Euclidean geometry.
(3)
(2) (1)
(a) (b)
(c) (d)
F
F
F F
(a) (b)
F
F
120◦ 60◦
F
(c) (d)
Figure 8.1.10: Four Euclidean 3-manifolds: (a) the quarter turn manifold;
(b) the half turn manifold; (c) the one-third turn manifold; and (d) the one-
sixth turn manifold.
Exercises
188 8. Cosmic Topology
1. Investigating quaternions.
a. Show that ij = k, jk = i, and ki = j.
b. Show that ji = −k, kj = −i, and ik = −j.
The conjugate of a quaternion q = a + bi + cj + dk is q = a − bi − cj − dk.
∗
c.
√
The modulus of q is |q| = a2 + b2 + c2 + d2 . Prove that q · q ∗ = |q|2 for any
quaternion.
d. Prove that |uv| = |u| · |v| for any quaternions. Thus, the transformations
in three-dimensional elliptic geometry, which have the form T (q) = uqv where
u, q, v are unit quaternions, do send a point of S3 to a point on S3 .
2. Suppose P = (a, b, c, d) and −P = (−a, −b, −c, −d) are diametrically
opposed points on S3 . Prove that
where φ : S3 → R
c3 is stereographic projection.
3. Verify that in the 3-torus of Example 8.1.2, all eight corners come together
at a single point.
6. What happens if you identify each pair of opposite faces in a cube with
a one-quarter twist? Convince yourself that not all 8 corners come together at
a single point, so that the result is not a Euclidean 3-manifold.
(2)
G
(3)
(3) (1)
E
(2)
Figure 8.2.1: Seeing multiple images of the same object in the torus.
In fact, if we suppose for a minute that we can see as far as we wish, then
we would be able to see G by looking in any direction that produces a line of
sight with rational slope. In reality, we can't see forever, and this limitation
produces a visual boundary. We will let robs denote the distance to which
we can see, which is the radius of our observable universe. To have any hope
8.2. Cosmic Crystallography 189
of seeing multiple images of the same object, the diameter of our observable
universe, 2robs , must exceed some length dimension of the universe.
Getting back to the torus, the easiest way to nd the directions in which
one can view G is to tile the plane with identical copies of the torus. Place the
Earth at the same point of each copy of the rectangle, and the same goes for
other objects such as G. Figure 8.2.2 displays a portion of the tiling, and our
visual boundary. According to Figure 8.2.2, in addition to the instance of G
in the fundamental domain, 5 of its images would be visible.
G G G
E E E
(2)
G (3) (1)
G G
E E E
robs
G G G
E E E
visual boundary
10
10 5 0 5 10
10
Rather than hunt for two copies of the same source, we compute the distance
between each pair of sources in the catalog (we must make an assumption about
the geometry of space to compute these distances), giving us N (N − 1)/2
distances. If the catalog contains no repeat images, then the distances ought
to follow a Poisson probability distribution. The histogram of the N (N − 1)/2
distances is called a pair separation histogram (PSH) and is given in
Figure 8.2.4.
0
2R
distance
10 10
5 5
10 5 0 5 10 10 5 0 5 10
5 5
10 10
(a) (b)
0 5 10 15 20
What causes these spikes? Look at the catalog plot again in Figure 8.2.5(b),
and nd an object near the top edge of the fundamental domain. We've
highlighted a group of objects that look a bit like a sled. There is a copy
of this object just below the bottom edge of the fundamental domain. The
distance between a point in this sled and its image below equals the length of
the width dimension of our torus, which is 14 in this simulation. Indeed, there
are lots of points for which we see an image displaced vertically in this manner,
so this distance will occur lots of time in the pair separation histogram.
Another copy of our big sled appears just to the right of the right edge of
the fundamental domain. So, the distance between a point and its horizontally
displaced image will be equal to the length dimension of the torus, which is
10 in this simulation. There are also lots of these pairs in the catalog - in fact
more than before since this dimension of the fundamental rectangle is smaller.
This accounts for the larger spike in the histogram above the distance 10.
Finally, there are some objects in the fundamental domain for which
a diagonally displaced copy is visible. The length of this diagonal is
√
102 + 142 ≈ 17.2, and this accounts for the third, smallest spike in the
histogram.
