Math 161 - Notes
Neil Donaldson
Spring 2024
1 Geometry and the Axiomatic Method
1.1 The Early Origins of Geometry: Thales and Pythagoras
We begin with a condensed overview of geometric history. The word geometry comes from the an-
cient Greek geo (Earth), and metros (measure). Measurement (of distance, area, height, angle) had
obvious practical benefits with regard to construction, taxation, commerce and navigation. Astron-
omy provided a related cultural driver of ancient geometry.
Ancient times (pre-500 BC) Egypt, Mesopotamia, China, India: basic rules for measuring lengths,
areas and volumes of simple shapes. Applications: surveying, tax collection, construction,
religious practice, astronomy, navigation. Typically worked examples without general for-
mulæ/abstraction.
Ancient Greece (from c. 600 BC) Philosophers such as Thales and Pythagoras began the process of
abstraction. General statements (theorems) formulated and proofs attempted. Concurrent de-
velopment of early scientific reasoning.
Euclid of Alexandria (c. 300 BC) Collected and expanded earlier work, especially that of the Pytha-
goreans. His compendium the Elements is one of the most important books in Western history
and remained a standard school textbook in to the 1900’s. The Elements is an early exemplar of
the axiomatic method at the heart of modern mathematics.
Later Greek Geometry Archimedes’ (c. 270–212 BC) work on area and volume included techniques
similar to those of modern calculus. Ptolemy (c. AD 100–170) writes the Almagest, a treatise on
astronomy which covers the foundations of trigonometry.
Post-Greek Geometry During the European Dark Ages, geometric understanding was developed
and enhanced by Indian and Islamic mathematicians who particularly developed trigonometry
and algebra.
Analytic Geometry In mid 1600’s France, Descartes and Fermat melded algebra with geometry with
the advent of co-ordinate systems (axes).
Modern Development Non-Euclidean geometries help provide the mathematical foundation for
Einstein’s relativity and the study of curvature. Following Klein (1872), modern geometry is
highly dependent on group theory.
1
Thales of Miletus (c. 624–546 BC) Thales was an olive trader from Miletus, a city-state on the west
coast of modern Turkey. Through trading and travelling, he absorbed mathematical ideas from
nearby cultures including Egypt and Mesopotamia. Here are five results partly attributable to Thales.
1. A circle is bisected by a diameter.
2. The base angles of an isosceles triangle are equal.
3. The pairs of angles formed by two intersecting lines are equal.
4. Two triangles are congruent if they have two angles and the included side equal.
5. An angle inscribed in a semicircle is a right angle.
The last is still known as Thales’ Theorem. Thales’ arguments were not rigorous by modern standards.
His real innovation was to state abstract, general principles: any circle is bisected by any of its diameters.
The Greek word θεωρεω (theoreo), from which we get theorem, has several meanings: ‘to look at,’
‘speculate,’ or ‘consider.’ Thales’ results were supposed to be clear just by looking at a picture.
β
α
α
β
The pictures show Thales’ Theorems 1, 2 and 5. Arguments for Theorems 1 and 2 could be as simple
as ‘fold.’ Theorem 5 follows from the observation that the radius of the circle splits the large triangle
into two isosceles triangles: Theorem 2 says that these have equal base angles (labelled), now check
that α + β is half the angles in a triangle, namely a right-angle.
Pythagoras of Samos (570—495 BC) Pythagoras grew up on Samos, an island in the Aegean Sea
not far from Miletus. He also travelled widely, eventually settling in Croton, southern Italy, around
530 BC where he founded a philosophical school devoted to the study of number, music and ge-
ometry. It has been claimed that the Pythagoreans first classified the regular (Platonic) solids and
developed the musical relationship between the length of a vibrating string and its pitch. While it
is difficult to verify such assertions, the Pythagorean obsession with number and the ‘music of the
universe’ certainly inspired later mathematicians and philosophers—particularly Euclid, Plato and
Aristotle—who believed they were refining and clarifying this earlier work.
Of course, Pythagoras is best known for the result that bears his name.
Theorem 1.1 (Pythagoras). The square on the hypotenuse of a right triangle equals the sum of the
squares on the remaining sides.
Two important clarifications are needed for modern readers.
1. By square, the Greeks meant an honest square! There is no algebra, no numerical lengths, and
the equation a
2
+ b
2
= c
2
won’t be seen for another 2000 years.
2
2. The word equals means equal area, though without a numerical concept of such. The Greeks
meant that the large square can be subdivided into pieces which may be rearranged to produce
the two squares on the remaining sides.
The result suddenly seems less easy! The pictures below provide a simple visualization.
A simple proof of Pythagoras’ Theorem
The problem with this ‘proof is that it relies on subtracting the areas of the four congruent triangles
from a very large square, whereas the Greeks idea of area was essentially additive. Book I of Euclid’s
Elements seems to have been structured precisely to correct this and provide a rigorous constructive
proof. Indeed it is possible (though very ugly!) to apply the 47 results leading up to and including his
proof of Pythagoras, in such as way as to explicitly subdivide the hypotenuse square and rearrange
it into the two smaller squares as required.
Much has been written about Pythagoras’ Theorem, including many, many proofs.
1
It is often
claimed that Pythagoras himself first proved the result, but this is generally considered incorrect—
the ‘proof most often attributed to the Pythagoreans is based on contradictory ideas about numbers
which were debunked by the time of Aristotle. Moreover, other cultures, particularly ancient China,
2
are known to have used the result, at least in example form, several hundred years before Pythago-
ras. Regardless, an argument over attribution is fruitless without first agreeing on what constitutes a
proof. This means that we need to spend some time considering Axiomatic Systems. . .
Exercises 1.1. 1. Let the side-lengths of the above triangles be a, b, c. Can you rephrase the proof
algebraically?
2. A theorem of Euclid states:
The square on the parts equals the sum of the squares on each part plus twice the
rectangle on the parts
By referencing the above picture, state Euclid’s result using modern algebra.
(Hint: let a and b be the ‘parts’. . . )
1
Including a proof by former US President James Garfield. Would that current presidents were so learned. . .
2
In China, Pythagoras’ Theorem is known as the gou gu, which refers to the two non-hypotenuse sides of the triangle.
3
1.2 Axiomatic Systems
Arguably the most revolutionary aspect of the Euclid’s Elements was its axiomatic presentation.
Definition 1.2. An axiomatic system comprises four types of object.
1. Undefined terms: Concepts accepted without definition/explanation.
2. Axioms: Logical statements regarding the undefined terms which are accepted without proof.
3. Defined terms: Concepts defined in terms of 1 & 2.
4. Theorems: Logical statements deduced from 1–3.
Examples 1.3. Here are two systems viewed informally in this framework. In each case we provide
only examples of each type of object, not a full description of an axiomatic system.
Basic Geometry 1. Line and point.
2. There exists a line joining any two given points.
3. A triangle may be defined in using three non-collinear points.
4. Thales’ and Pythagoras’ Theorems.
Chess 1. Pieces (as black/white objects) and the board.
2. Rules for how each piece moves.
3. Concepts such as check, stalemate or en-passant.
4. For example, Given a particular position, Black can win in 5 moves.
A proof is a logical argument demonstrating the truth of a theorem within an axiomatic system. In
practice, this is an ideal to which we aspire, and a proof is simply a convincing logical argument.
Definition 1.4. A model is a choice/definition of the undefined terms such that all axioms are true.
Models are often abstract in that they depend on another axiomatic system. In a concrete model, the
undefined terms are real-world objects (where contradictions are impossible(!)). The big idea is this:
Any theorem proved within an axiomatic system is true in any model of that system.
Mathematical discoveries often hinge on the realization that seemingly separate discussions can be
described in terms of models of a common axiomatic system.
Example 1.5. Monoid: If you’ve studied group theory, this should seem familiar.
1. A set G and a binary operation .
2. (A1) Closure: a, b G, a b G
(A2) Associativity: a, b, c G, a (b c) = (a b) c
(A3) Identity: e G such that a G, a e = e a = a
3. Concepts such as square a
2
= a a, or commutativity a b = b a.
4. For example, The identity is unique.
(G, ) = (Z, +) is an abstract model, where e = 0. If you really want a concrete model, consider a
single dot on the page, equipped with the operation = !
4
Definition 1.6. Certain properties are desirable in an axiomatic system.
Consistency The system is free of contradictions.
Independence An axiom is independent if it is not a theorem of the others. An axiomatic system is
independent if all its axioms are.
Completeness Every valid proposition within the theory is decidable: can be proved or disproved.
We unpack these ideas slightly. By necessity, our descriptions are vague. Many notions need to be
clarified (e.g., what is meant by a valid proposition) before these ideas can be made rigorous.
Consistency May be demonstrated by exhibiting a concrete model. An abstract model demonstrates
relative consistency, dependent on the consistency of the underlying system. An inconsistent
system is essentially useless.
Independence To demonstrate the independence of an axiom, exhibit two models: one in which all
axioms are true, the other in which only the considered axiom is false.
Completeness This is very unlikely to hold for most useful axiomatic systems in mathematics, though
examples do exist. To show incompleteness, an undecidable
3
statement is required, which can
be viewed as a new independent axiom of an enlarged system.
Example (1.5, cont). The axiomatic system for a monoid is:
Consistent We have a (concrete) model.
Independent Consider three models:
( N, +) satisfies axioms A1 and A2 but not A3.
({e, a, b}, ) defined by the following table satisfies A1 and A3 but not A2
e a b
e e a b
a a e a
b b b a
e.g. a (b b) = a a = e = a = a b = (a b) b
( Z \ {1}, +) satisfies axioms A2 and A3 but not A1.
Incomplete The proposition A monoid contains at least two elements is undecidable just from the axioms.
For instance, ({0}, +) and (Z, +) are models with one/infinitely many elements.
We could also ask if all elements have an inverse. That this is undecidable is the same as saying
that a new axiom is independent of A1, A2, A3.
(A4) Inverse: g G, g
1
G such that g g
1
= g
1
g = e.
The new system defined by the four axioms is also consistent and independent—this is the
structure of a group. Even this new system is incomplete; for instance, consider a new axiom of
commutativity. . .
3
A famous example of an undecidable statement from standard set theory is the Continuum Hypothesis, which states that
there is no uncountable set with cardinality strictly smaller than that of the real numbers.
5
Example 1.7 (Bus Routes). Here is a loosely defined axiomatic system. Discuss the questions with
your classmates.
Undefined Terms: Route, Stop
Axioms: (A1) Each route is a list of stops in some order. These are the stops visited by the route.
(A2) Each route visits at least four distinct stops.
(A3) No route visits the same stop twice, except the first stop which is also the last stop.
(A4) There is a stop called downtown that is visited by each route.
(A5) Every stop other than downtown is visited by at most two routes.
1. Construct a model of this system with three routes. What is the fewest number of stops you
can use?
2. Your answer to 1 shows that this system is: complete, consistent, inconsistent, independent?
3. Is the following a model for the Bus Routes system? If not, determine which axioms are satisfied
by the model and which are not?
Stops: Downtown, Walmart, Albertsons, Main St., CVS, Trader Joe’s, Zoo
Route 1: Downtown, Walmart, Main St., CVS, Zoo, Downtown
Route 2: Main St., CVS, Zoo, Albertsons, Downtown, Main St.
Route 3: Walmart, Main St., Downtown, Albertsons, Main St., Walmart
4. Show that A3 is independent of the other axioms.
5. Demonstrate that There are exactly three routes is not a theorem in this system by finding a
model in which it is not true.
We are only scratching the surface of axiomatics. If you really want to dive down the rabbit hole,
consider taking a class in formal logic or model theory. As an example of the ideas involved, we
finish with two results proved in 1931 by the German logician Kurt G
¨
odel.
Theorem 1.8 (G¨odel’s incompleteness theorems).
1. Any consistent system containing the natural numbers is incomplete.
2. The consistency of such a system cannot be proved within the system itself.
G
¨
odel’s first theorem tells us that there is no ultimate consistent complete axiomatic system. Perhaps
this is reassuring—there will always be undecidable statements, so mathematics will never be fin-
ished! However, the undecidable statements cooked up by G
¨
odel are analogues of the famous liar
paradox (‘This sentence is false’), so the profundity of this is a matter of debate.
G
¨
odel’s second theorem fleshes out the difficulty in proving the consistency of an axiomatic system.
If a system is sufficiently complex to describe the natural numbers, its consistency can at best be
proved relative to some other axiomatic system. While an inconsistent system might be essentially
useless, good luck showing that what you have really is consistent!
6
Exercises 1.2. 1. Between two players are placed several piles of coins. On each turn a player takes
as many coins as they want from one pile, as long as they take at least one coin. The player who
takes the last coin wins.
If there are two piles where one pile has more coins than the other, prove that the first player
can always win the game.
2. Consider a system where children in a classroom choose different flavors of ice cream. Suppose
we have the following axioms:
(A1) There are exactly five flavors of ice cream: vanilla, chocolate, strawberry, cookie
dough, and bubble gum.
(A2) Given any two distinct flavors, there is exactly one child who likes these.
(A3) Every child likes exactly two flavors of ice cream.
(a) How many children are in the classroom? Prove your assertion.
(b) Prove that any pair of children likes at most one common flavor.
(c) Prove that for each flavor, there are exactly four children who like that flavor.
3. Consider an axiomatic system that consists of elements in a set S and a set P of pairings of
elements (a, b) that satisfy the following axioms:
(A1) If (a, b) is in P, then (b, a) is not in P.
(A2) If (a, b) is in P and (b, c) is in P, then (a, c) is in P.
(a) Let S = {1, 2, 3, 4} and P = {(1, 2), (2, 3), (1, 3)}. Is this a model for the axiomatic system?
Why/why not?
(b) Let S be the set of real numbers and let P consist of all pairs (x, y) where x < y. Is this a
model for the system? Explain.
(c) Use the results of (a) and (b) to argue that the axiomatic system is incomplete. I.e., think
of another independent axiom that could be added to the axioms A1 and A2 for which S
and P in part (a) is a model, but for which S and P from part (b) is not a model.
4. The undefined terms of an axiomatic system are ‘brewery’ and ‘beer’. Here are some axioms.
(A1) Every brewery is a non-empty collection of at least two beers (every brewery
brews at least two beers!).
(A2) Any two distinct breweries have at most one beer in common.
(A3) Every beer belongs to exactly three breweries.
(A4) There exist exactly six breweries.
(a) Prove the following theorems.
i. There are exactly four beers.
ii. There are exactly two beers in each brewery.
iii. For each brewery, there is exactly one other brewery which has no beers in common.
(b) Prove that the axioms are independent.
(When negating A1, you should assume that a brewery is still a collection of beers, but that any
such could contain none or one beer)
7
2 Euclidean Geometry
2.1 Euclid’s Postulates and Book I of the Elements
Euclid’s Elements (c. 300 BC) formed a core part of European and Arabic curricula until the mid 20
th
century. Several examples are shown below.
Earliest Fragment c. AD 100 Full copy, Vatican, 9
th
C Pop-up edition, 1500s
Latin translation, 1572 Color edition, 1847 Textbook, 1903
Many of Euclid’s arguments can be found online, and you can read Byrne’s 1847 edition here: the
cover is Euclid’s proof of Pythagoras’. We present an overview of Book I.
Undefined Terms E.g., point, line, etc.
4
Axioms/Postulates
5
A1 If two objects equal a third, then the objects are equal (= is transitive)
A2 If equals are added to equals, the results are equal (a = c & b = d = a + b = c + d)
A3 If equals are subtracted from equals, the results are equal
A4 Things that coincide are equal (in magnitude)
A5 The whole is greater than the part
P1 A pair of points may be joined to create a line
P2 A line may be extended
P3 Given a center and a radius, a circle may be drawn
P4 All right-angles are equal
P5 If a straight line crosses two others and the angles on one side sum to less than two right-
angles then the two lines (when extended) meet on that side.
4
In fact Euclid attempted to define these: ‘A point is that which has no part,’ and ‘A line has length but no breadth.’
5
In Euclid, an axiom is considered somewhat more general than a postulate. Here the postulates contain the geometry.
8
The first three postulates describe the intuitive ruler and compass constructions. P4 allows Euclid to
compare angles at different locations. P5 is usually known as the parallel postulate.
Euclid’s system doesn’t quite fit the modern standard. Some axioms are vague (what are ‘things’?)
and we’ll consider several more-serious shortcomings later. For now we clarify two issues and intro-
duce some notation.
Segments To Euclid, a line had finite extent—we call such a (line) segment. The segment joining points
A, B is denoted AB. In modern geometry, a line extends as far as is permitted.
Congruence Euclid uses equal where modern mathematicians say congruent. We’ll express, say, con-
gruent angles as ABC
=
DEF rather than ABC = DEF.
Basic Theorems `a la Euclid
Theorems were typically presented as a problem. Euclid first provides a construction (P1–P3) before
proving that his construction solves the problem.
Theorem 2.1 (I. 1). Problem: to construct an equilateral triangle on a given segment.
The labelling I. 1 indicates Book I, Theorem 1.
Proof. Given a line segment AB:
By P3, construct circles centered at A and B with radius AB.
Call one of the intersection points C. By P1, construct AC and BC.
We claim that ABC is equilateral.
Observe that AB and AC are radii of the circle centered at A, while
AB and BC are radii of the circle centered at B. By Axiom A1, the
three sides of ABC are congruent.
A
B
C
Euclid proceeds to develop several well-known constructions and properties of triangles.
(I. 4) Side-angle-side (SAS) congruence: if two triangles have two pairs of congruent sides and
the angles between these are congruent, then the remaining sides and angles are congruent in
pairs.
AB
=
DE
ABC
=
DEF
BC
=
EF
=
AC
=
DF
BCA
=
EFD
CAB
=
FDE
(I. 5) An isosceles triangle has congruent base angles.
(I. 9) To bisect an angle.
(I. 10) To find the midpoint of a segment.
(I. 15) If two lines/segments cut one another, opposite angles are congruent.
Have a look at some of Euclid’s arguments online. These are worth reading despite there being logical
issues with Euclid’s presentation. We’ll revisit these results in the Exercises and next two sections.
9
Parallel Lines: Construction & Existence
Definition 2.2. Lines are parallel if they do not intersect. Segments are parallel if no extensions of
them intersect.
In Euclid, a line is not parallel to itself. The next result is one of the most important in Euclidean
geometry, in that it describes how to create a parallel line through a given point.
Theorem 2.3 (I. 16 Exterior Angle Theorem). If one side of a
triangle is extended, then the exterior angle is larger than either
of the opposite interior angles.
In the picture, we have δ > α and δ > β.
α
β
δ
Euclid did not quantify angles numerically: δ > α means that α is congruent to some angle inside δ.
Proof. Construct the bisector BM of AC (I. 10).
Extend BM to E such that BM
=
ME (I. 2) and connect CE (P1).
The opposite angles at M are congruent (I. 15).
SAS (I. 4) applied to AMB and CME says BAM
=
EMC,
which is clearly smaller than the exterior angle at C.
A
B
C
M
E
Bisect BC and repeat the argument to see that β < δ.
The proof in fact constructs a parallel (CE) to AB through C, as the next result shows.
Theorem 2.4 (I. 27). If a line falls on two other lines such that the
alternate angles (α, β) are congruent, then the two lines are parallel.
α
β
The alternate angles in the exterior angle theorem are those at A and C: CE really is parallel to AB.
Proof. If the lines were not parallel, they would meet on one
side. WLOG suppose they meet on the right side at C.
The angle β at B, being exterior to ABC, must be greater than
the angle α at A (I. 16): contradiction.
α
β
A
B
C
Euclid combines this with the vertical angles theorem (I. 15) to finish the first half of Book I.
Corollary 2.5 (I. 28). If a line falling on two other lines makes con-
gruent angles, then the two lines are parallel.