Spikes appear in the PSH precisely because the torus universe C/Γ is
constructed from isometries that move each point in the space by the same
distance. A transformation T of a metric space with the property that
for all points p and q in the space is called a Cliord translation . Any
translation in C is a Cliord translation; every point gets moved the same
distance. However, a rotation about the origin is not: the further a point is
from the origin, the further it moves. In the exercises you prove that non-trivial
isometries in the hyperbolic plane are not Cliord translations.
Recall that to create the PSH of a catalog in the cosmic crystallography
method we must rst make an assumption about the geometry of the universe.
The PSH in Figure 8.2.6 was generated by computing the Euclidean distances
between points in the catalog. If we assumed a hyperbolic universe and used
the hyperbolic metric to produce the PSH, the spikes would vanish.
In essence, the method of cosmic crystallography pours over catalogs of
astronomical objects, computing distances between all pairs of objects in the
catalog, and then looking for spikes in the pair separation histogram. The
192 8. Cosmic Topology
I T (q)
T (p)
II
II
q
p I
Type I pairs will not produce discernible spikes in the PSH. Even
in simulations for which several images of a pair of points are present,
8.2. Cosmic Crystallography 193
the spike generated by this set of pairs having the same distance is not
statistically signicant.
The collecting correlated pairs (CCP) method, outlined below,
attempts to detect the Type I pairs in a catalog.
Suppose a catalog has N objects, and let P = N (N − 1)/2 denote
the number of pairs generated from this set. Compute all P distances
between pairs of objects, and order them from smallest to largest. Let
∆i denote the dierence between the (i + 1)st distance and the ith
distance. Notice ∆i ≥ 0 for all i, and i runs from 1 up to P − 1.
Now, ∆i = 0 for some i if two dierent pairs of objects in the catalog
have the same separation. It could be that unrelated pairs happen to
have the same distance, or that the two pairs responsible for ∆i = 0 are
of the form {p, q} and {T (p), T (q)} (Type I pairs). (We're assuming
there are no type II pairs in the catalog.)
Let Z equal the number of the ∆i 's that equal zero. Then
Z
R=
P −1
denotes the proportion of the dierentials that equal zero. This single
number is a measure, in some sense, of the likelihood of living in a
multiconnected universe.
In a real catalog involving estimations of distances, one wouldn't
expect Type I pairs to produce identical distances, so instead of using
Z as dened above, one might let Z equal the number of the ∆i 's that
are less than , where is some small positive number. For more details
on this method, see [28].
Exercises
1. The Klein Bottle. We may view the Klein bottle as a quotient of C by
the group of isometries generated by T1 (z) = z + i and T2 (z) = z + 1 + i. A
fundamental domain for the quotient is the unit square in C. The edges of the
square are identied as pictured. The a edges are identied as they would be
for a torus, but the b edges get identied with a twist.
a
i
b b
a 1
a. Verify that T1 maps the bottom edge of the unit square to the top edge
of the unit square, and that T2 maps the left edge of the unit square to the
right edge with a twist.
b. Determine the inverse transformations T1−1 and T2−1 .
c. We may compose any number of these four transformations T1 , T2 , T1−1 ,
−1
and T2 to tile all of C with copies of the unit square. In the following gure
we have indicated in certain squares the transformation (built from the four
above) that moves the unit square into the indicated square. Complete the
194 8. Cosmic Topology
T12
T1 T1 ◦ T2
i
T2−1 T2 T22
0 1
T1−1 T1−1 ◦ T2
10
Pair Separation Histogram
10 5 0 5 10
0 5 10 15 20
10
4. A rough estimate for the number of images of an object one might see
in a catalog can be made by dividing the volume of space occupied by the
catalog by the volume of the Dirichlet domain at our position in the universe.
Suppose we live in an orientable two-dimensional universe. In fact, suppose we
8.3. Circles in the Sky 195
live in Hg for some g ≥ 2, and our Dirichlet domain is the standard 4g -gon as
in Figure 7.5.24. Set the curvature of the universe to k = −1.
a. According to Gauss-Bonnet, what is the area of the Dirichlet domain?
b. What is the area of the observable universe, as a function of robs ? That
is, what is the area of a circle in (D, H) with radius robs ?
c. Determine the ratio of the area of the observable universe (A(O.U.)) to
the area of the fundamental domain (A(F.D.)). Your ratio A(O.U.)/A(F.D.)
will depend on robs and g.
d. Complete the following table in the case g = 2. Assume robs has units
in light-years.
Table 8.2.10: Estimating the number of images of an object one might see
in a catalog
Immediately after the big bang, the universe was so hot that the usual
constituents of matter could not form. Photons could not move freely in
space, as they were constantly bumping into free electrons. Eventually, about
350,000 years after the big bang, the universe had expanded and cooled to
the point that light could travel unimpeded. This free radiation is called
the cosmic microwave background (CMB) radiation, and much of it is still
travelling today. The universe has cooled and expanded to the point that this
radiation has stretched to the microwave end of the electromagnetic spectrum,
having a wavelength of about 1 or 2 millimeters.
The CMB radiation is coming to us from every direction, and it has all
been travelling for the same amount of time - and at the same speed. This
means that it has all traveled the same distance to reach us at this moment.
Thus, we may think of the CMB radiation that we can detect at this instant
as having come from the surface of a giant 2-sphere with us at the sphere's
center. This giant 2-sphere is called the last scattering surface (LSS).
It is perhaps comforting to think that everyone in the universe has their
own last scattering surface, that everyone's LSS has the same radius, and that
this radius is growing in time.
The CMB radiation coming to us has a temperature that is remarkably
uniform: it is constant to a few parts in 100,000, which makes the temperature
of the radiation in the LSS very nearly perfectly uniform. As Craig Hogan
points out in [14], this is much smoother than a billiard ball. Nonetheless,
there are slight variations in the temperature. These variations, due to slight
imbalances in the distribution of matter in the early universe, were predicted
well before they were nally found (when our instruments became sensitive
196 8. Cosmic Topology
enough to detect them). These very slight temperature dierences might reveal
the shape of the universe.
Imagine our universe is a giant 3-torus. Assume a fundamental domain for
the universe is a rectangular box as shown in Example 8.1.2, and that this box
is our Dirichlet domain (we're at the center of this box). We may tile R3 with
copies of this fundamental domain, placing ourselves in the same position of
each copy of the fundamental domain. Now, imagine our last scattering surface
in the fundamental domain. In fact, there will be a copy of our last scattering
surface surrounding each copy of us in each copy of the fundamental domain.
If our last scattering surface is small relative to the size of the fundamental
domain, as in Figure 8.3.1(a), then it will not intersect any of its copies.
However, if the last scattering surface is large relative to the size of the
fundamental domain, as in Figure 8.3.1(b), then it will intersect one or more
of its copies. Moreover, adjacent copies of the LSS will intersect in a circle. In
this happy case, our last scattering surface will contain circles with matching
temperature distributions. Look again at Figure 8.3.1(b). We have three copies
of our fundamental domain pictured as well as three copies of the LSS (only one
of which is shaded to make the situation less cluttered). Two vertical circles
of intersection appear in the gure. From our point of view at the center of
the LSS, the two images of the circle will be directly opposite one another in
the sky. Since these circles are one and the same, the temperature distribution
around the two circles will agree. Therein lies the hope. Scan the temperature
distribution in the last scattering surface for matching circles.
(a) (b)
Figure 8.3.1: The LSS compared to the size of a 3-torus universe. In (a)
the LSS is small, so it won't reveal the shape of the universe; in (b) the LSS
intersects itself and will have circles of matching temperature distributions.
(θ1 , φ1 )
(θ2 , φ2 )
When comparing the size of the LSS relative to the size of space, it is
convenient to dene the following length dimension. The injectivity radius
at a point in a manifold, denoted rinj , is half the distance of the shortest closed
geodesic path that starts and ends at that point. A necessary condition, then,
for detecting matching circles in the LSS at our location is that our observable
radius robs exceeds our injectivity radius rinj .
Example 8.3.2: Detecting the 3-torus from the LSS.
If the universe is a 3-torus, and our LSS has diameter larger than some
dimension of the 3-torus, then the LSS will intersect copies of itself, and
the matching circles would be diametrically opposed to one another on
the LSS. Suppose our Dirichlet domain in a 3-torus universe is an a
by b by c box in R3 , where a < b < c. Our LSS is then centered at
the center of the box. The injectivity radius of the universe is a/2.
In the following gure, we assume that robs , the radius of the LSS, is
greater than a/2 but less than b/2. In this case, the circles-in-the-sky
method would detect one pair of matching circles in the temperature
distribution of the LSS. From the Earth E , we would observe that circle
C1 , when traced in the counterclockwise direction, matches circle C2
when traced in the clockwise direction, with no relative phase shift.
b
C1
E
a
C2
c
If the size is right, all six compact orientable Euclidean 3-manifolds would
have matching circles that are diametrically opposed to one another on the
LSS. The phase shift on these matching circles will be non-zero if the faces are
identied with a rotation.