Thus far, Euclid uses only postulates P1–P4. In any model in which these hold:
Given a line and a point C not on , there exists a parallel to through C
10
Parallel Lines: Uniqueness, Angle-sums & Playfairs Postulate
Euclid finally invokes the parallel postulate to prove the converse of I. 27, showing that the congruent
alternate angle approach is the only way to have parallel lines.
Theorem 2.6 (I. 29). If a line falls on two parallel lines, then the alternate angles are congruent.
Proof. Given the picture, we must prove that α
=
β.
Suppose not and WLOG that α > β.
But then β + γ < α + γ, which is a straight edge.
By the parallel postulate, the lines , m meet on the left side of
the picture, whence and m are not parallel.
m
α
γ
β
n
The most well-known result about triangles is now in our grasp, that the interior angles sum to a
straight edge. Euclid words this slightly differently.
Theorem 2.7 (I. 32). If one side of a triangle is extended, the exterior angle is congruent to the sum
of the opposite interior angles.
This is not a numerical sum, though for familiarity’s sake we’ll
often write 180° for a straight edge and 90° for a right-angle.
In the picture we’ve labelled angles as Greek letters for clarity.
The result amounts to showing that
e
α +
e
β
=
α + β.
A
B
C
D
E
β
˜
β
α
˜
α
Proof. Construct CE parallel to BA as in I. 16, so that
e
α
=
α.
BD falls on parallel lines AB and CE, whence
e
β
=
β (Corollary of I. 29).
Axiom A2 shows that ACD =
e
α +
e
β
=
α + β.
The parallel postulate is stated in the negative (angles don’t sum to a straight edge, therefore lines are
not parallel). Though we cannot be sure, Euclid possibly chose this formulation in order to facilitate
proofs by contradiction. Unfortunately the effect is to obscure the meaning of the parallel postulate.
Here is a more modern interpretation.
Axiom 2.8 (Playfairs Postulate). Given a line and a point C
not on , at most one parallel line m to passes through C.
m
C
Our discussion up to now shows that the parallel postulate
implies Playfair.
Let A, B and construct the triangle ABC.
The exterior angle theorem constructs E and thus a par-
allel m to by I. 27.
I. 29 invokes the parallel postulate to prove that this is
the only such parallel.
m
B
A
C
E
11
In fact the postulates are equivalent.
Theorem 2.9. In the presence of Euclid’s first four postulates, Playfair’s postulate and the parallel
postulate (P5) are equivalent.
Proof. (P5 Playfair) We proved this above.
(Playfair P5) We prove the contrapositive. Assume postulates P1–P4 are true and that P5 is false.
Using quantifiers, and with reference to the picture in I. 29, we restate the parallel postulate:
P5: pairs of lines , m and crossing lines n, β + γ < 180° = , m not parallel.
Its negation (P5 false) is therefore:
parallel lines , m and a crossing line n for
which β + γ < 180°
This is without loss of generality: if β + γ > 180°, consider
the angles on the other side of n.
By the the exterior angle theorem/I. 28, we may build a
parallel line
ˆ
to through the intersection C of m and n
(in the picture,
ˆ
β
=
β). Crucially, this only requires postu-
lates P1–P4!
m
γ
β
n
ˆ
m
C
ˆ
β
γ
β
Observe that
ˆ
and m are distinct since
ˆ
β + γ
=
β + γ < 180°. We therefore have a line and a
point C not on , though which pass (at least) two parallels to : Playfair’s postulate is false.
Non-Euclidean Geometry
That Euclid waited so long before invoking the uniqueness of parallels suggests he was trying to
establish as much as he could about triangles and basic geometry in its absence. By contrast, every-
thing from I. 29 onwards relies on the parallel postulate, including the proof that the angle sum in a
triangle is 180°. For centuries, many mathematicians believed, though none could prove it, that such
a fundamental fact about triangles must be true independent of the parallel postulate.
Loosely speaking, a non-Euclidean geometry is a model for which a parallel through an off-line point
either doesn’t exist or is non-unique. It wasn’t until the 17–1800s and the development of hyperbolic
geometry (Chapter 4) that a model was found in which Euclid’s first four postulates hold but for which
the parallel postulate is false.
6
We shall eventually see that every triangle in hyperbolic geometry has angle
sum less than 180°, though this will require a lot of work! For a more eas-
ily visualized non-Euclidean geometry consider the sphere. A rubber band
stretched between three points on its surface describes a spherical triangle: an
example with angle sum 270° is drawn. A similar game can be played on a
saddle-shaped surface: as in hyperbolic geometry, ‘triangles’ will have angle
sum less than 180°.
6
This shows that the parallel postulate is independent; in fact all Euclid’s postulates are independent. They are also
consistent (the ‘usual’ points and lines in the plane are a model), but incomplete: a sample undecidable is in Exercise 5.
12
Pythagoras’ Theorem
Following his discussion of parallels, Euclid shows that parallelograms with the same base and
height are equal (in area) (I. 33–41), before providing constructions of parallelograms and squares
(I. 42–46). Some of this is in Exercise 2. Immediately afterwards comes the capstone of Book I.
Theorem 2.10 (I. 47 Pythagoras’ Theorem). The square on the hypotenuse of a right triangle equals
(has the same area as) the sum of the squares on the other sides.
Proof. The given triangle ABC is assumed to have a right-angle at A.
1. Construct squares on each side of ABC (I. 46) and a
parallel AL to BD (I. 16).
2. AB
=
FB and BD
=
BC since sides of squares are con-
gruent. Moreover ABD
=
FBC since both contain
ABC and a right-angle.
3. Side-angle-side (I. 4) says that ABD and FBC are
congruent (identical up to rotation by 90°).
4. I. 41 compares areas of parallelograms and triangles
with the same base and height:
Area(ABFG) = 2 Area(FBC)
= 2 Area(ABD)
= Area
BOLD
5. Similarly Area(ACKH) = Area
OCEL
.
A
B
CO
LD E
F
G
H
K
6. Sum the rectangles to obtain BCED and complete the proof.
Euclid finishes Book I with the converse, which we state without proof. Euclid’s argument is very
sneaky—look it up!
Theorem 2.11 (I. 48). If the (areas of the) squares on two sides of a triangle equal the (area of the)
square on the third side, then the triangle has a right-angle opposite the third side.
The Elements contains thirteen books. Much of the remaining twelve discuss further geometric con-
structions, including in three dimensions. There is also a healthy dose of basic number theory includ-
ing what is now known as the Euclidean algorithm.
While undoubtedly a masterpiece of logical reasoning, Euclid’s presentation has several flaws. Most
problematic is his reliance on pictorial reasoning: for instance, he ‘proves’ the SAS and SSS congru-
ence theorems (I. 4 & 8) by laying one triangle on top of another, a process not justified by his axioms
(look it up online or Byrne). In a modern sense, Euclid’s approach is part axiomatic system and part
model: his reasoning requires a visual/physical representation of lines, circles, etc. Because of these
issues, we now turn to a more modern description of Euclidean geometry, courtesy of David Hilbert.
13
Exercises 2.1. 1. (a) Prove the vertical angle theorem (I. 15): if two lines cut one another, opposite
angles are congruent.
(Hint: This is one place where you will need to use postulate 4 regarding right-angles)
(b) Use part (a) to complete the proof of the exterior angle theorem: i.e., explain why β < δ.
2. To help prove Pythagoras’, Euclid makes use of the following results. Prove them as best as
you can. Full rigor is tricky, but the pictures should help!
(a) (I. 11) At a given point on a line, to construct a perpendicular.
(b) (I. 46) To construct a square on a given segment.
(c) (I. 35) Parallelograms on the same base and with the same height have equal area.
(d) (I. 41) A parallelogram has twice the area of a triangle on the same base and with the same
height.
A
B
D
C
A
B
C
D
A
B
C
D EF
Theorem I. 11 Theorem I. 46 Theorem I. 35
3. Consider spherical geometry (page 12), where lines are paths of shortest distance (great circles).
(a) Which of Euclid’s postulates P1–P5 are satisfied by this geometry?
(b) (Hard) Where does the proof of the exterior angle theorem fail in spherical geometry?
4. (a) State the negation of Playfair’s postulate.
(b) Prove that Playfair’s postulate is equivalent to the following statement:
Whenever a line is perpendicular to one of two parallel lines, it must be perpen-
dicular to the other.
5. The line-circle continuity property states:
If point P lies inside and Q lies outside a circle α, then the segment PQ intersects α.
By considering the set of rational points in the plane Q
2
= {(x, y) : x, y Q}, and making a
sensible definition of line and circle, show that the line-circle continuity property is undecidable
within Euclid’s system.
6. The standard proof of the converse of Pythagoras’ theorem (I. 48) is, in fact, a corollary of the
original! Look it up and explain the argument as best you can.
14
2.2 Hilbert’s Axioms I: Incidence and Order
The long process of identifying and correcting the errors and omissions in Euclid’s Elements culmi-
nated in the 1899 publication of David Hilbert’s Grundlagen der Geometrie (Foundations of Geometry).
In the next two sections we consider some of the details of Hilbert’s approach, providing a modern
and logically superior description of Euclidean geometry.
Hilbert’s axioms for plane geometry
7
are listed on the next page. The undefined terms consist of two
types of object (points and lines), and three relations (between , on and congruence
=
). For brevity
we’ll often use/abuse set notation, viewing a line as a set of points, though this is not necessary. At
various places, definitions and notations are required.
Definition 2.12. Throughout, A, B, C denote points and , m lines.
Line:
AB denotes the line through distinct A, B. This exists and is unique by axioms I-1 and I-2.
Segment: AB := {A, B} {C : A C B} consists of distinct endpoints A, B and all interior points C
lying between them.
Ray:
AB := AB {C : A B C} is a ray with vertex A. In essence we extend AB beyond B.
Triangle: ABC := AB BC CA where A, B, C are non-collinear. Triangles are congruent if their
sides and angles are congruent in pairs.
Sidedness: Distinct A, B, not on , lie on the same side of if AB = . Otherwise A and B lie on
opposite sides of .
Angle: BAC :=
AB
AC has vertex A and sides
AB and
AC.
Parallelism: Lines and m intersect if there exists a point lying on both: A m. Lines are parallel
if they do not intersect. Segments/rays are parallel when the corresponding lines are parallel.
The pictures represent these notions in the usual model of Cartesian geometry.
A
B
A
B
A
B
C
Line
AB Segment AB Ray
AB Triangle ABC
A
B
A
B
A
B
C
A
m
Same side Opposite sides Angle BAC Intersection A m
7
Like Euclid, Hilbert also covered 3D geometry—we only give the axioms for plane geometry. With regard to our
desired properties (Definition 1.6), his system is about as good as can be hoped. Essentially one only one model exists,
which is almost the same thing as completeness. In the absence of the continuity axiom, the axioms are consistent; in
line with G
¨
odel’s theorems (1.8), consistency cannot be proved once continuity is included. As stated, the axioms are not
quite independent, though this can be remedied: O-3 does not require existence (follows from Pasch’s axiom), C-1 does
not require uniqueness (follows from uniqueness in C-4) and C-6 can be weakened slightly.
15
Hilbert’s Axioms for Plane Geometry
Undefined terms
1. Points: use capital letters, A, B, C . . .
2. Lines: use lower case letters, , m, n, . . .
3. On: A is read A lies on
4. Between: A B C is read B lies between
A and C
5. Congruence:
=
is a binary relation on
segments or angles
Axioms of Incidence
I-1 For any distinct A, B there exists a line
on which lie A, B.
I-2 There is at most one line through distinct
A, B (A and B both on the line).
Notation: line
AB through A and B
I-3 On every line there exist at least two dis-
tinct points. There exist at least three
points not all on the same line.
Axioms of Order
O-1 If A B C, then A, B, C are distinct
points on the same line and C B A.
O-2 Given distinct A, B, there is at least one
point C such that A B C.
O-3 If A, B, C are distinct points on the same
line, exactly one lies between the others.
Definitions: segment AB and triangle ABC
O-4 (Pasch’s Axiom) Let ABC be a triangle
and a line not containing any of A, B, C.
If contains a point of the segment AB,
then it also contains a point of either AC
or BC.
Definitions: sides of line
AB and ray
AB
Axioms of Congruence
C-1 (Segment transference) Let A, B be distinct
and r a ray based at A
. Then there exists a
unique point B
r for which AB
=
A
B
.
Moreover AB
=
BA.
C-2 If AB
=
EF and CD
=
EF, then AB
=
CD.
C-3 If A B C, A
B
C
, AB
=
A
B
and
BC
=
B
C
, then AC
=
A
C
.
Definitions: angle ABC
C-4 (Angle transference) Given BAC and
A
B
, there exists a unique ray
A
C
on
a given side of
A
B
for which BAC
=
B
A
C
.
C-5 If ABC
=
GHI and DEF
=
GHI,
then ABC
=
DEF. Moreover, ABC
=
CBA.
C-6 (Side-angle-side) Given triangles ABC
and A
B
C
, if AB
=
A
B
, AC
=
A
C
,
and BAC
=
B
A
C
, then the triangles
are congruent.
8
Axiom of Continuity
Suppose a line/segment is partitioned into non-
empty subsets Σ
1
, Σ
2
such that no point of Σ
1
lies
between two points of Σ
2
and vice versa.
Then there exists a unique point O satisfying
P
1
O P
2
, if and only if O = P
1
, O = P
2
and
one of P
1
or P
2
lies in Σ
1
and the other in Σ
2
.
Playfairs Axiom
Definition: parallel lines
Given a line and a point P / , at most one line
through P is parallel to .
8
Its sides/angles are congruent in pairs. We extend congruence to other geometric objects similarly.
16
Axioms of Incidence: Finite Geometries
The axioms of incidence describe the relation on. An incidence geometry is any model satisfying axioms
I-1, I-2, I-3. Perhaps surprisingly, there exist incidence geometries with finitely many points!
Examples 2.13. By I-3, an incidence geometry requires at least three points.
A 3-point geometry exists, and is unique up to relabelling:
I-3 says the points A, B, C must be non-collinear. By I-1 and
I-2, each pair lies on a unique line, whence there are precisely
three lines
= {A, B}, m = {A, C}, n = {B, C}
Up to relabelling, there are two incidence geometries with four
points: one is drawn; how many lines has the other?
m
n
A
B
C
A
B
C
D
3 points, 3 lines 4 points, 6 lines
The final picture is a seven-point incidence geometry called the Fano
plane, which finds many applications particularly in combinatorics. Each
point lies on precisely three lines and each line contains precisely three
points—each dot is colored to indicate the lines to which it belongs.
Don’t be fooled by the black line looking ‘curved’ and seeming to cross
the blue line near the top, for the line really only contains three points!
We can even prove some simple theorems in incidence geometry.
Lemma 2.14. If distinct lines intersect, then they do so in exactly one point.
Proof. Suppose A, B are distinct points of intersection. By axiom I-2, there is at most one line through
A and B. Contradiction.
Lemma 2.15. Given any point, there exist at least two lines on which it lies.
The proof is an exercise. While incidence geometry is fun, our main goal is to understand Euclidean
geometry, so we move on to the next set of axioms.
Axioms of Order: Sides of a Line, Pasch’s Axiom & the Crossbar Theorem
The axioms of order describe the ternary relation between. Their inclusion in Hilbert’s axioms is due
in no small part to the work of Moritz Pasch, after whom Pasch’s axiom (O-4, c. 1882) is named. This
axiom is very powerful; in particular, it permits us to define the interiors of several geometric objects,
and to see that these are non-empty.
Lemma 2.16. Every segment contains an interior point.
We leave the proof to Exercise 5. By inducting on the Lemma, every segment contains infinitely many
points, whence the above finite geometries are not valid models once the order axioms are included.
17
To get much further, it is necessary to establish that a line has precisely two sides (Definition 2.12). This
concept lies behind several of Euclid’s arguments, without being properly defined in the Elements.
Theorem 2.17 (Plane Separation). A line separates all points not on into two half-planes: the two
sides of . To be explicit, suppose none of the points A, B, C lie on , then:
1. If A, B lie on the same side of and B, C lie on the same side, then A, C lie on the same side.
2. If A, B lie on opposite sides and B, C lie on opposite sides, then A, C lie on the same side.
3. If A, B lie on opposite sides and B, C lie on the same side, then A, C lie on opposite sides.
A
B
C
A
B
C
A
B
C
Case 1 Case 2 Case 3
Proof. We prove the contrapositive of case 1. Suppose A, B, C are non-collinear. If AC intersects ,
then intersects one side of ABC. By Pasch’s axiom, it also intersects either AB or BC.
The other cases are exercises, and we omit the tedious collinear possibilities.
Plane separation/sidedness allows us to properly define interiors of angles and triangles.
Definition 2.18. A point I is interior to angle BAC if:
I lies on the same side of
AB as C, and,
I lies on the same side of
AC as B.
Otherwise said, I lies in the intersection of two half-planes.
A
B
C
I
A point I is interior to triangle ABC if it is interior to all three of its angles ABC, BAC and
ACB. Otherwise said, I lies in the triple intersection of three of the half-planes defined by the
triangle’s sides.
Interior points permit us to compare angles: if I is interior to BAC, then BAI < BAC has obvious
meaning without resorting to numerical angle measure.
Corollary 2.19. Every angle has an interior point.
Proof. Given BAC, consider any interior point I of the segment BC. This plainly lies on the same
side of
AB as C and on the same side of
AC as B.
In Exercise 8, we check that the interior of a triangle is non-empty.
Pasch’s axiom could be paraphrased: If a line enters a triangle, it must come out. We haven’t quite
established this crucial fact, however. What if the line passes through a vertex?
18
Theorem 2.20 (Crossbar Theorem). Suppose I is interior to BAC.
Then
AI intersects BC.
In particular, if a line passes through a vertex and an interior point of
a triangle, then it intersects the side opposite the vertex.
A
B
C
Proof. Extend AB to a point D such that A lies between B and D (O-2). Since C is not on
BD =
AB we
have a triangle BCD. Since
AI intersects one edge of BCD at A and does not cross any vertices
(think about why. . . ), Pasch says it intersects one of the other edges (BC or CD) at some point M.
The result follows from applying plane separation to the lines
AB =
BD and
AC. First observe:
Since I, M lie on the same side of
AB =
BD as C, it follows that IM does not intersect
AB.
Since A, I, M are collinear and A
AB, it follows that A / I M.
If M BC, we are done. Our goal is to show that M CD is a contradiction.
A
B
C
I
D
M
A
B
C
I
D
M
correct arrangement
9
contradiction
Suppose, for contradiction, that M CD. Relative to
AC:
I and B lie on the same side since I is interior to BAC;
B and D lie on opposite sides, since B A D and
AC =
BD = AB;
D and M lie on the same side since M CD and
CD =
AC.
By plane separation, I, M lie on opposite sides of
AC. The collinearity of A, I, M then forces the
contradiction A IM.
Euclid repeatedly uses the crossbar theorem without justification,
including in his construction of perpendiculars and angle/segment
bisectors (Theorems I. 9+10). We sketch the latter here.
Given BAC, construct E such that AB
=
AE. Construct D using
an equilateral triangle (I. 1). SSS (I. 8) shows that BAC is bisected,
and SAS (I. 4) that
AD bisects BE.
Quite apart from Euclid’s arguments for SAS and SSS being suspect
(we’ll deal with these in the next section), he gives no argument for
why D is interior to BAC or why
AD should intersect BE!
A
B
C
D
E
Even with Pasch’s axiom and the crossbar theorem, it requires some effort to repair Euclid’s proof. No
matter, we’ll provide an alternative construction of the bisector once we’ve considered congruence.
9
The pictures could be modified: e.g., I = M and A I M are also correct arrangements (M BC).
19
Exercises 2.2. 1. Label the vertices in the Fano plane 1 through 7 (any way you like). As we did in
Example 2.13 for the 3-point geometry, describe each line in terms of its points.