5
1 2
c2 2
c1
4 3
5
3
LSS
In general, the matching circles we (might) see can depend not only on the
shape of the universe, but also on where we happen to be in the universe.
This follows because the Dirichlet domain can vary from point to point
(see Example 7.7.14 for the two-dimensional case). Of the 10 Euclidean 3-
manifolds, only the 3-torus has the feature that the Dirichlet domain is location
independent. Some (but not all) elliptic 3-manifolds have this feature, and in
any hyperbolic 3-manifold, the Dirichlet domain depends on your location.
Thus, if we do observe matching circles, it can not only reveal topology but
also the Earth's location in the universe.
Searches to date have focused on circles that are diametrically opposed to
one another on the LSS (or nearly so). This restriction reduces the search
space from six parameters to four. Happily, most detectable universe shapes
would have matching circles diametrically opposite one another, or nearly so.
At the time of this writing, no matching circles have been found, and this
negative result places bounds on the size of our universe. For instance, an
article written by the people who rst realized a small nite universe would
imprint itself on the LSS [19], concludes from the absence of matching circles
that the universe has topology scale (i.e., injectivity radius rinj ) bigger than
24 gigaparsecs, which works out to 24 × 3.26 × 109 ≈ 78 billion light-years. So,
a geodesic closed path trip in the universe would be at least 156 billion light
years long.
This stupendous distance is hard to fathom, but it appears safe to say
that we might abandon the possibility of gazing into the heavens and seeing a
distant image of our beloved Milky Way Galaxy.
In addition to [19], other accessible papers have been written on the circles-
in-the-sky method, as well as the cosmic crystallography method. (See [23],
[26], and [24].) Je Weeks also discusses both programs of research in The
Shape of Space, [12].
so) after the big bang, the universe expanded at a stupendous rate, causing
the universe to appear homogeneous, isotropic, and also at.
The assumptions of isotropy and homogeneity are remarkably fruitful when
one approaches the geometry and topology of the universe from a mathematical
point of view. Under these assumptions, three possibilities exist for the
geometry of the universe - the three geometries that have been the focus of
this text. Each geometry type has possible universe shapes attached to it, and
Section 8.1 showcases some of the leading compact candidates.
The mathematical point of view gives us our candidate geometries, but
attempts at detecting the geometry from the mathematical theory have proved
unsuccessful. For instance, no enormous, cosmic triangle involving parallax
has produced an angle sum suciently dierent from π radians to rule out
Euclidean geometry.
Adopting a physical point of view, we have another way to approach the
geometry of the universe. Einstein's theory of general relativity ties the
geometry of the universe to its mass-energy content. If the universe has a
high mass-energy content, then the universe will have elliptic geometry. If the
universe has a low mass-energy content, then it will be hyperbolic. If it has
just a precise amount, called the critical mass density, the universe will be
Euclidean. From a naive point of view, this makes it seem highly unlikely that
our universe is Euclidean. If our mass-energy content deviates by the mass
of just one hydrogen atom from this critical amount, our universe fails to be
Euclidean.
A little notation might be helpful. It turns out that from Einstein's eld
equations, the mass-energy density of the universe, ρ, is related to its curvature
k by the following equation, called the Friedmann Equation:
8πG k
H2 = ρ − 2.
3 a
Here G is Newton's gravitational constant; H is the Hubble constant
measuring the expansion rate of the universe; k = −1, 0, or 1 is the curvature
constant; and a is a scale factor. In fact, a and H are both changing in time,
but may be viewed as constant during the present period. Current estimates
of H (see [17] or [22]) are in the range of 68 to 70 kilometers per second per
megaparsec, where 1 megaparsec is 3,260,000 light-years.
In a Euclidean universe, k=0 and solving the Friedmann equation for ρ
gives us the critical density
3H 2
ρc = .
8πG
This critical density is about 1.7 × 10−29 grams per cubic centimeter, and
is the precise density required in a Euclidean universe.
We let Ω equal the ratio of the actual mass-energy density ρ of the universe
to the critical one ρc . That is,
ρ
Ω= .
ρc
Then, if Ω < 1 the universe is hyperbolic; if Ω > 1 the universe is elliptic;
and if Ω = 1 on the nose then it is Euclidean.