2. Prove Lemma 2.15.
3. Give a model for each of the 5-point incidence geometries. How many are there?
(Hint: remember that order doesn’t matter, so the only issue is how many points lie on each line)
4. Consider the proof of the crossbar theorem. Explain how we know that
AI does not contain
any of the vertices of BCD.
5. You are given distinct points A, B. Using the axioms of incidence and order and Lemma 2.14
(follows from I-2), show the existence of each of the points C, D, E, F in the picture in alphabetical
order. Hence conclude the existence of a point F lying between A and B (Lemma 2.16).
During your construction, address the following issues:
(a) Explain why D does not lie on
AB.
(b) Explain why E does not lie on ABD.
(c) Explain why E = C (whence
CE exists).
(d) Explain why F lies on AB and not on BD.
A
B
C
D
E
F
6. We complete the proof of the plane separation theorem (2.17).
(a) Prove part 3 (it is almost a verbatim application of Pasch’s axiom).
(b) Suppose a line intersects all three sides of ABC but no vertices.
This results in a very strange picture (we’ve labelled the intersec-
tions D, E, F and WLOG chosen D E F).
Apply Pasch’s axiom to DBF and
AC to obtain a contradiction.
Hence establish part 2 of the plane separation theorem.
A
B
C
D
E
F
7. Suppose A, B, C are distinct points on a line .
(a) Explain why there exists a line m = such that B m.
(b) Prove that A B C A and C lie on opposite sides of m.
(c) Suppose A B C. Use part (b) to prove the following:
i. B is the only point common to the rays
BA and
BC.
ii. If D is any point other than B, prove that D lies in precisely one of
BA or
BC.
8. Prove that the interior of a triangle is non-empty.
(Hint: use Exercise 5 to construct a suitable I, then prove that it lies on the correct side of each edge)
20
2.3 Hilbert’s Axioms II: Congruence
Hilbert’s congruence axioms address two primary issues in Euclid.
1. Euclid’s use of equal is confusing. In Hilbert, segments/angles are now equal only when they
are precisely the same (this amounts to the reflexivity part of the next result).
2. Euclid’s frequent and unjustified use of pictorial reasoning. We previously discussed Euclid’s
erroneous approach to the SAS and SSS triangle congruence theorems. It was eventually real-
ized that one of the triangle congruences has to be an axiom: SAS is Hilbert’s C-6.
We start with a small piece of bookkeeping.
Lemma 2.21. Congruence of segments/angles is an equivalence relation.
Proof. (Reflexivity) Let AB be given. Apply C-1 to obtain A
B
such that AB
=
A
B
. We sneakily use
this twice and apply C-2 to obtain
AB
=
A
B
and AB
=
A
B
= AB
=
AB
(Symmetry) Assume AB
=
CD. By reflexivity, CD
=
CD. By C-2 we have CD
=
AB.
(Transitivity) Suppose AB
=
CD and CD
=
EF. By symmetry, EF
=
CD. Axiom C-2 now shows that
AB
=
EF.
Axioms C-4 and C-5 say essentially the same thing for angles (see Exercise 2).
Segment/Angle Transfer and Comparison
Hilbert’s axioms of segment and angle transference are crucial for comparing non-collinear segments
and angles with distinct vertices.
Definition 2.22. Let segments AB and CD be given.
By axiom C-1, let E be the unique point on
CD such that CE
=
AB:
we have transferred AB onto
CD.
We write AB < CD if E lies between C and D, etc.
C
D
A
B
E
By O-3, any two segments are comparable: given AB & CD, precisely one of the following holds,
AB < CD, CD < AB, AB
=
CD
C-3 says that congruence respects the ‘addition’ of adjacent congruent segments. Unique angle trans-
fer, comparison and addition follow similarly from axiom C-4 and Definition 2.18 (interior points).
Neither Hilbert nor Euclid use or require an absolute notion of length/angle-measure: the compari-
son AB < CD does not indicate a relationship between numerical quantities (lengths). Introducing
numerical length requires the inclusion of the real numbers (and thus far more axioms)—for purity
reasons, we postpone this until Section 2.5.
21
The Triangle Congruence Theorems: SAS, ASA, SSS & SAA
Hilbert assumes side-angle-side (SAS) and proceeds to prove the remainder. Here is the first of these;
we’ll cover SSS momentarily and SAA in Exercise 6.
Theorem 2.23 (Angle-Side-Angle/ASA, Euclid I. 26, case I). Suppose ABC and DEF satisfy
ABC
=
DEF, AB
=
DE, BAC
=
EDF
Then the triangles are congruent (ACB
=
DFE, AC
=
DF and BC
=
EF).
Hilbert’s approach modifies Euclid’s: instead of laying ABC on top of DEF, he creates a new
triangle DEG
=
ABC and proves that G = F.
Proof. Segment transfer provides the unique point G
EF such that EG
=
BC.
SAS applied to AB
=
DE, ABC
=
DEG (= DEF) , BC
=
EG,
says BAC
=
EDG (
=
EDF) (this last is by assumption).
Since F and G lie on the same side of
DE, angle transfer (C-4) says
they lie on the same ray through D.
But then F and G both lie on two distinct lines (
EF =
EG and
DF =
DG). We conclude that F = G.
By SAS we conclude that ABC
=
DEF.
A
B
C
D E
F
G
Geometry Without Circles
Circles are at the heart of Euclid’s constructions. Yet, for reason we’ll address in Section 2.4, Hilbert
essentially ignores them. We sketch a few of his alternative approaches to Euclid’s basic results.
Theorem 2.24 (Euclid I. 5). An isosceles triangle has congruent base angles.
Isosceles means equal legs: two sides of the triangle are congruent. The remaining side is the base.
Euclid’s argument relies on a famously complicated construction (look it up!). Hilbert does things
more speedily and sneakily, by relabelling the original triangle and applying SAS.
Proof. Suppose ABC is isosceles where AB
=
AC. Consider a ‘new’
triangle A
B
C
= ACB where the base points are switched:
A
:= A, B
:= C, C
:= B
Observe:
BAC
=
CAB (axiom C-5) = BAC
=
B
A
C
.
AB
=
AC = AB
=
A
B
and AC
=
A
C
.
SAS says that ABC
=
A
B
C
=
ACB.
A = A
B = C
C = B
22
Dropping a Perpendicular As with the majority of Book I, Euclid accomplishes this using circle
intersections.
10
Hilbert instead uses segment/angle transference and the concept of sidedness.
Suppose we are given a line and a point P not on . Our goal is to
construct a point M such that PM intersects in a right-angle.
Let A, B be distinct points on (axiom I-3) so that =
AB.
By axioms C-4 and C-1, we may transfer AP to the other side of at
A, creating a new point Q.
Since P and Q lie on opposite sides of , the line intersects PQ at some
point M. There are two cases to consider.
In the generic case M = A (pictured), SAS applied to MAP
and MAQ shows that AMP
=
AMQ. Since these angles
sum to a straight edge (PQ), they are both right-angles.
In the extreme case M = A, there are no triangles and SAS
cannot be applied. Instead, observe that B does not lie on PQ
(which axioms/results make this clear?!) and apply the above
argument with B instead of A.
A
B
M
P
Q
A generalization of this construction facilitates a corrected argument for the SSS triangle congruence.
Theorem 2.25 (Side-Side-Side/SSS, Euclid I. 8). Suppose ABC and DEF have sides congruent
in pairs:
AB
=
DE, BC
=
EF, AC
=
DF
Then the triangles are congruent (ABC
=
DEF, BCA
=
EFD, CAB
=
FDE).
The strategy is similar to the proof of ASA. Hilbert creates a new triangle DEG
=
ABC, though
this time with G on the opposite side of
DE to F.
Proof. Transfer BAC to D on the other side of
DE from F to
obtain G (axioms C-4 and C-1).
SAS (AB
=
DE, BAC
=
EDG, AC
=
DG) shows that
EG
=
BC
=
EF. Otherwise said, DEG
=
ABC.
Join FG to produce isosceles triangles FDG and FEG
with base FG, both with congruent angles at F and G.
Sum angles at F and G and apply SAS (DF
=
DG, DFE
=
DGE, EF
=
EG) to see that DEF
=
DEG.
We conclude that ABC
=
DEG
=
DEF, as required.
To be completely formal, we should also carefully deal with
the situations where the sum is a subtraction or the triangle
is right-angled at A or B.
D E
F
G
A
B
C
10
Consider the picture for Thm. I. 11 in Exercise 2.2.2.
23
Exterior Angle Theorem (Thm. 2.3, I. 16) Euclid’s approach uses a bisector which he obtains from
circles. Hilbert does things a little differently.
Proof. Given ABC, extend AB to D such that AC
=
BD. For clarity, we label angles with Greek let-
ters as in the first picture below. We show that γ < δ by proving that the alternatives are impossible.
1. (δ γ) Assume δ
=
γ. By SAS, ACB
=
DBC; in particular ϵ
=
β. Since A and D lie on
opposite sides of
BC, we see that ϵ + γ
=
β + δ is a straight edge. But then A, D are distinct
points lying on two lines! Contradiction.
2. (δ < γ) Assume δ < γ. Transfer δ to C as shown to obtain η
=
δ. By the crossbar theorem, we
obtain an intersection point E. But now δ is an exterior angle of EBC congruent to an interior
angle η of the same triangle, contradicting part 1.
A
B
C
D
α
β
γ
δ
ϵ
A
B
C
D
E
α
β
η
δ
ϵ
Step 1: δ
=
γ is a contradiction Step 2: δ < γ is a contradiction
Take the vertical angle to δ at B and repeat the argument to see that α < δ.
The proof also shows that the sum of any two angles in a triangle is strictly less than a straight edge:
α + β < δ + β = 180°.
Is Euclid now fixed? Almost! In the exercises we show how the following may be achieved:
Construction of an isosceles triangle on a segment AB. With this one can construct segment
and angle bisectors (Euclid I. 9+10).
SAA congruence (Euclid I. 26, case II), the last remaining triangle congruence theorem.
We’ve now recovered almost all of Book I prior to the application of the parallel postulate. Including
Playfair’s axiom completes the remainder, including Pythagoras’, all without circles!
Exercises 2.3. Except for question 8, answer everything without reference to the continuity axiom,
circles, or the uniqueness of parallels (e.g., Playfair’s axiom, (tri)angle sum = 180°).
1. Draw pictures to suggest why you don’t expect Angle-Angle-Angle (AAA) and Side-Side-
Angle (SSA) to be triangle congruence theorems.
2. Use Hilbert’s axioms C-4 and C-5 to prove that congruence of angles is an equivalence relation.
3. (a) Use ASA to prove that if the base angles are congruent then a triangle is isosceles.
(b) Find an alternative argument that relies the exterior angle theorem.
(Hint: this is essentially the same as the proof of Exercise 5 (a))
(c) Explain why the base angles of an isosceles triangle are acute (less than a right-angle).
24
4. Given AB, axiom I-3 says C
AB.
If ABC is not isosceles, then WLOG assume ABC < BAC.
Transfer ABC to A to produce D on the same side of
AB as C with
ABC
=
BAD, BC
=
AD
A
B
C
D
M
(a) Explain why rays
AD and
BC intersect (at some point M).
(b) Why is MAB isosceles?
(c) Describe how to produce the perpendicular bisector of AB.
(d) Explain how to construct an angle bisector using the above discussion.
5. We prove Theorems I. 18, 19 and 20 on comparisons
of angles and sides in a triangle. For clarity, suppose
ABC has sides and angles labelled as in the picture.
c
b
a
A
B
C
α
β
γ
(a) (I. 18) Assume a < c. Prove that α < γ.
(Hint: let D on AB satisfy BD
=
a)
(b) (I. 19: converse to I. 18) α < γ = a < c.
(Hint: Prove the contrapositive)
(c) (I. 20: triangle inequality) a + b > c.
(Hint: Let E lie on
BC such that CE
=
b and apply I. 19)
6. Prove the SAA congruence. If ABC and DEF satisfy
AB
=
DE, ABC
=
DEF and BCA
=
EFD
then the triangles are congruent: ABC
=
DEF.
(Hint: Let G
BC be such that BG
=
EF, apply SAS and the exterior
angle theorem)
G
B
A
C
7. Hilbert’s published SAS axiom is weaker than we’ve stated.
Given triangles ABC and A
B
C
, if AB
=
A
B
, AC
=
A
C
, and BAC
=
B
A
C
, then ABC
=
A
B
C
.
Use this to prove the full SAS congruence theorem (axiom C-6 as we’ve stated it).
(Hint: try a trick similar to the proof of ASA)
8. Construct the picture on the right, where BF is the perpendicular
bisector of AC, and
BD
=
AB, BE
=
AD, BF
=
AE
Use Pythagoras’ Theorem to prove that ACF is equilateral.
This construction requires Playfairs axiom and thus unique parallels. It
does not require circle intersections (continuity) like Theorem 2.1 (I. 1).
A
B
C
D
E
F
25
2.4 Circles and Continuity
Definition 2.26. Let O and R be distinct points. The circle C with center O and
radius OR is the collection of points A such that OA
=
OR.
A point P lies inside the circle C if P = O or OP < OR.
A point Q lies outside if OR < OQ.
Since all segments are comparable, any point lies inside, outside or on a given
circle.
O
R
A
P
Q
A major weakness of Euclid is that many of his proofs rely on circle intersections, rather than lines.
To use circles in this manner requires the Axiom of Continuity. This much more technical than the
other axioms. It is likely for this reason that Hilbert barely mentions circles, instead wanting to build
as much geometry as possible using only the simplest axioms.
Here are the facts necessary before Euclid’s approach can be followed.
Theorem 2.27. Suppose C and D are circles.
1. (Elementary/line-circle Continuity Principle) If P lies inside and Q outside C, then PQ inter-
sects C in exactly one point.
2. (Circular Continuity Principle) If D contains a point inside and another outside C, then they
intersect in exactly two points. These lie on opposite sides of the line joining the circle centers.
The idea of the first principle is to partition PQ into two pieces:
Σ
1
consists of the points lying on or inside C
Σ
2
consists of the points lying outside C
One shows that Σ
1
and Σ
2
satisfy the assumptions of the axiom. The
unique point O then exists and is shown to lie on C itself. Some of this are
in Exercise 6. The circular continuity principle is harder.
P
Q
O
Σ
1
Σ
2
C
What is perhaps more interesting is to consider a geometry in which the axiom of continuity is false.
Example 2.28. The geometry Q
2
= {(x, y) R
2
: x, y Q} of points in the plane with rational
co-ordinates satisfies almost all of Hilbert’s axioms, however C-1 and continuity are false.
Axiom C-1 Given points A = (0, 0), B = (1, 0) and C = (1, 1), we see
that O = (
1
2
,
1
2
) is the unique point (in R
2
) on the ray r =
AC
such that AC
=
AB. Clearly O is an irrational point and therefore
not in the geometry.
Continuity The circle centered at A = (0, 0) with radius 1 does not
intersect the segment AC. More properly, AC = Σ
1
Σ
2
may be
partitioned as shown and yet no point O in the geometry separates
Σ
1
, Σ
2
.
Σ
1
Σ
2
A
B
C
O
26
Equilateral triangles We can finally correct Euclid’s proof of the first proposition of the Elements!
Theorem 2.29 (Euclid I.1). An equilateral triangle many be constructed on a given segment AB.
Proof. Following Euclid, take the circles α and β centered at A and B, with radii AB.
Axiom O-2: D such that A B D.
Axiom C-1: let C
BD be such that BC
=
AB.
Circular continuity principle: β contains A (inside α) and
C (outside α) so the circles intersect in two points P, Q.
Since P lies on both circles (and is therefore distinct from
A and B), we have AB
=
AP
=
BP whence ABC is
equilateral.
A
B
C
D
P
Q
α
β
If one allows Playfair’s axiom on unique parallels, Euclid’s result can be proved without using circles
or the continuity axiom (see Exercise 2.3.8). Nevertheless, we are finally able to say that every result
in Book I of Euclid is correct, even if his original axioms and arguments are insufficient!
Basic Circle Geometry
We continue our survey of Euclidean geometry with a few results about circles, many of which are
found in Book III of the Elements. From this point on, we assume all Hilbert’s axioms including
Playfair and continuity. Indeed what follows often relies on their consequences, particularly angle-
sums in triangles and the circular continuity principle.
Definition 2.30. With reference to the picture:
A chord AB is a segment joining two points on a circle.
A diameter BC is a chord passing through the center O.
An arc
)
AB is part of the circular edge between chord points
(major or minor by length).
AOB is a central angle and APB an inscribed angle.
ABP is inscribed in its circumcircle.
A
B
C
P
O
Since these ideas shouldn’t be new, most of the details are left as exercises.
Theorem 2.31 (III. 20). The central angle is twice the inscribed angle: AOB = 2APB.
For a sketch proof, join OP, breaking ABP into three isosceles triangles and count angle sums.
Corollary 2.32. 1. (III. 21) If inscribed triangles share a side, the opposite angles are congruent.
2. (III. 22) An inscribed quadrilateral has opposite angles supplementary (summing to 180°).
3. (Thales’ Theorem III. 31) A triangle in a semi-circle is right-angled.
27
Theorem 2.33. Any triangle has a unique circumcircle.
This is similar to III. 1: construct the perpendicular bisectors of
two sides as in the picture.
A
B
C
O
Definition 2.34. A line is tangent to a circle if it intersects the circle exactly once.
Theorem 2.35 (III. 18, 19 (part)). A line is tangent to a circle if and only if it is perpendicular to the
radius at an intersection point.
Proof. () Suppose through T is perpendicular to the radius OT.
Let P be any another point on . But then (Exercise 2.3.5),
OPT < 90° = OTP = OT < OP
thus P lies outside the circle. Every point on except T lies outside
the circle, so T is the unique intersection, and is therefore tangent.
The converse is an exercise.
T
O
P
Theorem 2.36. Through a point outside a circle, exactly two lines are tangent to the circle.
Exercises 2.4. 1. Give formal proofs of all parts of Corollary 2.32.
2. Prove Theorem 2.33.
3. Complete the proof of Theorem 2.35 by showing the () direction.
(Hint: suppose T is an intersection and that the angle isn’t 90°, and drop a perpendicular to . . . )
4. Given a circle centered at O and a point P outside the circle, draw the circle centered at the
midpoint of OP passing through O and P. Explain why the intersections of these circles are the
points of tangency required in Theorem 2.36, and hence complete its proof.
5. (a) Prove Theorem 2.31 when O is an interior point to ABP.
(b) Prove Theorem 2.31 when O is an exterior point to ABP.
6. Suppose A C B and that O /
AB. Use Exercise 2.3.5 to show that
OC < max
OA, OB
If A, B are interior to a circle centered at O, conclude that C is also.
(This is part of what’s needed to demonstrate the elementary continuity principle:
no point of Σ
2
lies between two points of Σ
1
. Can you prove the other condition?)
A
B
C
O
7. If a line contains a point inside a circle, show that it intersects the circle in two points.
(Hint: first construct two points on the line lying outside the circle)
28
2.5 Similar Triangles, Length and Trigonometry
In the geometry of Euclid & Hilbert, there are no numerical measures of length or angle. Relative
measure is built in (Definition 2.22), and we’ve denoted right-angles and straight edges by 90° & 180°
purely for convenience. To avoid continued frustration it is time we introduced explicit numerical
measure, though to do so properly requires more axioms!
Axioms 2.37 (Length and Angle Measure).
L1 To each segment AB corresponds a unique length
|
AB
|
, a positive real number
L2
|
AB
|
=
|
CD
|
AB
=
CD
L3
|
AB
|
<
|
CD
|
AB < CD (Definition 2.22)
L4 If A B C, then
|
AB
|
+
|
BC
|
=
|
AC
|
A1 To each ABC corresponds a unique degree measure mABC, a real number between 0 and 180
A2 mABC = mDEF ABC
=
DEF
A3 mABC < mDEF ABC < DEF (Definition 2.18)
A4 If P is interior to ABC, then mABP + mPBC = mABC
A5 Right-angles measure 90°
Don’t memorize these axioms, just observe how they fit your intuition. Angle measure in Euclidean
geometry has two notable differences from what you might expect:
(A1) All angles measure strictly between and 180°. A straight edge isn’t an angle, though
such is commonly denoted 180°, and there are no reflex angles (> 180°).
(A2) Angles are non-oriented, measuring the same in reverse (mABC = mCBA).
The axioms for length and angle follow the same pattern except for us explicitly fixing the scale of
angle measure (A5). To do the same for length requires only a choice of a reference segment of length
1. The following is a consequence of the continuity axiom.
Theorem 2.38 (Uniqueness of measure).
1. Given OP, there is a unique way to assign a length to every segment such that
|
OP
|
= 1.
2. There is a unique way to assign a degree measure to every angle.
The segment OP in part 1 provides a length-scale
for a ruler. We measure the length of any segment
by moving this rule on top of the desired segment
(congruence!).
0
1
2
3
4
5
1
O
P
0
1
2
3
4
5
|QR| = 4.6
Q
R
29
Area Measure If we also include Playfair’s axiom, then the discussion at the end of Book I of Euclid
becomes valid, and rectangles can be defined (see e.g., Exercise 2.1.2).
Definition 2.39. The area (measure) of a rectangle is the product of its base and height (measures).
Given a length measure, a square with side length 1 necessarily has area 1. Relative to a base segment,
the height of a triangle is the length of the perpendicular dropped from the vertex.
Since every rectangle is a parallelogram and a triangle half a parallelogram, Eu-
clid’s discussion (Thm. I. 35) amounts to the familiar area formulæ:
area(parallelogram) = bh, area(triangle) =
1
2
bh
While these expressions are nice to have, they are not necessary. Indeed every-
thing that follows depends only on area ratios: e.g.,
1
2
bh
1
1
2
bh
2
=
h
1
h
2
.
h
b
Lemma 2.40. If triangles have congruent bases, then their areas are in the same ratio as their heights.
The same holds with the roles of heights and bases reversed.
Similarity and the AAA Theorem Similar triangles are the concern of Book VI of the Elements.
Definition 2.41. Triangles are similar, written ABC XYZ, if
their sides are in the same length ratio
|
AB
|
|
XY
|
=
|
BC
|
|
YZ
|
=
|
CA
|
|
ZX
|
A
B
C
X
Y
Z
Euclid discusses these using non-numerical ratios of segments (e.g., AB : XY = BC : YZ). This is
unnecessarily confusing for modern readers, indeed some of the most difficult parts of the Elements
are where he describes what this should mean, particularly for irrational ratios (Books V & X).
Our primary result comes in two versions, where the second (which we’ll prove) is a special case of
the first.
Theorem 2.42 (Angle-Angle-Angle/AAA, Euclid VI. 2–5).
1. Triangles are similar if and only if their angles come in mutually con-
gruent pairs.
2. Suppose a line intersects two sides of a triangle. The smaller triangle
so created is similar to the original if and only if the line is parallel to
the third side of the triangle.
A
B
C
The picture should convince you that 1 2 follows from the uniqueness of parallels (Playfair’s
axiom, Corollary 2.5 & Theorem 2.6). This reliance is crucial! We should not expect AAA similarity
in non-Euclidean geometry, and indeed shall see later that it is false in hyperbolic geometry (Chapter
4), where AAA is a theorem for congruent triangles! The converse (2 1) is left to Exercise 10.
30
Proof (AAA similarity, part 2). Suppose intersects ABC at
points D, E as shown. Drop perpendiculars to create distances
h, k, d
1
, d
2
as indicated. We prove:
is parallel to BC ABC ADE
( ) Suppose is parallel to BC. Playfair
11
tells us that d
1
= d
2
.
By Lemma 2.40, triangles with the same height have areas
proportional to their bases:
d
1
d
2
h
k
A
B
C
D
E
|
BD
|
|
AD
|
=
area(BDE)
area(ADE)
(BDE, ADE have same height h)
|
CE
|
|
AE
|
=
area(CDE)
area(ADE)
(CDE, ADE have same height k)
Since BDE and CDE share base DE, we see that
is parallel to BC d
1
= d
2
area(BDE) = area(CDE)
|
BD
|
|
AD
|
=
|
CE
|
|
AE
|
()
Add 1 =
|
AD
|
|
AD
|
=
|
AE
|
|
AE
|
to both sides to obtain one part of the similarity ratio
|
AB
|
|
AD
|
=
|
AD
|
+
|
BD
|
|
AD
|
=
|
AE
|
+
|
CE
|
|
AE
|
=
|
AC
|
|
AE
|
It remains to see that this ratio equals
|
BC
|
|
DE
|
. Again using common heights (h, k) of triangles,
|
AB
|
|
BD
|
=
area(ABE)
area(BDE)
|
CE
|
|
AE
|
=
area(BCE)
area(ABE)
(†)
=
|
AB
|
|
AD
|
()
=
|
AB
|
|
BD
|
|
CE
|
|
AE
|
(†)
=
area(BCE)
area(BDE)
=
|
BC
|
|
DE
|
where the last equality follows since BCE and BDE have common height d
1
= d
2
.
( ) Suppose ABC ADE. By Playfair, let m be the
unique parallel to BC through D. This intersects AC at a
point G. We must prove that G = E (consequently m = ).
By the () direction above,
ABC ADG
d
1
d
2
m
A
B
C
D
E
G
However, is plainly transitive (it is an equivalence relation), whence ADE ADG. The
similarity ratio is 1 =
|
AD
|
|
AD
|
, whence
|
AE
|
|
AG
|
= 1 =
|
AE
|
=
|
AG
|
= E = G
11
d
1
= d
2
parallel to BC is Playfair! Compare Exercise 2.1.2 (Thm I. 46) on the construction of a square. . .
31
Applications of Similarity: Trigonometric Functions, Cevians and the Butterfly Theorem
We finish with several applications of similarity which hopefully give an idea of what can be done
without co-ordinates. None of these ideas were known to Euclid.
Definition 2.43. Given an acute angle ABC (mABC < 90°), drop a perpendicular from A to
BC
at D so that ADB is a right-angle. Define
sin ABC :=
|
AD
|
|
AB
|
cos ABC :=
|
BD
|
|
AB
|
Early trigonometry dates to a few hundred years after Euclid, though the approach was different.
12
Theorem 2.44. Angles have the same sine (cosine) if and only if they are congruent.
Proof. Assume ABC
=
A
B
C
as in the picture, and drop perpendiculars to D, D
.
Since ABD and A
B
D
have two pairs of mutually
congruent angles, the third pair is congruent also. AAA
applies, the triangles are similar and
|
AD
|
|
AB
|
=
|
A
D
|
|
A
B
|
|
BD
|
|
AB
|
=
|
B
D
|
|
A
B
|
A
B
C
D
A
B
C
D
In particular, sin ABC = sin A
B
C
and cos ABC = cos A
B
C
.
The converse is an exercise.
After Giovanni Ceva (1647–1734), a cevian is a segment joining a vertex to the opposite side of a
triangle. Here is a beautiful result from the height of Euclidean geometry—good luck trying to prove
it using co-ordinates!
Theorem 2.45 (Ceva’s Theorem). Given ABC and cevians AX, BY, CZ,
|
BX
|
|
XC
|
|
CY
|
|
YA
|
|
AZ
|
|
ZB
|
= 1 the cevians meet at a common point P
Proof. () This is simply a repeated application of Lemma 2.40.
area(ABX)
area(AXC)
=
|
BX
|
|
XC
|
=
area(PBX)
area(PXC)
=
|
BX
|
|
XC
|
()
=
area(ABX) area(PBX)
area(AXC) area(PXC)
=
area(ABP)
area(APC)
Repeat for the other ratios and multiply to get 1.
A simple justification of () and the converse are an exercise.
A
B
C
X
Y
Z
P
12
The ancient forerunners of sine and cosine were defined using chords of circles rather than triangles. The word
trigonometry (literally triangle measure) wasn’t coined until 1595.
32
Theorem 2.46 (Butterfly Theorem). We are given the following data
as in the picture:
PQ is a chord of a circle with midpoint M.
AC and BD are chords meeting at M.
X, Y are the chord-intersections as shown.
Then M is the midpoint of XY.
A
B
C
D
M
P
Q
X
Y
This beautiful result dates to 1803-5 and has several proofs. We present an argument relying on
similar triangles.
Proof. For convenience we introduce several numerical
lengths:
z =
|
PM
|
=
|
MQ
|
, x =
|
XM
|
and y =
|
YM
|
.
Drop perpendiculars from X, Y to chords AD, BC,
and label the lengths x
1
, x
2
, y
1
, y
2
as shown.
The four colored pairs of angles are congruent: vertical an-
gles at M, and inscribed angles at A, B, C, D.
We compare sides of several similar triangles:
x
x
1
=
y
y
1
and
x
x
2
=
y
y
2
=
x
y
=
x
1
y
1
=
x
2
y
2
|
AX
|
|
BY
|
=
x
1
y
2
and
|
XD
|
|
YC
|
=
x
2
y
1
x
y
x
1
x
2
y
1
y
2
A
B
C
D
M
P
Q
X
Y
The result follows by putting these observations together and applying Exercise 1 twice:
13
x
2
y
2
=
x
1
x
2
y
1
y
2
=
|
AX
||
XD
|
|
BY
||
YC
|
=
|
PX
||
XQ
|
|
PY
||
YQ
|
=
( z x)(z + x)
( z + y)(z y)
=
z
2
x
2
z
2
y
2
= x
2
( z
2
y
2
) = y
2
( z
2
x
2
)
= x
2
= y
2
= x = y
as required.
Exercises 2.5. 1. Let AD and PQ be chords of a circle which intersect at X. Use similar triangles to
prove that
|
AX
||
XD
|
=
|
PX
||
XQ
|
13
The denominators are equal by applying the Exercise to the chords BC and PQ.
33
2. Let ABC have a right-angle at C. Drop a perpendicular from C to
AB at D.
(a) Prove that D lies between A and B.
(b) Prove that you have three similar triangles
ACB ADC CDB
(c) Use these facts to prove Pythagoras’ Theorem.
(Use the picture, where a, b, c, x, y are lengths)
a
b
c = x + y
x
y
A
B
C
D
3. Prove a simplified version of the SAS similarity theorem:
|
AB
|
|
AG
|
=
|
AC
|
|
AH
|
ABC AGH
(Hint: construct BJ parallel to GH and appeal to AAA)
A
B
C
G
H
4. By excluding the other possibilities, prove the converse of length axiom L4:
If A, B, C are distinct and
|
AB
|
+
|
BC
|
=
|
AC
|
, then B lies between A and C.
5. Use Pythagoras’ to prove that sin 45° = cos 45° =
1
2
, that sin 60° =
3
2
and cos 60° =
1
2
.
6. Prove the converse of Theorem 2.44: if sin ABC = sin A
B
C
, then ABC
=
A
B
C
.
(Hint: create right-triangles and prove they are similar. Label the side lengths o, a, h, etc.)
7. We complete the proof of Ceva’s theorem.
(a) If p, q, r, s are non-zero real numbers, verify that α =
p
q
=
r
s
= α =
pr
qs
.
(b) Assume X, Y, Z satisfy Ceva’s formula.
Define P as the intersection of BY and CZ and let
AP
meet BC at X
.
Prove the () direction of Ceva’s theorem by using
the () direction to show that X
= X.
A
B
C
X
X
Y
Z
P
8. (a) A median of a triangle is a segment from a vertex to the midpoint of the opposite side. Use
Ceva’s Theorem to prove that the medians of a triangle meet at a point (the centroid).
(b) (Hard) Medians split a triangle into six sub-triangles. Prove that all have the same area.
(c) Prove that the centroid is exactly 2/3 of the distance along each median.
9. Prove that similarity of triangles is an equivalence relation.
(Don’t use AAA since its proof requires this fact!)
10. (Hard) Explain how to prove (2 1) in the AAA Theorem (2.42).
34
3 Analytic Geometry
Geometry in the style of Euclid and Hilbert is synthetic: axiomatic, without co-ordinates or explicit
formulæ for length, area, volume, etc. By contrast, the practice of elementary geometry nowadays
is typically analytic: reliant on co-ordinates & algebra, vectors. The critical invention was the axis,
developed by Ren
´
e Descartes and Pierre de Fermat in the early 1600s: a fixed reference ruler against
which objects can be measured using co-ordinates.
3.1 The Cartesian Co-ordinate System
Since Cartesian geometry (Descartes’ geometry) should be familiar, we merely sketch the core ideas.
Perpendicular axes meet at the origin O.
The co-ordinates of a point are measured by projecting onto the axes;
since these are real numbers we denote the set of these
R
2
=
(x, y) : x, y R
E.g., P has co-ordinates (1, 2), we usually just write P = (1, 2).
Algebra is introduced via addition and scalar multiplication
2
1
1
2
3
y
2 1 1 2 3
x
P
P + Q = (p
1
, p
2
) + (q
1
, q
2
) = (p
1
+ q
1
, p
2
+ q
2
) λP = (λp
1
, λp
2
)
The length of a segment uses Pythagoras’ Theorem
d(P, Q) =
|
PQ
|
=
q
( q
1
p
1
)
2
+ (q
2
p
2
)
2
In the picture
|
OP
|
=
1
2
+ 2
2
=
5. As in Section 2.5, segments are congruent if and only if
they have the same length.
Curves are defined using equations. E.g. x
2
+ y
2
= 1 describes a circle.
Analytic geometry was conceived as a computational toolkit built on top of Euclid. At first, math-
ematicians felt the need to justify analytic arguments synthetically lest no-one believe their work.
14
Synthetic geometry is not without its benefits, but its study has increasingly become a fringe activity;
co-ordinates are just too useful to ignore.
We may therefore assume anything from Euclid and mix strategies as appropriate. To see this at
work, consider a simple result.
Lemma 3.1. Non-collinear points O = (0, 0), A = (x, y), B = (v, w) and
C := (x + v, y + w), form a parallelogram OACB.
Proof. Opposite sides have the same length (
|
BC
|
=
p
x
2
+ y
2
=
|
OA
|
, etc.)
and are thus congruent. SAS shows OAC
=
CBO. Euclid’s discussion
of alternate angles (pages 10–11) forces opposite sides to be parallel.
O
A
B
C
14
This attitude persisted for some time. For instance, when Issac Newton published his groundbreaking Principia in
1687, his presentation was largely synthetic, even though he had used co-ordinates in his derivations.
35
Lemma 3.2. The points X
t
on the line
PQ are in 1–1 correspondence with the real numbers via
X
t
= P + t(Q P) = ( 1 t)P + tQ
Moreover, d(P, X
t
) =
|
t
||
PQ
|
so that t measures the (signed) distance along the line.
The proof is an exercise. As an example of how easy it can be to work in analytic geometry, we
repeatedly apply the Lemma to re-establish a famous result.
Theorem 3.3. The medians of a triangle meet at a point 2/3 of the way along each median.
Proof. Given ABC, label the midpoints of each side as shown. By the Lemma, these are
M =
1
2
(B + C), N =
1
2
(A + B), P =
1
2
(A + C)
The point
2
3
of the way along median AM is then
A +
2
3
(M A) = A +
2
3
(B + C 2A) =
1
3
(A + B + C)
By symmetry (check directly if you like!), this is also the point
2
3
of the
way along the other two medians.
A
C
B
M
P
N
G
The three points are therefore identical: the medians meet at the centroid G =
1
3
(A + B + C).
Compare this to Exercise 2.5.8 where we used Ceva’s Theorem!
Exercises 3.1. 1. By completing the square, identify the curve described by the equation
x
2
+ y
2
4x + 2y = 10
2. (a) Perform a pure co-ordinate proof of Theorem 3.3. For simplicity, arrange the triangle so
that A = (0, 0) is the origin, and B points along the positive x-axis.
(b) Descartes and Fermat did not have a fixed perpendicular second axis! Their approach was
equivalent to choosing a second axis at an angle which made the problem as simple as
possible.
Given ABC, let A be the origin and choose axes which point along the edges AB and
BC. What are the co-ordinates of B and C with respect to these axes? Now give an even
simpler proof of the centroid theorem.
3. Prove Lemma 3.2.
4. A parabola is a curve whose points are equidistant from a fixed
point F, the focus, and a fixed line d (the directrix). Choose axes as
shown in the picture so that F = (0, a) and d has equation y = a.
Find the equation of the parabola.
y
x
a
d
F
P
36
3.2 Angles and Trigonometry
Angles are defined differently to Section 2.5, though the approach should feel familiar.
Definition 3.4. Suppose A, B, C are distinct points in the plane. Take
any circular arc centered at A and define the radian measure
BAC :=
arc-length
radius
[0, 2π)
where arc-length is measured counter-clockwise from
AB to
AC.
θ
2π θ
A
B
C
Since arc-length scales with radius, the definition is independent of the radius of the circular arc. It
is important to appreciate the difference between angle measures in our two geometries.
Euclidean geometry All angles < 180°. Reversed legs congruent angles and same degree measure:
CAB
=
BAC mCAB = mBAC
Analytic geometry Reflex angles exist ( π). Reversed legs different radian measure:
CAB = 2π θ = 2π BAC = BAC (unless a straight edge)
In the picture, CAB is not the radian measure (θ) of CAB! However,
Angles congruent radian measures equal and < π
As such, it is common to label angles in a triangle by their radian measure;
standard convention is shown: e.g., (A, a, α) for (point,length,angle).
α
β
γ
a
b
c
A
B
C
Definition 3.5 (Trigonometric Functions). Let O be the origin and I = (1, 0).
Let P = (x, y) lie on a circle of radius r and θ = IOP. We define:
cos θ :=
x
r
sin θ :=
y
r
tan θ :=
y
x
(x = 0)
AAA similarity (Thm. 2.42) says these are well-defined, independent of r.
x
y
r
θ
P
O
I
Example 3.6. Basic trig identities should be obvious from the picture: e.g.,
cos
2
θ + sin
2
θ = 1 (Pythagoras!) and sin θ = cos(
π
2
θ)
What well-known facts regarding sine and cosine do the following illustrate?
θ
π θ
θ
cos θ
sin θ
1
1
1
2
1
3
2
37
Solving Triangles A triangle is described by six values: three side lengths and three angle mea-
sures. Euclid’s triangle congruence theorems (SAS, ASA, SSS, SAA) say that three of these in suitable
combination is enough to recover the rest. In analytic geometry, these calculations typically use the
sine and cosine rules.
Theorem 3.7. Label the sides/angles of ABC following the standard convention (page 37):
Sine Rule If d is the diameter of the circumcircle (Defn. 2.30), then
sin α
a
=
sin β
b
=
sin γ
c
=
1
d
Cosine Rule c
2
= a
2
+ b
2
2ab cos γ
Proof. We prove the sine rule and leave the cosine rule as an exercise.
Everything relies on Corollary 2.32. Draw the circumcircle of ABC.
Construct BCD with diameter BD; this is right-angled at C by Thales’
Theorem. There are two cases:
1. If A lies on the same side of
BC as D, then A and D share the same
arc, whence BDC = α and
a = d sin BDC = d sin α
2. If A lies on the opposite side, then the quadrilateral ABDC lies on
a circle. Opposite angles at A, D are supplementary, whence
sin α = sin(π α) = sin BDC =
a
d
The two other angle-side combinations follow by permutation.
a
d
α
α
A
B
C
D
a
d
α
π α
A
B
C
D
Examples 3.8. 1. The SSS congruence corresponds to solving a triangle using the cosine rule. For
instance, the given triangle has angles
α =
6
2
+ 7
2
3
2
2 ·6 ·7
= cos
1
19
21
25° β =
3
2
+ 7
2
6
2
2 ·3 ·7
= cos
1
11
21
58°
γ =
3
2
+ 6
2
7
2
2 ·3 ·6
= cos
1
1
9
96°
Once you have α, you could alternatively switch to the sine rule to find β, before
computing γ = π α β.