Until the late 1990s all estimates of the mass-energy content of the universe
put the value of Ω much less than 1, suggesting a hyperbolic universe. In fact,
dierent observational techniques for estimating the total mass-energy content
200 8. Cosmic Topology
of the universe put the value of Ω at about 1/3, contrary to the value of 1
predicted by the inationary universe model.
But at the dawn of the 21st century, this all changed with detailed mea-
surements of the cosmic microwave background radiation, and the remarkable
discovery that the universe is expanding at an accelerated rate.
Careful analysis of the WMAP data on the cosmic background radiation
1
suggests that the universe is at, or nearly so, in agreement with the theory of
ination. (This analysis is dierent than the circles in the sky method, which
searches for shape.) The ve-year estimate from the WMAP data (see [21])
put the value of Ω at
Ω = 1.0045 ± .013.
So if the mass-energy density is about 1/3 of what is required to get us
to a Euclidean universe, but it appears from the CMB that the universe is
Euclidean, or nearly so, some other form of energy must exist. Evidence for
this was presented in 1999 (see, for instance, [18]) in the form of observations
of distant exploding stars called Type Ia supernovae. These distant supernovea
are fainter than expected for a universe whose expansion rate is slowing down,
suggesting that the universe is accelerating its expansion. In 2011, Saul
Perlmutter, Brian Schmidt, and Adam Riess won the Nobel Prize in Physics
for their work on this discovery.
When Einstein rst proposed the 3-sphere as the shape of the universe,
his theory predicted that the 3-sphere should be expanding or collapsing. The
idea of a static universe appealed to him, and he added a constant into his
eld equations, called the cosmological constant, whose role was to counteract
gravity and prevent the universe from collapsing in on itself. But at roughly the
same time, Edwin Hubble, Vesto Slipher and others discovered that galaxies
in every direction were receding from us. Moreover, galaxies farther away were
receding at a faster rate, implying that the universe is expanding. Einstein
withdrew his constant.
But now the constant has new life, as it can represent the repulsive dark
energy that seems to be counteracting gravity and driving the accelerated
expansion of the universe. So the density parameter in the Friedmann equation
can have two components: ρM , which is the mass-energy density associated
with ordinary and dark matter (the mass-energy density that cosmologists
have been estimating by observation); and ρΛ , which is the dark energy, due
to the cosmological constant. In this case, Friedmann's equation becomes
8πG k
H2 = (ρM + ρΛ ) − 2
3 a
and dividing by H2 we have
8πG 8πG k
1= ρM + ρΛ − .
3H 2 3H 2 (aH)2
We let
ρM ρΛ −k
ΩM = , ΩΛ = , Ωk = .
ρc ρc (aH)2
So the simple equation
1 = ΩM + ΩΛ + Ωk
1 The Wilkinson Microwave Anisotropy Probe was launched in 2001 to carefully plot the
temperature of the cosmic microwave background radiation.
8.4. Our Universe 201
with ΩΛ ≈ .72 and ΩM ≈ 0.28. As the universe evolves, the values of the
density parameters may change, though the sum will always equal one.
Now is probably as good a time as any to tell you that the fate of the
universe is tied to its mass-energy content and the nature of the dark energy,
which is tied to the the geometry of the universe, which is tied to the topology
of the universe.
If the cosmological constant is zero then the relationship is simple: if Ω>1
so that we live in an elliptic 3-manifolds, the universe will eventually begin to
fall back on itself, ultimately experiencing a big crunch. If Ω = 1, a nite
Euclidean universe will have one of 10 possible shapes (6 if we insist on an
orientable universe) and its expansion rate will asymptotically approach 0, but
it will never begin collapsing. If Ω < 1, the universe will continue to expand
until everything is so spread out we will experience a big chill.
With a nontrivial vacuum energy the situation changes. While the
gravitational force due to the usual mass-energy content of the universe tends
to slow the expansion of the universe, it seems the dark energy causes it to
accelerate. Whether the universe is hyperbolic, elliptic, or Euclidean, if the
dark energy wins the tug of war with gravity, the curvature of the universe
would approach 0 as it continued to accelerate its expansion, and the density
of matter in the universe would approach 0.
The European Space Agency launched the Planck satellite in 2009. In its
orbit at a distance of 1.5 million kilometers from the Earth, the Planck satellite
has given us improved measurements of the CMB temperature, enabling
sharper estimates of cosmological parameters, as well as more rened data
on which to run circles-in-the-sky tests. Alas, no circles have been detected.