7
3
6
α
β
γ
2. To solve a triangle with data corresponding to the ASA congruence,
find the remaining angle γ = π
π
4
π
3
=
5π
12
and apply the sine rule
sin
π
4
a
=
sin
π
3
b
= sin
5π
12
= cos
π
12
= a =
1
2 cos
π
12
0.732
b =
3
2 cos
π
12
0.897
1
a
b
π
4
π
3
γ
38
Multiple-angle formulæ The picture provides a very simple
proof of the expressions
sin(α + β) = sin α cos β + cos α sin β
cos(α + β) = cos α cos β sin α sin β
at least when α + β <
π
2
. A little algebraic manipulation pro-
duces the double-angle and difference formulæ, and verifies
that these hold for all possible angle inputs.
α
β
1
α
α
cos β
sin β
sin α cos β
cos α sin β
sin 2α = 2 sin α cos α sin(α β) = sin α cos β cos α sin β
cos 2α = cos
2
α sin
2
α = 2 cos
2
α 1 cos(α β) = cos α cos β + sin α sin β
Exercises 3.2. 1. A triangle has angle of
2π
3
radians between sides of lengths 2 and
3 1. Find the
length of the remaining side, and the remaining angles.
2. Describe how to solve a triangle given data in line with the SAA congruence theorem.
3. Two measurements for the height of a mountain are taken at sea level 5000 ft apart in a line
pointing away from the mountain. The angles of elevation to the mountain top from the hori-
zontal are 15° and 13° respectively. What is the height of the mountain?
4. Use a multiple angle formula to find an exact value for cos
π
12
and thus exact values for the side
lengths of the triangle in Example 3.8.2.
5. The area of a triangle is
1
2
(base)·(height). By using each side of the triangle alternately as the
‘base,’ find an alternative proof of the sine rule without the relationship to the circumcircle.
6. You are given SSA data for a triangle: sides with lengths a = 1 and b =
3 and angle α =
π
6
.
Show that there are two triangles satisfying this data. Can you generalize to general SSA data?
7. (a) By dropping a perpendicular from B to
AC at D, construct a
proof of the cosine rule.
(Hint: apply Pythagoras’ to the two right-triangles)
(b) Is your argument valid if D is not interior to AC?
a
c
b
h
γ
B
AC
D
8. The dot product of A = (a
1
, a
2
) and B = (b
1
, b
2
) is A · B := a
1
b
1
+ a
2
b
2
. Apply the cosine rule to
OAB to prove that
A · B =
|
OA
||
OB
|
cos AOB
9. Derive the multiple-angle formula for sin(α β).
(Remember that 0 α, β, α β < 2π so you can’t simply switch the sign of β!)
10. Given the arrangement pictured, find x, the radian-measure α and
the exact value of cos α.
(Hint: first show that you have similar isosceles triangles)
1
x
x x
1 x
α
39
3.3 Isometries
At the heart of elementary geometry is congruence, the idea that geometric figures can be essentially
the same without necessarily being equal. In analytic geometry, congruence may be described alge-
braically using functions. This follows from the idea that two segments have the same length if and
only if they are congruent.
Definition 3.9. A function f : R
2
R
2
is a (Euclidean) isometry if it preserves lengths:
15
P, Q R
2
, d
f (P), f (Q)
=
|
PQ
|
Two figures (segments, angles, triangles, etc.) are congruent precisely when there is an isometry
f : R
2
R
2
mapping one to the other.
Example 3.10. We check that the map f (x, y) =
1
5
3x + 4y, 4x 3y
+ (3, 1) is an isometry. If
P = (x, y) and Q = (v, w), then
d
f (P), f (Q)
2
=
3v + 4w 3x 4y
5
2
+
4v 3w 4x + 3y
5
2
=
3
2
+ 4
2
5
2
( v x)
2
+ (w y)
2
=
|
PQ
|
2
Isometric segments are certainly congruent. We should make sure the same holds for angles.
Lemma 3.11. Isometries preserve (non-oriented) angles: if f : R
2
R
2
is an isometry, then
PQR
=
f (P) f (Q) f (R)
Proof. Since f is an isometry, the sides of PQR and f (P) f (Q) f (R) are mutually congruent in
pairs. The SSS triangle congruence theorem says that the angles are also mutually congruent.
Example (3.10, cont). Warning: Isometries can reverse orientation!
In the picture,
ABC =
π
2
but f (A) f (B) f (C) =
3π
2
= 2π ABC
A
B
C
f
Our next task is to confirm our intuition that isometries are rotations, reflections and translations.
Given an isometry f , define g(X) = f (X) f (O), where O is the origin. Then g is an isometry
g(P) g(Q) = f (P) f (Q) = d
g(P), g(Q)
= d
f (P), f (Q)
=
|
PQ
|
which moreover fixes the origin: g(O) = O. We conclude that every isometry f is the composition of
an origin-preserving isometry g followed by a translation +C:”
f (X) = g(X) + C
15
In ancient Greek, iso-metros is literally same measure (length/distance).
40
It thus suffices to describe the origin-preserving isometries g. For these, we make two observations.
1. Suppose
|
OQ
|
= 1 and let X
r
= rQ for some r R. Then
g(X
r
) is a distance
|
r
|
=
|
OX
r
|
from the origin O = g(O).
g(X
r
) is a distance
|
1 r
|
=
|
QX
r
|
from g(Q).
g(X
r
) therefore lies on the intersection of two circles, which in-
tersect at a single point: we conclude that
g(rQ) = rg(Q)
The picture shows the case 0 < r < 1, where the uniqueness of
intersection follows from 1 =
|
r
|
+
|
1 r
|
.
|r|
|1 r |
|1 r |
O
Q
X
r
g(Q)
g(X
r
)
2. g(1, 0) lies on the unit circle and therefore has the form
g(1, 0) = S
θ
:=
cos θ, sin θ
for some θ [0, 2π). By preservation of length and angle
(Lemma 3.11), any other point S
ϕ
= (cos ϕ, sin ϕ) on the unit
circle must therefore be mapped to one of two points
g( S
ϕ
) = S
θ±ϕ
=
cos(θ ± ϕ), sin(θ ±ϕ)
The angle ϕ is transferred to one side of the ray
OS
θ
.
0
1
y
0 1
x
S
ϕ
S
θ
S
θ+ϕ
S
θϕ
ϕ
θ
Putting these together by writing X = rS
ϕ
= (r cos ϕ, r sin ϕ) in polar co-ordinates, we conclude that
g has one of two forms:
y
x
X = rS
ϕ
g(X) = rS
θ+ϕ
ϕ
θ
y
x
θ
2
ϕ
θ
2
ϕ
θ
2
ϕ
X = rS
ϕ
g(X) = rS
θϕ
Rotation counter-clockwise by θ Reflection across the line making
angle
θ
2
with positive x-axis
Theorem 3.12. Every isometry of R
2
has the form
f (X) = g(X) + C
where g is either a rotation about the origin, or a reflection across a line through the origin.
41
Calculating with isometries
This benefits from column-vector notation and matrix multiplication. Writing x =
(
x
y
)
=
r cos ϕ
r sin ϕ
for
the position vector of X
r
= (x, y) = rS
ϕ
and applying the multiple-angle formulæ, rotation becomes
g( x) = r
cos(θ + ϕ)
sin(θ + ϕ)
= r
cos θ cos ϕ sin θ sin ϕ
sin θ cos ϕ + cos θ sin ϕ
=
cos θ sin θ
sin θ cos θ
x
For reflections, the sign of the second column is reversed:
cos θ sin θ
sin θ cos θ
. Every isometry therefore has
the form f (x) = Ax + c where A is an orthogonal matrix.
16
Examples 3.13. 1. We revisit Example 3.10 in matrix format:
f (x) =
1
5
3x + 4y
4x 3y
+
3
1
=
1
5
3 4
4 3
x
y
+
3
1
Since
sin θ
cos θ
=
4/5
3/5
=
4
3
, we see that its effect is to reflect across the line through the origin making
angle
1
2
tan
1
4
3
26.6° with the positive x-axis, before translating by (3, 1).
2.
a
has vertices (0, 0), (1, 0), (2, 1) and is congruent to
b
, two of whose vertices are (1, 2) and
(1, 3). Find all isometries transforming
a
to
b
and the location(s) of the third vertex of
b
.
Let f = Ax + c be the isometry. Since d
(1, 2), (1, 3)
= 1 these points must be the images
under f of (0, 0) and ( 1, 0). There are four distinct isometries:
Cases 1, 2: If f (0, 0) = (1, 2) and f (1, 0) = (1, 3), then c = f
0
0
=
1
2
and
A
1
0
+ c =
1
3
= A
1
0
=
0
1
= A =
0 a
12
1 a
22
for some a
12
, a
22
. Since A is orthogonal, the options are A =
0 1
1 0
and
we obtain two possible isometries:
f
1
( x) =
0 1
1 0
x +
1
2
rotates by 90°, then translates by
1
2
.
f
2
( x) =
0 1
1 0
x +
1
2
reflects across y = x, then translates by
1
2
.
The third point of
b
is f
1
(2, 1) = (2, 4) or f
2
(2, 1) = (0, 4).
Cases 3, 4: f (0, 0) = (1, 3) and f ( 1, 0) = (1, 2) results in two further
isometries f
3
and f
4
. The details are an exercise.
All four possible triangles
b
are drawn in the picture.
1
0
1
2
3
4
1 2
a
b
In 1872, Felix Klein suggested that the geometry of a set is the study of its invariants: properties
preserved by its group of structure-preserving transformations. In Euclidean geometry, this is the
group of Euclidean isometries (Exercise 9). Klein’s approach provided a method for analyzing and
comparing the non-Euclidean geometries beginning to appear in the late 1800s. By the mid 1900s, the
resulting theory of Lie groups had largely classified classical geometries. Klein’s algebraic approach
remains dominant in modern mathematics and physics research.
16
An orthogonal matrix satisfies A
T
A = I. All such have the form
cos θ sin θ
sin θ ±cos θ
=
a b
b ±a
where a
2
+ b
2
= 1.
42
Exercises 3.3. 1. Let f : R
2
R
2
be the isometry, “reflect across the line through the origin making
angle
π
3
with the positive x-axis.” Find a 2 ×2 matrix A such that f (x) = Ax.
2. Describe the geometric effect of the isometry f (x) =
1
2
1
3
3 1
x +
3
2
3. Find the remaining isometries f
3
, f
4
and the third points of
b
in Exercise 3.13.2.
4. Find the reflection of the point (4, 1) across the line making angle
1
2
tan
1
12
5
33.7° with the
positive x-axis.
(Hint: if tan θ =
12
5
, what are cos θ and sin θ?)
5. An origin-preserving isometry f (v) = Av moves the point (7, 4) to (1, 8).
(a) If f is a rotation, find the matrix A. Through what angle does it rotate?
(b) If f is a reflection, find the matrix A. Across which line does it reflect?
6. Let ABCD be the rectangle with vertices A = (0, 0), B = (4, 0), C = (4, 3), D = (0, 3). Suppose
an isometry f : R
2
R
2
maps ABCD to a new rectangle PQRS where
P = f (A) := (2, 4) and R = f (C) := (2, 9)
Find all possible isometries f and the remaining points Q = f (B) and S = f (D).
7. (a) If A =
cos θ sin θ
sin θ cos θ
and p is constant, explain why f (x) = A(x p) + p = Ax + (I A)p
rotates by θ around the point with position vector p.
(b) Suppose f (x) = Ax + c rotates the plane around the point P = (2, 1) by an angle θ =
tan
1
3
4
. Find A and c.
(c) Suppose f rotates by θ around p and g rotates by ϕ around q where θ, ϕ are non-zero.
i. If θ + ϕ = 2π, show that f g is a rotation: by what angle and about which point?
ii. What happens instead if θ + ϕ = 2π?
8. Make an argument involving circle intersections (see page 41) to prove that for any isometry f ,
f
(1 t)P + tQ
= (1 t) f (P) + t f (Q)
9. Throughout this question, we use the notation f
A,c
: x 7 Ax + c.
(a) Prove that isometries obey the composition law f
A,c
f
B,d
= f
AB,c+Ad
.
(b) Find the inverse function of the isometry f
A,c
. Otherwise said, if f
A,c
f
C,d
= f
I,0
, where I
is the identity matrix, how do B, d depend on A, c?
(c) Verify that the following composition f
A,c
f
I,d
f
1
A,c
is a translation.
Part (a) can be written using augmented matrices: (A |c)(B |d) := (AB |c + Ad).
If you know group theory, parts (a) and (b) are the closure and inverse properties of the group of Euclidean
isometries E. Part (c) says the translations T form a normal subgroup. We may therefore write E as a
semi-direct product of T and the orthogonal group of origin-preserving isometries
E = T O
2
(R )
43
3.4 The Complex Plane
Complex numbers date to 16
th
century Italy. Their application to geometry really begins with Leon-
hard Euler (1707–1783) who identified the set of complex numbers C with the plane (what is now
known as the Argand diagram).
Definition 3.14. Let i be an abstract symbol satisfying the property
i
2
= 1.
Given real numbers x, y, the complex number z = x + iy is simply the
point (x, y) in the standard Cartesian plane.
17
Given z = x + iy, its:
Complex conjugate z = x iy is its reflection across the real axis.
Modulus
|
z
|
=
zz =
p
x
2
+ y
2
is its distance from the origin.
Argument arg(z) is the angle (measured counter-clockwise) be-
tween the positive real axis and the ray
0z.
y
1 2 3 4
z = 2 3i
z = 2 + 3i
|z | =
13
x
i
2i
3i
i
2i
3i
arg(z) = tan
1
3
2
Addition, scalar multiplication (by real numbers) and complex multiplication follow the usual algebraic
rules while using i
2
= 1 to simplify.
Example 3.15. A simple example of multiplication of complex numbers:
(2 + 3i)(4 + 5i) = 2 ·4 + 2 ·5i + 3i ·4 + 3i ·5i (multiply out)
= 8 + 10i + 12i 15 (use i
2
= 1 to simplify)
= 7 + 22i
The algebra screams geometry! Definition 3.14 already length, angle and reflection in the real axis.
Two other aspects of basic geometry are immediate:
Addition by z translates all points by z.
Scalar multiplication scales distances from the origin (similarity).
The algebraic property distinguishing the complex numbers from the standard Cartesian plane is
complex multiplication. To start visualizing this, consider multiplication by i,
iz = i(x + iy) = y + ix
This is the result of rotating z counter-clockwise
π
2
radians about the origin. To obtain all rotations
and reflections, we need an alternative description of a complex number.
Lemma 3.16. 1. (Euler’s Formula) For any θ R, e
iθ
= cos θ + i sin θ.
2. (Exponential laws) e
iθ
e
iϕ
= e
i(θ+ϕ)
and (e
iθ
)
n
= e
inθ
for any n Z.
Evaluating at θ = π yields the famous Euler identity e
iπ
= 1. Part 1 can be taken as a definition. To
see that it is a reasonable definition requires either power series or elementary differential equations,
topics best described elsewhere. Part 2 is an exercise.
17
In the language of linear algebra, C is a vector space over R with basis {1, i}.
44
Definition 3.17. Let z = x + iy be a non-zero complex number.
Writing x = r cos θ and y = r sin θ, we obtain the polar form
z = re
iθ
= r(cos θ + i sin θ)
where r =
|
z
|
is the modulus and θ = arg(z) the argument of z.
0 1 2
1 +
3i = 2e
iπ
3
2
π
3
i
0
2i
Now consider the effect of multiplying a complex number z = re
iϕ
by e
iθ
= cos θ + i sin θ: according
to the Lemma
e
iθ
z = re
iθ
e
iϕ
= re
i(θ+ϕ)
which has the same modulus (r) as z but a new argument.
Theorem 3.18. The complex number e
iθ
z is the result of rotating z counter-clockwise about the origin
through an angle θ.
Example 3.19. To rotate z = 1 + 2i counter-clockwise by
3π
4
radians, we multiply by
e
3πi
4
= cos
3π
4
+ i sin
3π
4
=
1
2
( 1 + i)
to obtain
e
3πi
4
z =
1
2
( 1 + i)(1 + 2i) =
1
2
(3 + i)
2 1 1
z
e
3πi
4
z
3π
4
i
2i
i
You could try to keep things in polar form, though it doesn’t result in a nice answer:
z =
5e
i tan
1
2
= e
3πi
4
z =
5e
3πi
4
+i tan
1
2
Reflections may be described by combining rotations with complex con-
jugation. To reflect across the line making angle θ with the positive real
axis, we rotate the plane so that the reflection appears to be vertical:
1. Rotate the plane clockwise by θ, that is z 7 e
iθ
z.
2. Reflect across the real axis by complex conjugation.
3. Rotate counter-clockwise by θ.
Combining these steps gives the formula.
ϕ
1
2
3
Theorem 3.20. To reflect z across the line making angle θ with the positive real axis, we compute
z 7 e
iθ
( e
iθ
z) = e
2iθ
z
45
Example 3.21. Reflect z = 2 + 3i across the line through the origin and w =
3 + i.
First compute θ = arg(w) = tan
1
1
3
=
π
6
. The desired point is therefore
e
iπ
3
( 2 3i) =
1
2
+
3
2
i
!
( 2 3i) =
3
3
2
1
!
3 +
3
2
i
To describe general rotations and reflections about arbitrary points/lines, we combine our approach
with translations (compare Exercise 3.3.7).
Corollary 3.22. 1. To rotate z by θ about a point w, compute z 7 e
iθ
( z w) + w.
2. To reflect z across the line with slope θ through a point w, compute z 7→ e
2iθ
( z w) + w.
Example 3.23. The combination of translation by i, rotation by
π
3
around the origin, then translation
by 1, may be expressed
z 7 e
π
3
z i
+ 1 = i + e
π
3
z i
+ 1 i
Alternatively, this is rotation by
π
3
around i followed by translation by 1 i.
We have now described all the Euclidean isometries of the previous section in the language of com-
plex numbers. Here is the full dictionary.
18
Isometry/Transformation Complex numbers Matrices/vectors
Addition/Translation z + w = (x + iy) + (u + iv) z + w =
x
y
+
u
v
Scaling
λz = (λx) + i(λy) λz =
λx
λy
Rotation CCW by
π
2
z 7 iz z 7
0 1
1 0
z
Rotation CCW by θ z 7 e
iθ
z z 7
cos θ sin θ
sin θ cos θ
z
Vertical reflection z 7 z z 7→
1 0
0 1
z
Reflection across line with slope
θ
2
z 7 e
iθ
z z 7
cos θ sin θ
sin θ cos θ
z
It is perhaps surprising to modern readers, but complex numbers came before vectors and matrix-
geometry! During the 1800s mathematicians tried unsuccessfully to replicate the complex number
approach in higher dimensions. This ultimately led (via Hamilton’s quaternions) to the adoption of
vectors and linear algebra/matrix calculations.
One reason for the desire to keep the complex number description is that it may be used to describe
further (non-isometric) transformations of the plane: for instance z 7 z
1
is reflection in a circle! We’ll
discuss some of this at the end of Chapter 4.
18
Scaling isn’t an isometry, but it is worth including nonetheless!
46
Exercises 3.4. 1. Use complex numbers to compute the result of the following transformations: you
can answer in either standard or polar form.
(a) Rotate 3 5i counter-clockwise around the origin by
3π
4
radians.
(b) Reflect 2 i across the line joining 1 + i
3 and the origin.
(c) Reflect 1 + i across the line through the origin making angle
π
5
radians with the positive
real axis.
2. Find the reflection of the point (2, 3) across the line making angle
3π
8
with the positive x-axis.
Give your answer using both complex numbers and matrices/vectors.
3. Repeat the previous question for the point (3, 4) and the angle
5π
12
= 75°.
4. Describe the geometric effect of the map z 7
1
2
( 1 i)
z 3 + 4i
.
(Hint: compare Example 3.23)
5. (Hard) Consider the line through the origin and
p
2 +
2,
p
2
2
. Compute the result
of reflecting 2 + 3i across .
6. By letting n = 3 in Lemma 3.16, prove that
cos 3θ = 4 cos
3
θ 3 cos θ
Find a corresponding trigonometric identity for sin 3θ.
7. Prove part 2 of Lemma 3.16.
(Hint: use the multiple-angle formulae (page 39) to expand e
i(θ+ϕ)
)
47
3.5 Birkhoff’s Axiomatic System for Analytic Geometry (non-examinable)
Analytic geometry was originally conceived as an addition to Euclidean geometry. In 1932, courtesy
of George David Birkhoff, it was axiomatized in its own right.
Background Assume the usual properties/axioms of the real numbers as a complete ordered field.
Birkhoff’s system is typical of modern axiomatic systems in that it is built on top of pre-existing
systems (set theory, complete ordered fields, etc.).
Undefined terms Two objects: Point, line. Two function: distance d, angle measure . If the set of
points is S, then,
d : S ×S R
+
0
, : S ×S × S [ 0, 2π)
Axioms Euclidean Given two distinct points, there exists a unique line containing them.
Ruler Points on a line are in bijective correspondence with the real numbers in such a way that if
t
A
, t
B
correspond to A, B , then
|
t
A
t
B
|
= d(A, B).
Protractor The rays emanating from a point O are in bijective correspondence with the set [0, 2π) so
that if α, β correspond to rays
OA,
OB, then AOB β α (mod 2π). This correspondence is
continuous in A, B.
SAS similarity
19
If triangles have a pair of angles with equal measure, and the sides adjacent to said
angles are in the same ratio, then the remaining angles have equal measure and the final sides
are in the same ratio.
Definitions As with Hilbert, some of these are required before later axioms make sense. In partic-
ular, the definition of ray is required before the protractor axiom.
Betweenness B lies between A and C if d(A, B) + d(B, C) = d(A, C)
Segment AB consists of the points A, B and all those between
Ray
AB consists of the segment AB and all points C such that B lies between A and C.
Basic shapes Triangles, circles, etc.
Analytic Geometry as a Model
The axioms should feel familiar. Being shorter than Hilbert’s list, and being built on familiar notions
such as the real line, it is somewhat easier for us to understand what the axioms are saying and to
visualize them. There is something to prove however; indeed the major point of Birkhoff’s system!
Theorem 3.24. Cartesian analytic geometry is a model of Birkhoff’s axioms.
Recall what this requires: we must provide a definition of each of the undefined terms and prove that
these satisfy each of Birkhoff’s axioms. Here are suitable definitions for Cartesian analytic geometry:
19
As with Hilbert, Birkhoff makes SAS an axiom: Birkhoff’s version is stronger, for it also applies to similar triangles
48
Point An ordered pair (x, y) of real numbers.
Distance d(A, B) =
q
(A
x
B
x
)
2
+ (A
y
B
y
)
2
Line All points satisfying a linear equation ax + by + c = 0.
Angle Define column vectors as differences (v = P O and w = Q O) and consider the matrix
J =
0 1
1 0
. Now define angle via
cos POQ =
v · w
|
v
||
w
|
where POQ
(
[0, π] w · Jv 0
( π, 2π) w · Jv < 0
()
In essence, J is ‘rotate counter-clockwise by
π
2
.’ Cosine may be defined using power series, so
no pre-existing geometric meaning is required.
Proof. (Euclidean axiom) If (x
1
, y
1
) and (x
2
, y
2
) satisfy ax + by + c = 0 then
a(x
1
x
2
) + b(y
1
y
2
) = 0
whence a = y
1
y
2
, b = x
2
x
1
up to scaling. It follows that the line has equation
( y
1
y
2
)x + (x
2
x
1
) y + x
1
y
2
x
2
y
1
= 0
unique up to multiplication of all three of a, b, c by a non-zero constant.
The remaining axioms are exercises.
Exercises 3.5. 1. Prove that the ruler axiom is satisfied:
(a) First show that if P = Q lie on , then any point A on the line has the form
A = P +
t
A
d(P, Q)
(Q P) where t
A
R
(b) Use this formula to verify that d(A, B)
2
= (t
A
t
B
)
2
.
2. Let i =
1
0
. Given any non-zero point B, define b = B O and let β = cos
1
i·b
|
b
|
in accordance
with (). This is a continuous function of b.
(a) If
ˆ
B is any other point on the same ray
OB, explain why we get the same value β.
(β is thus a continuous function of B)
(b) If B = (x, y), what is are values of cos β and sin β?
(c) Suppose A corresponds to α under this identification. Evaluate cos(β α) and therefore
prove that the protractor axiom is satisfied.
3. Use the cosine rule (Theorem 3.7) to prove that the SAS similarity axiom is satisfied.
49
4 Hyperbolic Geometry
4.1 History: Saccheri, Lambert and Absolute Geometry
For 2000 years after Euclid, many mathematicians believed that his parallel postulate could not be
an independent axiom. Rigorous work on this problem was undertaken by Giovanni Saccheri (1667–
1733) & Johann Lambert (1728–1777); both attempted to force contradictions by assuming the nega-
tion of the parallel postulate. While this approach ultimately failed, their insights supplied the foun-
dation of a new non-Euclidean geometry. Before considering their work, we define some terms and
recall our earlier discussion of parallels (pages 10–13).
Definition 4.1. Absolute or neutral geometry is the axiomatic system comprising all of Hilbert’s
axioms except Playfair. Euclidean geometry is therefore a special case of neutral geometry.
A non-Euclidean geometry is (typically) a model satisfying most of Hilbert’s axioms but for which
parallels might not exist or are non-unique:
There exists a line and a point P through which there are no parallels or at least two.
For instance, spherical geometry is non-Euclidean since there are no parallel lines—Hilbert’s axioms
I-2 and O-3 are false, as is the exterior angle theorem.
Results in absolute geometry The conclusions of Euclid’s first 28 theorems are valid.
Basic constructions: bisectors, perpendiculars, etc.
Triangle congruence theorems: SAS, ASA, SAA, SSS.
Exterior angle theorem and its consequences:
m
P
β
α
Side/angle comparison and triangle inequality (Exercise 2.3.5).
Existence of a parallel m to a line through a point P ∈ via congruent angles
α
=
β = m
Arguments making use of unique parallels The following results were proved using Playfair’s
axiom or the parallel postulate, whence the arguments are false in absolute geometry:
A line crossing parallel lines makes congruent angles: in the picture, m = α
=
β. This is
the uniqueness claim in Playfair: the parallel m to through P is unique.
Angles in a triangle sum to 180°.
Constructions of squares/rectangles.
Pythagoras’ Theorem.
While our arguments for the above are false in absolute geometry, we cannot instantly claim that the
results are false, for there might be alternative proofs! To show that these results truly require unique
parallels, we must exhibit a model in which they are false—such will be described in the next section.
The existence of this model explains why Saccheri and Lambert failed in their endeavors; the parallel
postulate (Playfair) is indeed independent of Euclid’s (Hilbert’s) other axioms.
50
The Saccheri–Legendre Theorem
We work in absolute geometry, starting with an extension of
the exterior angle theorem based on Euclid’s proof.
Suppose ABC has angle sum Σ
and construct M and E fol-
lowing Euclid to the arrangement pictured. Observe:
M
A
B
C
E
1. ACB + CAB = ACB + ACE < 180° is the exterior angle theorem. More generally, the
exterior angle theorem says that the sum of any two angles in a triangle is strictly less than 180°.
2. ABC and EBC have the same angle sum
Σ
=
+
+
+
Just look at the picture—remember that we do not know whether Σ
= 180°!
3. EBC has at least one angle (EBC or BEC) measuring
1
2
ABC.
Iterate this construction: if EBC
1
2
ABC, start by bisecting CE; otherwise bisect BC . . . The result
is an infinite sequence of triangles
1
= EBC,
2
,
3
, . . . with two crucial properties:
(a) All triangles have same angle sum Σ
= Σ
1
= Σ
2
= ···.
(b)
n
has at least one angle measuring α
n
1
2
n
ABC.
Now suppose Σ
= 180° + ϵ is strictly greater than 180°. Since lim
1
2
n
= 0, we may choose n large
enough to guarantee α
n
< ϵ. But then the sum of the other two angles in
n
would be greater than
180°, contradicting the exterior angle theorem (observation 1)! We have proved a famous result.
Theorem 4.2 (Saccheri–Legendre). In absolute geometry, triangles have angle sum Σ
180°.
Saccheri’s failed hope was to prove equality without invoking the parallel postulate.
Saccheri and Lambert Quadrilaterals
Two families of quadrilaterals in absolute geometry are named in honor of these pioneers.
Definition 4.3. A Saccheri quadrilateral ABCD satisfies
AD
=
BC and DAB = CBA = 90°
AB is the base and CD the summit.
The interior angles at C and D are the summit angles.
A Lambert quadrilateral has three right-angles; for instance AMND
in the picture.
B
A
M
N
C
D
We draw these with curved sides to indicate that the summit angles need not be right-angles, though
we haven’t yet exhibited a model which shows they could be anything else. Regardless of how they
are drawn, AD, BC and CD are all segments!
51
The apparent symmetry of a Saccheri quadrilateral is not an illusion.
Lemma 4.4. 1. If the base and summit of a Saccheri quadrilateral
are bisected, we obtain congruent Lambert quadrilaterals.
2. The summit angles of a Saccheri quadrilateral are congruent.
3. In Euclidean geometry, Saccheri and Lambert quadrilaterals
are rectangles (four right-angles).
B
A
M
N
C
D
Parts 1 and 2 are exercises. We could interpret part 3 as saying that Saccheri and Lambert quadrilat-
erals are as close as we can get to rectangles in absolute geometry.
Proof of 3. By part 1 we need only prove this for a Saccheri quadrilateral. Following the exterior angle
theorem,
AB is a crossing line making congruent right-angles, whence AD BC.
However
CD also crosses the same parallel lines. By the parallel postulate, the summit angles sum
to a straight edge. Since these are congruent, they are both right-angles.
We now show that drawing acute summit angles is justified by the Saccheri–Legendre Theorem.
Theorem 4.5. The summit angles of a Saccheri quadrilateral measure 90°.
Proof. Suppose ABCD is a Saccheri quadrilateral with base AB.
Extend CB to E (opposite side of AB to C) such that BE
=
DA.
Let M be the midpoint of AB.
SAS implies DAM
=
EBM; the vertical angles at M are con-
gruent, whence M lies on DE.
A
B
C
D
E
M
By Saccheri–Legendre, the (congruent) summit angles at C and D sum to
ADC + DCB = ADM + EDC + DCE = CED + EDC + DCE 180°
Exercises 4.1. Work in absolute geometry; you cannot use Playfair’s Axiom or the parallel postulate!
1. Use the first picture to prove parts 1 and 2 of Lemma 4.4.
2. Use the first picture to give an alternative proof of Theorem 4.5.
3. Suppose ABCD has four right-angles (second picture). Show
that AC splits ABCD into two congruent triangles, and con-
clude that the opposite sides are congruent.
Why is this question easier in Euclidean geometry?
4. Suppose two Saccheri quadrilaterals have congruent bases
(AB) and perpendicular sides (AD, BC). Prove that they are
congruent.
5. (Hard!) Suppose Saccheri quadrilaterals have congruent sum-
mits and perpendicular sides. Prove that the quadrilaterals are
congruent.
B
A
M
N
C
D
A
B
C
D
52
4.2 Models of Hyperbolic Geometry
In the 1820-30s, J
´
anos Bolyai, Carl Friedrich Gauss and Nikolai Lobachevsky independently took the
next step, each describing versions of non-Euclidean geometry.
20
Rather than attempting to establish
the parallel postulate as a theorem within Euclidean geometry, a new geometry was defined based on
the first four of Euclid’s postulates plus an alternative to the parallel postulate:
Axiom 4.6 (Bolyai–Lobachevsky/Hyperbolic Postulate). Given a line and a point P , there
exist at least two parallel lines to through P.
The resulting axiomatic system
21
is called hyperbolic geometry. Consistency was proved in the late
1800s by Beltrami, Klein and Poincar
´
e, each of whom created models by defining point, line, etc., in
novel ways. One of the simplest is named for Poincar
´
e, though was first proposed by Beltrami.
Definition 4.7. The Poincar´e disk is the interior of the unit circle
(x, y) R
2
: x
2
+ y
2
< 1
or
z C :
|
z
|
< 1
A hyperbolic line is a diameter or a circular arc meeting the unit circle at
right-angles.
In the picture we have a hyperbolic line and a point P: also drawn
are several parallel hyperbolic lines to passing through P.
Points on the boundary circle are termed omega-points: these are not in
the Poincar
´
e disk and are essentially ‘points at infinity.’
P
Since it depends only on the incidence axioms, there exists a unique hyperbolic line joining any two
points in the Poincar
´
e disk.
Hyperbolic lines may straightforwardly be described using equations in analytic geometry.
Lemma 4.8. Every hyperbolic line in the Poincar
´
e disk model is
one of the following:
A diameter passing through (c, d) = (0, 0) with Euclidean
equation dx = cy.
The arc of a (Euclidean) circle with equation
x
2
+ y
2
2ax 2by + 1 = 0 where a
2
+ b
2
> 1
and (Euclidean) center and radius
C = (a, b) and r =
p
a
2
+ b
2
1
1
r
O
P
Q
C = (a, b)
20
Bolyai indeed is the source of the term ‘absolute geometry.’
21
We assume all of Hilbert’s axioms, replacing Playfair axiom with the hyperbolic postulate.
53
Example 4.9. We compute the hyperbolic line through P = (0,
1
2
) and Q = (
1
2
,
1
3
) in the Poincar
´
e
disk: this is the picture shown in Lemma 4.8.
Substitute into x
2
+ y
2
2ax 2by + 1 = 0 to obtain a system of equations for a, b:
(
1
4
b + 1 = 0
1
4
+
1
9
a
2
3
b + 1 = 0
= (a, b) =
19
36
,
5
4
The required hyperbolic line
PQ therefore has equation
x
2
+ y
2
19
18
x
5
2
y + 1 = 0 or
x
19
36
2
+
y
5
4
2
=
545
648
The undefined terms point, line, on and between now make sense. To complete the model, we need to
define congruence of hyperbolic segments and angles.
Definition (4.7 continued). The hyperbolic distance between points P, Q in the Poincar
´
e disk is
22
d(P, Q) := cosh
1
1 +
2
|
PQ
|
2
(1
|
P
|
2
)(1
|
Q
|
2
)
!
where
|
PQ
|
is the Euclidean distance and
|
P
|
,
|
Q
|
are the Euclidean
distances of P, Q from the origin.
Hyperbolic segments are congruent if they have the same length.
The angle between hyperbolic rays is that between their tangent lines:
angles are congruent if they have the same measure.
θ
Lemma 4.10. The hyperbolic distance of P from the origin is
d(O, P) = cosh
1
1 +
|
P
|
2
1
|
P
|
2
= ln
1 +
|
P
|
1
|
P
|
Example 4.11. We calculate the sides and angles in the isosceles right-triangle
with vertices O = (0, 0), P = (
1
2
, 0) and Q = (0,
1
2
).
|
P
|
=
1
2
=
|
Q
|
,
|
PQ
|
2
=
1
4
+
1
4
=
1
2
d(O, P) = d(O, Q) = ln
1 +
1
2
1
1
2
= ln 3 = cosh
1
5
3
1.099
d(P, Q) = cosh
1
1 +
2 ·
1
2
(1
1
4
)
2
!
= cosh
1
25
9
1.681
θ
P
Q
O
22
It seems reasonable for hyperbolic functions to play some role in hyperbolic geometry! As a primer:
cosh x =
e
x
+ e
x
2
, sinh x =
e
x
e
x
2
, tanh x =
sinh x
cosh x
=
e
x
e
x
e
x
+ e
x
and cosh
1
x = ln(x +
p
x
2
1)
54
To find the interior angle θ, implicitly differentiate the equation for the hyperbolic line
PQ:
x
2
+ y
2
5
2
x
5
2
y + 1 = 0 =
dy
dx
P
=
4x 5
5 4y
P
=
3
5
= θ = tan
1
3
5
30.96°
By symmetry, we have the same angle at Q. With a right-angle at O, we conclude that the angle sum
is approximately Σ
= 151.93°!
As a sanity check, we compare data for OPQ and the Euclidean triangle with the same vertices
Property Hyperbolic Triangle Euclidean Triangle
Edge lengths 1.099 : 1.099 : 1.681 0.5 : 0.5 : 0.707
Relative edge ratios 1 : 1 : 1.530 1 : 1 : 1.414
Angles 30.06°, 30.96°, 90° 45°, 45°, 90°
The hyperbolic triangle has longer sides and a relatively longer hypotenuse. Moreover, its side lengths
do not satisfy the Pythagorean relation a
2
+ b
2
= c
2
(though cosh a cosh b = cosh c . . .).
The next result is an exercise; it says that distance increases
smoothly as one moves along a hyperbolic line.
Lemma 4.12. Fix P and a hyperbolic line through P. Then the
distance function Q 7→ d(P, Q) maps the set of points on one side
of P differentiably and bijectively onto the interval (0, ).
The Lemma means that hyperbolic circles are well-defined and
look like one expects: the circle of hyperbolic radius δ centered at
P is the set of points Q such that d(P, Q) = δ.
In the picture are several hyperbolic circles and their centers; one has several of its radii drawn.
Observe how the centers are closer (in a Euclidean sense) to the boundary circle than one might
expect: this is since hyperbolic distances measure greater the further one is from the origin.
In fact (Exercise 4.2.5) hyperbolic circles in the Poincar
´
e disk model are also Euclidean circles! Their
hyperbolic radii moreover intersect the circles at right-angles, as we’d expect.
Theorem 4.13. The Poincar
´
e disk is a model of hyperbolic geometry.
Sketch Proof. A rigorous proof would require us to check the hyperbolic postulate and all Hilbert’s
axioms except Playfair. Instead we verify Euclid’s postulates 1–4 and the hyperbolic postulate 5.
1. Lemma 4.8 says we can join any given points in the Poincar
´
e disk by a unique segment.
2. A hyperbolic segment joins two points inside the (open) Poincar
´
e disk. The distance formula in-
creases (Lemma 4.12) unboundedly as P moves towards the boundary circle, so we can always
make a hyperbolic line longer.
3. Hyperbolic circles are defined above.
4. All right-angles are equal since the notion of angle is unchanged from Euclidean geometry.
5. The first picture on page 53 shows multiple parallels!
55
Other Models of Hyperbolic Space: non-examinable
There are several other models of hyperbolic space. Here are three of the most common.
Klein Disk Model This is similar to the Poincar
´
e disk, though lines are chords of the unit circle
(‘Euclidean’ straight lines!) and the distance function is different:
d
K
(P, Q) =
1
2
ln
|
PΘ
||
Q
|
|
P
||
QΘ
|
where , Θ are where the chord
PQ meets the boundary circle.
The cost is that the notion of angle is different. The picture shows
perpendicularity: Given a hyperbolic line find the tangents to where
it meets the boundary circle. Any chord whose extension passes
through the intersection of these tangents is perpendicular to the orig-
inal line. Measuring other angles is difficult!
Θ
P
Q
Gauss’ famous theorem egregium says that this problem is unavoidable; there is no model in which
lines and angles both have the same meaning as in Euclidean geometry.