Regarding the curvature of the universe, the Planck team concludes in [17], in
agreement with the WMAP team, that our universe appears to be at to a one
standard deviation accuracy of 0.25%.
In short, the universe appears to be homogeneous, isotropic, nearly at, and
dominated by dark energy. The current estimates for Ωk leave the question of
the geometry of the universe open, though just barely. It is still possible that
our universe is a hyperbolic or an elliptic manifold, but the curvature would
have to be very close to 0. If the universe is a compact, orientable Euclidean
manifold, we have six dierent possibilities for its shape. Since Euclidean
manifold volumes aren't xed by curvature, there is no reason to expect the
dimensions of a Euclidean manifold to be close to the radius of the observable
universe. But if the size is right, the circles-in-the-sky method would reveal
the shape of the universe through matching circles.
Perhaps we will be treated to matching circles some day. Perhaps not.
Perhaps the universe is just too big. In any event, pursuing the question of the
shape of the universe is a remarkable feat of the human intellect. It is inspiring
to think, especially looking up at a clear, star lled night sky, that we might
be able to determine the shape of our universe, all without leaving our tiny
planet.
A
List of Symbols
202
References
Textbooks
[1] Anderson, J.W. Hyperbolic Geometry, 2nd Ed. Springer, London, 2005.
[2] Boas, R.P. Invitation to Complex Analysis. Random House, New York,
1987.
[3] Coxeter, H.S.M. Introduction to Geometry, 2nd Ed. Wiley, New York,
1961.
[4] Heath, T.L. The Thirteen Books of Euclid's Elements, 2nd Ed. Dover,
New York, 1956.
[6] Hilbert, D., and H. Cohn-Vossen, Geometry and the Imagination. AMS
Chelsea Publishing, Rhode Island, 1952.
[12] Weeks, J.R. The Shape of Space, 2nd Ed.. Dekker, New York, 2002.
Other Books
[13] Gray J., ed. The Symbolic Universe: Geometry and Physics 1890-1930.
Ofxord University Press, New York, 1999.
[14] Hogan, C.J. The Little Book of the Big Bang. Copernicus Springer-
Verlag, New York, 1998.
[15] Rucker, R. Geometry, Relativity and the 4th Dimension. Dover, New
York, 1977.
Articles
[16] Adams, C. and J. Shapiro. The Shape of the Universe: Ten Possibilities.
American Scientist, Sept - Oct 2001, 89, 443453.
[17] Ade, P. A. R. and others. Planck 2015 results. XIII. Cosmological
parameters. e-Print: arXiv:1502.01589v3 [astro-ph.CO], 2016.
203
204 A. List of Symbols
[19] Cornish, N.J., D.N. Spergel, G.D. Starkman, and E. Komatsu. Con-
straining the Topology of the Universe. Physical Review Letters, (2004),
92 no. 20, 1302.
[20] Foote, R. L. A Unied Pythagorean Theorem in Euclidean, Spherical,
and Hyperbolic Geometries. Math. Magazine, February 2017 90 no. 1,
5969.
[23] Levin, J. Topology and the Cosmic Microwave Background. Phys. Rep.,
2002, 365, 251333.
[24] Luminet, J.-P. and M. Lachieze-Rey. Cosmic Topology. Phys. Rep.,
(1995), 254, 135214.
[25] Luminet, J.-P., J.R. Weeks, A. Riazuelo, R. Lehoucq, R., and J.-P.
Uzan. Dodecahedral space topology as an explanation for weak wide-angle
temperature correlations in the cosmic microwave background. Nature,
(2003), 425, 59395.
[26] Rebouças, M.J., and G.I. Gomero. Cosmic Topology: A Brief Overview.
Phys. Rep., December, 2004, 34 no. 4A, 13581366.
[27] Schwarzschild, K. On the Permissible Curvature of Space. Vierteljahrschrift
d. Astronom. Gesellschaft, (1900), 35, 33747.
[28] Uzan, J.-P., R. Lehoucq, and J.-P. Luminet. New developments in the
search for the topology of the Universe. Nuclear Physics B Proceedings
Supplements, January 2000, 80, C425+.
Web Resources
[29] Joyce, D.E. Euclid's Elements, http://aleph0.clarku.edu/~djoyce/java/
elements/elements.html (1994)
Index
205
206 Index