Poincar´e Half-plane Model Widely used in complex analysis, the
points comprise the upper half-plane (y > 0) in R
2
, while hyperbolic
lines are verticals or semicircles centered on the x-axis
x = constant or (x a)
2
+ y
2
= r
2
and angles are the same as in Euclidean space. The expression for
hyperbolic distance remains horrific! The picture shows several hy-
perbolic lines and a hyperbolic triangle.
y
x
Hyperboloid Model Points comprise the upper sheet (z 1) of the hyperboloid x
2
+ y
2
= z
2
1.
A hyperbolic line is the intersection of the hyperboloid with a plane through the origin. Isometries
(congruence) can be described using matrix-multiplication and hyperbolic distance is relatively easy:
given P = (x, y, z) and Q = (a, b, c), hyperbolic distance is
d(P, Q) = cosh
1
( cz ax by)
Difficulties include working in three dimensions and the fact
that angles are awkward.
The relationship to the Poincar
´
e disk is via projection. Place
the disk in the x, y-plane centered at the origin and draw a
line through the disk and the point (0, 0, 1). The intersec-
tion of this line with the hyperboloid gives the correspon-
dence.
56
Exercises 4.2. Answer all questions within the Poincar
´
e disk model.
1. (a) Find the equation of the hyperbolic line joining P = (
1
4
, 0) and Q = (0,
1
2
).
(b) Find the side lengths of the hyperbolic triangle OPQ where O = (0, 0) is the origin.
(c) The triangle in part (b) is right-angled at O. If o, p, q represent the hyperbolic lengths of
the sides opposite O, P, Q respectively, check that the Pythagorean theorem p
2
+ q
2
= o
2
is false. Now compute cosh p cosh q: what do you observe?
2. Let P =
1
2
,
q
5
12
and Q =
1
2
,
q
5
12
(a) Compute the hyperbolic distances d(O, P), d(O, Q) and d(P, Q), where O is the origin.
(b) Compute the angle POQ.
(c) Show that the hyperbolic line =
PQ has equation
x
2
10
3
x + y
2
+ 1 = 0
(d) Calculate
dy
dx
and hence show that a tangent vector to at P is
15i + 7j. Use this to
compute OPQ.
3. We extend Example 4.11. Let c (0, 1) and label O = (0, 0), P = (c, 0) and Q = (0, c).
(a) Compute the hyperbolic side lengths of OPQ.
(b) Find the equation of the hyperbolic line joining P = (c, 0) and Q = (0, c).
(c) Use implicit differentiation to prove that the interior angles at P and Q measure tan
1
1c
2
1+c
2
.
What happens as c 0
+
and as c 1
?
4. Let 0 < r < 1 and find the hyperbolic side lengths and interior angles of the equilateral triangle
with vertices (r, 0), (
r
2
,
3r
2
) and (
r
2
,
3r
2
).What do you observe as r 0
+
and r 1
?
5. (a) Use the cosh distance formula to prove that the hyperbolic circle of hyperbolic radius
ρ = ln 3 and center C = (
1
2
, 0) in the Poincar
´
e disk has Euclidean equation
x
2
5
2
+ y
2
=
4
25
(b) Prove that every hyperbolic circle in the Poincar
´
e disk is in fact a Euclidean circle.
6. We sketch a proof of Lemma 4.12.
(a) Prove that f (x) = cosh
1
x = ln(x +
x
2
1) is strictly increasing on the interval (1, ).
(b) By part (a), it is enough to show that
|
PQ
|
2
1
|
Q
|
2
increases as Q moves away from P along a
hyperbolic line. Appealing to symmetry, let P = (0, c) lie on the hyperbolic line with
equation x
2
+ y
2
2by + 1 = 0. Prove that
|
PQ
|
2
1
|
Q
|
2
=
( b c)y + bc 1
1 by
and hence show that this is an increasing function of y when c < y <
1
b
.
57
4.3 Parallels, Perpendiculars & Angle-Sums
From now on, all examples will be illustrated within the Poincar
´
e
disk model. Recall (page 50) that we may use anything from ab-
solute geometry; as a sanity check, think through how the picture
illustrates the following result.
Lemma 4.14. Through a point P not on a line there exists a
unique perpendicular to .
We now consider a major departure from Euclidean geometry.
P
B
A
Q
R
M
m
Theorem 4.15 (Fundamental Theorem of Parallels). Given P ∈ , drop
the perpendicular PQ. Then there exist precisely two parallel lines m, n to
through P with the following properties:
1. A ray based at P intersects if and only if it lies between m and n in
the same fashion as
PQ.
2. m and n make congruent acute angles µ with
PQ.
R
Q
P
n
m
µ
Definition 4.16. The lines m, n are the limiting, or asymptotic, parallels to through P. Every other
parallel is an ultraparallel. The angle of parallelism at P relative to is the acute angle µ.
More generally, parallel lines , m are limiting if they ‘meet’ at an omega-point.
The proof depends crucially on ideas from analysis, particularly continuity & suprema. As you read
through, consider how everything except the last line is valid in Euclidean geometry!
Proof. Points R are in continuous bijective correspondence with the real numbers (Lemma 4.12).
It follows that we have a continuous increasing function
f : R (90°, 90°) where f (r) = QPR
By Saccheri–Legendre, ±90° range f . Since dom f = R is an interval, the intermediate value
theorem forces range f to be a subinterval I (90°, 90°).
Given R , transfer QR to the other side of Q to obtain S . By SAS,
QPS = QPR whence I = range f is symmetric: θ I θ I.
Define µ := sup I (0°, 90°] to be the least upper bound; by symmetry,
inf I = µ. Let m and n be the lines making angles ±µ respectively.
Plainly every ray making angle θ (µ, µ) intersects .
Suppose m intersected at M. Let
˜
M lie on the other side of M from
Q. Since f is increasing, we see that QP
˜
M > µ, which contradicts
µ = sup I. It follows that m is parallel to . Similarly n and we have
part 1.
R
S
Q
P
n
m?
M
˜
M
µ
Finally m = n µ = 90°. In such a case there would exist only one parallel to through P,
contradicting the hyperbolic postulate.
58
The picture suggests a bijective relationship between µ and the perpendicular distance. Here it is; we
postpone a simplified argument to Exercise 4.3.3; the full result follows from a discussion of omega-
triangles in the next section.
Corollary 4.17. The perpendicular distance δ = d(P, Q) and the angle of parallelism are related via
cosh δ = csc µ or equivalently tan
µ
2
= e
δ
Examples 4.18. 1. Let be the hyperbolic line x
2
+ y
2
4x + 1 = 0.
Intersect with x
2
+ y
2
= 1 to find =
1
2
,
3
2
and Θ =
1
2
,
3
2
.
By symmetry, the perpendicular from P = (0, 0) to has equation
y = 0 and results in Q = (2
3, 0).
The limiting parallels through P have equations y = ±
3x, from
which the angle of parallelism is µ = tan
1
3 = 60°.
In accordance with Corollary 4.17, we easily verify that
P
Θ
Q
δ = d(P, Q) = ln
1 + (2
3)
1 (2
3)
= ln
3 e
δ
=
1
3
= tan
µ
2
2. We find the limiting parallels and the angle of parallelism when
P =
3
10
,
4
10
and x
2
+ y
2
+ 2x + 4y + 1 = 0
First find the omega-points by intersecting with x
2
+ y
2
= 1:
= (1, 0), Θ =
3
5
,
4
5
Plainly
PΘ is the diameter y =
4
3
x with slope
4
3
.
P
Θ
Q
µ
For
P, substitute into the usual expression x
2
+ y
2
2ax 2by + 1 = 0 and implicitly differ-
entiate:
x
2
+ y
2
+ 2x
13
8
y + 1 = 0 =
dy
dx
P
=
16( 1 + x)
13 16y
P
=
16 ·
7
10
13
64
10
=
56
33
The angle of parallelism is half that between the tangent vectors
33
56
and
3
4
:
µ =
1
2
cos
1
33
56
·
3
4
33
56
3
4
=
1
2
cos
1
5
13
33.69°
Corollary 4.17 can now be used to find the perpendicular distance d(P, Q) = ln
3+
13
2
.
Without the development of later machinery, it is very tricky to compute Q. If you want a serious
challenge, see if you can convince yourself that Q =
93(29+2
117)
1865
,
26(29+2
117)
1865
.
59
Angles in Triangles, Rectangles and the AAA Congruence
We finish this section three important differences between hyperbolic and Euclidean geometry.
Theorem 4.19. In hyperbolic geometry:
1. There are no rectangles (quadrilaterals with four right-angles). In particular, the summit angles
of a Saccheri quadrilateral are acute.
2. The angles in a triangle sum to strictly less than 180°.
3. (AAA congruence) If the angles of ABC and DEF are congruent in pairs, then the triangles
are congruent (ABC
=
DEF).
Note that AAA is a congruence theorem in hyperbolic geometry, not a similarity theorem (compare
with Theorem 2.42). Also revisit the observations on page 50. These results largely show that Euclid’s
arguments making use of the parallel postulate actually require it!
Proof. Given a rectangle ABCD, reflect across CD (Exercise 4.1.4) and repeat to obtain an infinite
family of congruent rectangles. Let P CD and drop perpendiculars to R AB and C
1
as shown.
PRBC is a rectangle: if not, then one of ARPD or PRBC would have angle sum exceeding 360°,
contradicting Saccheri–Legendre (Theorem 4.2). Similarly DPC
1
D
1
is a rectangle.
By Exercise 4.1.3,
BP splits PRBC into a pair of congruent triangles. In particular,
BP crosses CD at
the same angle as it leaves B. We find ourselves in the original configuration: the ray
BP emanates
from the upper-right vertex of DPC
1
D
1
at the same angle as it does for ABCD!
Iterate the process to obtain the picture, each time dropping the perpendicular from P
k
to CD to
produce the equidistant sequence Q
1
, Q
2
, Q
3
, . . . Since CD is finite, this eventually
23
passes D: some
Q
n
lies on the opposite side of
AD. It follows that P
n
BP does also, whence
BP intersects
AD.
A
B
C
D
R
P
P
1
P
2
P
3
P
4
Q
1
Q
2
Q
3
Q
4
C
1
D
1
C
2
D
2
C
3
D
3
Since P CD was generic, we see that any ray based at B on the same side as AD intersects
AD. The
angle of parallelism of B with respect to
AD is therefore 90°, contradicting the hyperbolic postulate.
Parts 2 and 3 are addressed in Exercises 4 and 5.
23
This is the Archimedean property from analysis: a > b = n N such that nb > a.
60
Exercises 4.3. 1. Prove the following in hyperbolic geometry (use Theorem 4.19).
(a) Two hyperbolic lines cannot have more than one common perpendicular.
(b) Saccheri quadrilaterals with congruent summits and summit angles are congruent.
2. Let be the line x
2
+ y
2
4x + 2y + 1 = 0 and drop a perpendicular from O to Q .
(a) Explain why Q has co-ordinates (
2
5
t,
1
5
t) for some t (0, 1).
(b) Show that the hyperbolic distance δ = d(O, Q) of from the origin is ln
1+
5
2
.
(c) = (0, 1) is an omega-point for . Compute the angle of parallelism µ = QO
explicitly and check that cosh δ = csc µ.
3. Suppose P = (0, 0) is the origin, let 0 < r < 1 and Q = (r, 0). Also let be the hyperbolic line
passing through Q at right-angles to PQ.
(a) Find the equation of and prove that its limiting parallels through P have equations
±2ry = (1 r
2
)x
(Hint: what does symmetry tell you about the location of the Euclidean center of ?)
(b) Let µ be the angle of parallelism of P relative to and δ = d(P, Q) the hyperbolic distance.
Prove that cosh δ = csc µ.
(Hint: csc
2
µ = 1 + cot
2
µ = 1 +
1
tan
2
µ
= . . .)
(c) By differentiating, verify the claim that δ and µ are bijectively related.
4. We work in absolute geometry.
(a) Suppose A, B and P are non-collinear and drop the per-
pendicular from P to Q
AB.
If P lies between the perpendiculars , m to
AB through A
and B, prove that Q is interior to AB.
(Hint: show that the other cases are impossible)
(b) Suppose there exists a triangle with angle sum 180°. Show
that there exists a right-triangle with angle sum 180° and
therefore a rectangle.
m
A
B
P
Q
(Since rectangles are impossible in hyperbolic geometry, this proves part 2 of Theorem 4.19)
5. We prove the AAA congruence theorem (Theorem 4.19, part 3).
Suppose ABC and DEF are non-congruent but have angles congruent in pairs. WLOG as-
sume DE < AB. By uniqueness of angle/segment transfer, there exist unique points G AB
and H
AC such that (SAS) DEF
=
AGH.
The picture shows the three possible arrangements.
(a) H is interior to AC.
(b) H = C.
(c) C lies between A and H.
In each case, explain why we have a contradiction.
A
B
C = H?
G
H?
H?
61
4.4 Omega-triangles
Recall that limiting parallels (Definition 4.16) ‘meet’ at an omega-point.
Definition 4.20. An omega-triangle or ideal-triangle is a ‘triangle’ one or
more of whose vertices is an omega-point. At least two of the sides of
an omega-triangle form a pair of limiting parallels.
The three types of omega-triangle depend on how many omega-points
they have. In the picture, PQ has one omega-point, PΘ has
two and ΘΞ three!
P
Θ
Ξ
Q
Amazingly, many of the standard results of absolute geometry also apply to omega-triangles! The
first can be thought of as the AAA congruence theorem where one ‘angle’ is zero.
Theorem 4.21 (Angle-Angle Congruence for Omega-triangles). Suppose AB and PQΘ are
omega-triangles, each with a single omega-point. If the angles are congruent in pairs
AB
=
PQΘ BA
=
QPΘ
then the finite sides of each triangle are also congruent: AB
=
PQ.
It doesn’t really make sense to speak of the ‘infinite’ sides, or the ‘angles’ at omega-points, being
congruent. However, if one defines congruence in terms of isometries (Section 4.6), then this idea is
more reasonable.
Proof. Let T
AB be such that AT
=
PQ. If T = B we are done.
Otherwise, WLOG and for contradiction, assume AB < AT. The hy-
pothesis asserts that the angles at B and T are congruent.
Let M be the midpoint of BT and drop the perpendicular to N
B.
Let L
T be on the opposite side of
BT to N such that TL
=
BN.
We now have Side-Angle-Side data, whence NBM
=
LTM. The
vertical angles at M are congruent, whence M lies on LN and we have
a right-angle(!) at L.
The angle of parallelism of L relative to
B is now NL = 90°, which
contradicts Theorem 4.15.
There are two other possible orientations:
N could lie on the opposite side of B from . In this case SAS is
applied to the same triangles but with respect to the congruent
magenta angles.
A
M
N
L
T
B
In the special case that N = B, the cyan angles are right-angles and the same contradiction
appears: the angle of parallelism of T with respect to
B is 90°.
62
Theorem 4.22 (Exterior Angle Theorem for Omega-Triangles). Suppose PQ has a single
omega-point and let P Q R. Then RQ > QP.
Proof. We show that the two other cases are impossible.
(RQ
=
QP) This is the contradictory arrangement described in
the previous proof: P = T, Q = B, R = A, and the magenta an-
gles cannot be congruent.
(RQ < QP) Transfer the latter to Q to produce a ray
QX interior to
PQ with RQX
=
QP.
Since
Q is a limiting parallel to
P, the Fundamental Theorem
says that
QX intersects
P at some point Y.
We now have PQY contradicting the standard exterior angle the-
orem (RQY
=
QPY).
Y
R
P
Q
X
The final congruence theorem is an exercise based on the previous picture.
Corollary 4.23 (Side-Angle Congruence for Omega-triangles). Suppose PQ and ABΘ have
a single omega-point. If QP
=
BAΘ and PQ
=
AB then PQ
=
ABΘ.
A triangle with one omega-point only has three pieces of data: two finite angles and one finite edge.
The AA and SA congruence theorems say that two of these determine the third.
Other observations
Pasch’s Axiom: Versions of this are theorems for omega-triangles.
If a line crosses a side of an omega-triangle and does not pass through any vertex (including
), then it must pass through exactly one of the other sides.
(Omega Crossbar Thm) If a line passes through an interior point and exactly one vertex (includ-
ing ) of an omega-triangle, then it passes through the opposite side. This is partly embedded
in the proof of Theorem 4.22.
Perpendicular Distance and the Angle of Parallelism: Applied to right-angled omega-triangles, the AA
and SA theorems prove that the angle of parallelism is a bijective function of the perpendicular dis-
tance. Moreover, by transferring the right-angle to the positive x-axis and the other vertex to the
origin, we obtain the arrangement in Exercise 4.3.3, thus completing the proof of Corollary 4.17.
Exercises 4.4. 1. Let PQ be an omega-triangle. Prove that PQ + QP < 180°
2. Let and m be limiting parallels. Explain why they cannot have a common perpendicular.
3. Prove the Side-Angle congruence theorem for omega-triangles with one omega-point.
4. What would an ‘omega-triangle’ look like in Euclidean geometry? Comment on the three re-
sults in this section: are they still true?
63
4.5 Area and Angle-defect
In this section we consider one of the triumphs of Johann Lambert: the relationship between the
sum of the angles in a triangle and its area. We start with a loose axiomatization of area as a relative
measure. Until explicitly stated otherwise, we work in absolute geometry.
Axiom I Two geometric figures have the same area if and only if they may be sub-divided into
finitely many pairs of mutually congruent triangles.
24
Axiom II The area of a triangle is positive.
Axiom III The area of a union of disjoint figures is the sum of the areas of the figures.
Definition 4.24 (Angle defect). Let Σ
be the sum of the angles in a triangle. Measured in radians,
the angle-defect of is π Σ
.
Since triangles in absolute geometry have Σ
π (Theorem 4.2), it follows that
0 π Σ
π
In Euclidean geometry the defect is always zero, while in hyperbolic geometry the defect is strictly
positive (Theorem 4.19). A ‘triangle’ with three omega-points would have defect π.
Lemma 4.25. Angle-defect is additive: If a triangle is split into two sub-
triangles, then the defect of the whole is the sum of the defects of the parts.
This is immediate from the picture:
[
π (α + γ + ϵ)
]
+
[
π (β + δ + ζ)
]
= π (α + β + γ + δ)
since ϵ + ζ = π. Notice that angle-sum is not additive!
C
B
A
Q
γ
β
α
δ
ζ
ϵ
Theorem 4.26 (Area determines angle-sum in absolute geometry). If triangles have the same area,
then their angle-sums are identical.
Of course this trivial in Euclidean geometry where all triangles have the same angle-sum!
Proof. The lemma provides the induction step: if
1
and
2
have the same area, then their interiors
are disjoint unions of a finite collection of mutually congruent triangles:
1
=
n
[
k=1
1,k
and
2
=
n
[
k=1
2,k
where
1,k
=
2,k
Each pair
1,k
,
2,k
has the same angle-defect, whence the angle-defects of
1
and
2
are equal:
defect(
1
) =
n
k=1
defect(
1,k
) =
n
k=1
defect(
2,k
) = defect(
2
)
24
To allow infinitely many infinitesimal sub-triangles would require ideas from calculus and complicate our discussion.
64
Angle-sum determines area in hyperbolic geometry
The converse in hyperbolic geometry relies on a beautiful and reversible construction relating tri-
angles and Saccheri quadrilaterals. The construction itself is valid in absolute geometry, though the
ultimate conclusion that angle-sum determines area is not. If the initial discussion seems difficult,
pretend you are in Euclidean geometry and think about rectangles.
Lemma 4.27. 1. Given ABC, choose a side BC. Bisect the remaining sides at E, F and drop per-
pendiculars from A, B, C to
EF. Then HICB is a Saccheri quadrilateral with base HI.
2. Conversely, given a Saccheri quadrilateral HICB with summit BC, let A be any point such that
HI bisects AB at E. Then the intersection F =
HI AC is the midpoint of AC.
Both constructions yield the same picture and the following
conclusions:
The triangle and quadrilateral have equal area.
The sum of the summit angles of the quadrilateral
equals the angle sum of the triangle.
We chose BC to be the longest side of ABC—this isn’t nec-
essary, though it helpfully forces E, F to lie between H, I.
A
B
C
D
E
F
G
H I
Proof. 1. Two applications of the SAA congruence (follow the arrows!) tell us that
BEH
=
AEG and CFI
=
AFG
We conclude that BH
=
AG
=
CI whence HICB is a Sac-
cheri quadrilateral. The area and angle-sum correspon-
dence is immediate from the picture.
2. Suppose the midpoint of AC were at J = F. By part 1, we
may create a new Saccheri quadrilateral with summit BC
using the midpoints E, J.
The perpendicular bisector of BC bisects the bases of both
Saccheri quadrilaterals (Lemma 4.4), creating EUV
with two right-angles: contradiction.
J
U
V
A
B
C
D
E
F
G
H I
We now prove a special case of the main result.
Lemma 4.28. Suppose hyperbolic triangles ABC and PQR have congruent sides BC
=
QR and
the same angle-sum. Then the triangles have the same area.
Proof. Construct the quadrilaterals corresponding to ABC and PQR with summits BC
=
QR.
These have congruent summits and summit angles: by Exercise 4.3.1b they are congruent.
The final observation is what makes this special to hyperbolic geometry. In the Euclidean case, Saccheri
quadrilaterals are rectangles, and congruent summits do not force congruence of the remaining sides.
65
Theorem 4.29. In hyperbolic geometry, if ABC and PQR have the same angle-sum then they
have the same area.
Proof. If the triangles have a pair of congruent edges, the previous result says we are done. Other-
wise, we use Lemma 4.27 to create a new triangle LBC which matching the same Saccheri quadri-
lateral as ABC.
WLOG suppose
|
AB
|
<
|
PQ
|
and construct the Saccheri quadrilateral with summit BC. Select K on
EF such that
|
BK
|
=
1
2
|
PQ
|
and extend such that K is the midpoint of BL.
By Lemma 4.27,
Area(LBC) = Area(HICB) = Area(ABC)
By Theorem 4.26, LBC has the same angle-sum as
ABC and thus PQR.
LBC and PQR share a congruent side (LB
=
PQ)
and have the same angle-sum. Lemma 4.28 says their
areas are equal.
A
B
C
D
E
F
G
H
I
L
K
Since both area and angle-defect are additive, we immediately conclude:
Corollary 4.30. The angle-defect of a hyperbolic triangle is an additive function of its area. By
normalizing the definition of area,
25
we may conclude that
π Σ
= Area
Note finally how the AAA congruence (Theorem 4.19, part 3) is related to the corollary:
ABC
=
DEF
AAA
angles congruent in pairs
equal area
Cor 4.30
same angle-defect
25
We have really only proved that π Σ
is proportional to Area . However, it can be seen that these quantities are
equal if we use the area measure arising naturally from the hyperbolic distance function (see page 79).
Corollary 4.30 is a special case of the famous Gauss–Bonnet theorem from differential geometry: for any triangle on a
surface with Gauss curvature K, we have
Σ
π =
ZZ
K dA
We’ve now met all three special constant-curvature examples of this:
Euclidean space is flat (K = 0) so the angle-defect is always zero.
Hyperbolic space has constant negative curvature K = 1, whence
RR
dA = (Σ
π) is the angle-defect.
Spherical geometry A sphere of radius 1 has constant positive curvature K = 1 and
RR
dA is the angle-excess Σ
π.
66
Example (4.11, cont). The isosceles right-triangle with vertices O, P = (
1
2
, 0) and Q = (0,
1
2
) has
angle-sum and area
π
2
+ 2 tan
1
3
5
151.93° = area = π
π
2
+ 2 tan
1
3
5
=
π
2
2 tan
1
3
5
0.490
A Euclidean triangle with the same vertices has area
1
2
·
1
2
·
1
2
=
1
8
= 0.125.
Generalizing this (Exercise 4.2.3), the triangle with vertices O, P =
( c, 0) and Q = (0, c) has area
π
π
2
+ 2 tan
1
1 c
2
1 + c
2
=
π
2
2 tan
1
1 c
2
1 + c
2
As expected, lim
c0
+
area(c) = 0. In the other limit, the triangle becomes
an omega-triangle with two omega-points and lim
c1
area(c) =
π
2
: an
infinite ‘triangle’ with finite ‘area’!
O
The limit c 1
Our discussion in fact provides an explicit method for cutting a triangle into sub-triangles and rear-
ranging its pieces to create a triangle with equal area.
A
B
C
L
K
S
1
S
3
1
3
P
R
Q
S
2
2
Suppose
1
and
2
have equal area and construct the quadrilaterals S
1
and S
2
. Let L, K be chosen
so that BL
=
QR and K is the midpoint of BL. We now have:
1
,
2
,
3
, S
1
, S
2
, S
3
have the same area.
The summit angles of S
1
, S
2
, S
3
are congruent (half the angle-sum of each triangle).
S
2
, S
3
are congruent since they have congruent summits and summit angles.
We can now follow the steps in Lemma 4.27 to transform
1
to
2
:
1
S
1
3
S
3
=
S
2
2
where each arrow represents cutting off two triangles and moving them. Indeed this works even for
triangles in Euclidean geometry!
67
Exercises 4.5. 1. Use Corollary 4.30 to find the area of the hyperbolic triangle with given vertices.
Your answers to exercises from Section 4.2 should supply the angles!
(a) O = (0, 0), P = (
1
2
,
q
5
12
) and Q = (
1
2
,
q
5
12
).
(b) O = (0, 0), P = (
1
4
, 0), Q = (0,
1
2
).
(c) P = (r, 0), Q =
r
2
,
3r
2
, R =
r
2
,
3r
2
where 0 < r < 1.
2. In the proof of Theorem 4.29, explain why we can find K such that
|
BK
|
=
1
2
|
PQ
|
.
3. Show that there is no finite triangle in hyperbolic geometry that achieves the maximum area
bound π.
(Hard!) For a challenge, try to prove that omega-triangles also satisfy the angle-defect formula:
Area = π Σ
, so that only triangles with three omega-points have maximum area.
4. Let
1
, . . . ,
n
be n distinct omega-points arranged counter-clockwise around the boundary
circle of the Poincar
´
e disk. A region is bounded by the n hyperbolic lines
1
2
,
2
3
, . . . ,
n
1
What is the area of the region? Hence argue that the ‘area’ of hyperbolic space is infinite.
5. An omega-triangle has vertices O = (0, 0), = (1, 0) and P = (0, h) where h > 0.
(a) Prove that the hyperbolic segment P is an arc of a circle with equation
(x 1)
2
+ (y k)
2
= k
2
for some k > 0.
(b) Prove that the area of OP is given by
A(h) = sin
1
2h
1 + h
2
6. Suppose two Saccheri quadrilaterals in hyperbolic geometry have the same area and congruent
summits. Prove that the quadrilaterals are congruent.
68
4.6 Isometries and Calculation
There are (at least!) two major issues in our approach to hyperbolic geometry.
Calculations are difficult In analytic (Euclidean) geometry we typically choose the origin and orient
axes to ease calculation. We’d like to do the same in hyperbolic geometry.
We assumed too much We defined distance, angle and line separately, but these concepts are not inde-
pendent! In Euclidean geometry, the distance function, or metric, defines angle measure via the
dot product,
26
and (with some calculus) the arc-length of any curve. One then proves that the
paths of shortest length (geodesics) are straight lines: the metric defines the notion of line!
Isometries provide a related remedy for these issues. To describe these it is helpful to use an alterna-
tive definition of the Poincar
´
e disk and its distance function.
Definition 4.31. The Poincar´e disk is the set D := {z C :
|
z
|
< 1}
equipped with the distance function
d( z, w) :=
ln
|
z
||
w Θ
|
|
z Θ
||
w
|
where , Θ are the omega-points for the hyperbolic line through z, w
(defined as circular arcs intersecting the boundary perpendicularly).
z
w
Θ
We’ll see shortly (Corollary 4.38) that this is the same as the original cosh formula (page 54); it is
already easy to check that d(z, 0) = ln
1+
|
z
|
1
|
z
|
as in Lemma 4.10 (if w = 0, then , Θ = ±
z
|
z
|
).
For candidate isometries we need functions f : D D for which d
f (z), f (w)
= d(z, w). These
follow from some standard results of complex analysis that we state without proof.
Theorem 4.32 (M¨obius/fractional-linear transformations). If a, b, c, d C and ad bc = 0, then the
function f (z) =
az+b
cz+d
has the following properties:
1. (Invertibility) f : C {} C {} is bijective, with inverse f
1
( z) =
dzb
cz+a
.
2. (Conformality) If curves intersect, then their images under f intersect at the same angle.
3. (Line/circle preservation) Every line/circle
27
is mapped by f to another line/circle.
4. (Cross-ratio preservation) Given distinct z
1
, z
2
, z
3
, z
4
, we have
f (z
1
) f (z
2
)
f (z
3
) f (z
4
)
f (z
2
) f (z
3
)
f (z
4
) f (z
1
)
=
( z
1
z
2
)(z
3
z
4
)
( z
2
z
3
)(z
4
z
1
)
26
Writing
|
u
|
=
|
PQ
|
for the length of a line segment, we see that for any u, v,
u ·v =
1
2
|
u + v
|
2
|
u
|
2
|
v
|
2
so that the metric defines the dot product. Now define angle measure via u ·v =
|
u
||
v
|
cos θ.
27
In C {} a line is just a circle containing !
69
The isometries of the Poincar
´
e disk are a subset of the M
¨
obius transformations.
Theorem 4.33. The orientation-preserving
28
isometries of the Poincar
´
e disk have the form
f (z) = e
iθ
α z
αz 1
where
|
α
|
< 1 and θ [0, 2π) ()
All isometries can be found by composing f with complex conjugation (reflection in the real axis).
Referring to the properties in Theorem 4.32:
1. The isometries are precisely the set of M
¨
obius transformations which map D bijectively to itself;
omega-points are also mapped to omega-points.
2. Isometries preserve angles.
3. The class of hyperbolic lines is preserved: any circle or line intersecting the unit circle at right-
angles is mapped to another such (angle-preservation is used here).
4. If , Θ are the omega-points on
zw, then (by 2 and 3), f () and f ( Θ) are the omega-points for
the hyperbolic line through f (z), f (w). Preservation of the cross-ratio says that f is an isometry:
d( f (z), f (w)) =
ln
|
f (z) f ()
||
f (w) f (Θ)
|
|
f (z) f (Θ)
||
f (w) f ()
|
=
ln
|
z
||
w Θ
|
|
z Θ
||
w
|
= d(z, w)
How does this help us compute? The isometry () moves α to the origin, where calculating distances
and angles is easy!
Example 4.34. Let P =
1
2
and Q =
2
3
+
2
3
i. Move P to the origin using
an isometry
29
with α = P:
f (z) =
α z
αz 1
=
1 2z
z 2
= f (P) = O
f (Q) =
1
4
3
2
2
3
i
2
3
2 +
2
3
i
=
1 + 2
2i
4 +
2i
=
i
2
Let us compare distances:
P
Q
O = f (P)
f (Q)
f
d
f (P), f (Q)
= ln
1 +
|
f (Q)
|
1
|
f (Q)
|
= ln
1 +
1
2
1
1
2
= ln
2 + 1
2 1
= ln(3 + 2
2) (Definition 4.31)
d(P, Q) = cosh
1
1 +
2
|
PQ
|
2
(1
|
P
|
2
)(1
|
Q
|
2
)
!
= cosh
1
1 +
2
4
(1
1
4
)(1
2
3
)
!
= cosh
1
3 = ln(3 +
p
3
2
1) = ln(3 + 2
2) = d
f (P), f (Q)
If we trust the original cosh-formula (page 54), then the points really are the same distance apart!
Indeed the hyperbolic segment PQ has been transformed by f to a segment f (P) f (Q) of the y-axis.
28
If C is to the left of
AB, then f (C) is to the left of
f (A) f (B). This is the usual ‘right-hand rule.’
29
We could also include a rotation (e
iθ
= i) to move f (Q) to the positive x-axis, but there is no real benefit.
70
Recall (e.g., Example 4.11) how we previously computed angles. Isometries make this much easier.
Example 4.35. Given A =
i
2
, B =
i
5
and C =
1
5
(3 i), we find d(A, B), d(A, C) and BAC.
Start by moving A to the origin and consider f (B), f (C):
f (z) =
i
2
z
i
2
z 1
=
2z + i
2 iz
= f (B) =
2i
5
+ i
2
1
5
=
i
3
f ( C) =
2
5
(3 i) + i
2
i
5
(3 i)
=
2( 3 i) + 5i
10 i(3 i)
=
2 + i
3 i
=
(2 + i)(3 + i)
10
=
1 + i
2
By mapping A to the origin, two sides of the triangle are now Eu-
clidean straight lines and the computations are easy:
d(A, B) = d
O, f (B)
= ln
1 +
1
3
1
1
3
= ln 2
d(A, C) = d
O, f (C)
= ln
1 +
1
2
1
1
2
= 2 ln(
2 + 1)
BAC = f (B) f (A) f (C) = arg
i
3
arg
1 + i
2
=
π
2
π
4
=
π
4
To compute the final side and angles, isometries moving B and then
C to the origin could be used.
A
B
C
O = f (A)
f (B)
f (C)
Interpretation of Isometries (non-examinable)
As in Euclidean geometry, isometries can be interpreted as rotations, reflections and translations.
Here is the dictionary in hyperbolic space.
Translations Move α to the origin via T
α
( z) =
αz
αz1
The picture shows repeated applications of T
α
to seven initial points.
Compose these to translate α to β:
T
β
T
α
( z) =
( αβ 1)z + α β
( α β)z + αβ 1
Rotations R
θ
( z) = e
iθ
z rotates counter-clockwise around the origin. To ro-
tate around α, one computes the composition
T
α
R
θ
T
α
The picture shows repeated rotation by 30° =
π
6
around α.
Reflections P
θ
( z) = e
2iθ
z reflects across the line making angle θ with the real
axis. Composition permits more general reflections, e.g.,
T
α
P
θ
T
α
α
O
α
71
Hyperbolic Trigonometry
By employing isometries in the abstract, we can develop formulas
30
allowing us to solve triangles
without the explicit requirement to compute with isometries at all!
Given a right-triangle, we may suppose an isometry has already moved the right-angle to the origin
and the other sides to the positive axes as in the picture. The non-hypotenuse side-lengths are
a = ln
1 + p
1 p
= cosh
1
1 + p
2
1 p
2
, b = cosh
1
1 + q
2
1 q
2
To measure the hypotenuse, translate p to the origin via an isometry
f (z) =
p z
pz 1
= f (iq) =
p iq
ipq 1
=
|
f (iq)
|
2
=
p
2
+ q
2
p
2
q
2
+ 1
a
b
c
B
A
b
a
c
B
A
p
iq
0
f (p)
f (iq)
We therefore see that
cosh c =
1 +
|
f (iq)
|
2
1
|
f (iq)
|
2
=
1 + p
2
q
2
+ p
2
+ q
2
1 + p
2
q
2
p
2
q
2
=
1 + p
2
1 p
2
·
1 + q
2
1 q
2
= cosh a cosh b
Moreover, applying the hyperbolic identity sinh
2
b = cosh
2
b 1, we obtain
sinh b =
2q
1 q
2
= tanh b =
sinh b
cosh b
=
2q
1 + q
2
Writing f (iq) in real and imaginary parts allows us to find the slope
f (iq) =
p iq
ipq 1
=
p(1 + q
2
) + iq(1 p
2
)
p
2
q
2
+ 1
= tan B =
q(1 p
2
)
p(1 + q
2
)
=
tanh b
sinh a
Applying trigonometric identities such as csc
2
B = 1 + cot
2
B, we eventually conclude:
Theorem 4.36. In a hyperbolic right-triangle with adjacent a, opposite b, and hypotenuse c,
sin B =
sinh b
sinh c
cos B =
tanh a
tanh c
tan B =
tanh b
sinh a
cosh c = cosh a cosh b
This last is Pythagoras’ Theorem for hyperbolic right-triangles.
Example 4.37. A right-triangle has non-hypotenuse sides a = cosh
1
3 1.76, b = cosh
1
5 2.29.
cosh c = cosh a cosh b = 15 = c = cosh
1
15 3.40
sin A =
sinh a
sinh c
=
s
cosh
2
a 1
cosh
2
c 1
=
s
3
2
1
15
2
1
=
1
2
7
= A 10.9°
sin B =
sinh b
sinh c
=
s
cosh
2
b 1
cosh
2
c 1
=
s
5
2
1
15
2
1
=
r
3
28
= B 19.1°
You could use the other trig expressions to calculate: e.g., tan A =
tanh a
sinh b
=
sinh a
cosh a sinh b
=
8
3
24
=
1
3
3
. . .
30
Treat the formulas of hyperbolic trigonometry as open-book—they are not worth memorizing!
72
The goal of trigonometry is to ‘solve’ triangles: given minimal numerical data, to compute the re-
maining sides and angles. As in Euclidean geometry, you can attack general problems by dropping
perpendiculars and using the results of Theorem 4.36, though it is helpful to generalize this by de-
veloping the sine and cosine rules.
Corollary 4.38 (Sine/Cosine rules and the Cosh-distance formula). Label a general triangle with
angle-measures A, B, C opposite sides with (hyperbolic) lengths a, b, c.
Sine Rule Drop a perpendicular from C and observe that sin A =
sinh h
sinh b
and
sin B =
sinh h
sinh a
. Eliminate sinh h to obtain the first equality in
sinh a
sin A
=
sinh b
sin B
=
sinh c
sin C
Drop a different altitude for the other equality.
a
b
c
h
B
A
Cosine Rule I Repeat the argument of Theorem 4.36 for a triangle with vertices 0, p and qe
iC
to obtain
cosh c = cosh a cosh b sinh a sinh b cos C
Expressing the right-hand side in terms of p, q (cosh a =
1+p
2
1p
2
, etc.) and applying the Euclidean
cosine rule (cos C = ···) yields the original cosh-formula for distance (page 54).
Cosine Rule II Hyperbolic geometry admits a second version:
cos C = sin A sin B cosh c cos A cos B
A proof is in Exercise 14.
In hyperbolic geometry, the triangle congruence theorems (SAS, ASA, SSS, SAA and AAA) pro-
vide suitable minimal data. The second version of the cosine rule has no analogue in Euclidean
geometry—it is particularly helpful for solving triangles given ASA or AAA data.
31
Examples 4.39. 1. (SAS) An isosceles triangle has angle C =
π
3
and sides a = b = cosh
1
2 1.32.
We have sinh a = sinh b =
p
cosh
2
a 1 =
3. By the cosine rule,
cosh c = 2 ·2
3
3 ·
1
2
=
5
2
= c = cosh
1
5
2
1.57
Apply the sine rule for the final angle:
sin B = sin A =
sin C sinh a
sinh c
=
3
2
·
3
21/4
=
3
21
=
r
3
7
= A = B 40.9°
The area of the triangle is therefore π
π
3
2 sin
1
q
3
7
0.67
cosh
1
2
cosh
1
2
c
A
B
π
3
31
Unlike in Euclidean geometry, knowing two angles doesn’t automatically give you the third! For SAS and SSS start
with the cosine rule. SAA data is best solved by dropping a perpendicular and using Theorem 4.36.
73
2. (Equilateral AAA) An equilateral triangle has interior angles 30°. We compute its side-length
using the second version of the cosine rule:
cosh c =
cos A cos B + cos C
sin A sin B
=
3
2
·
3
2
+
3
2
1
2
·
1
2
= 3
3
= a = b = c = cosh
1
(3
3) 2.33
a
b
c
30°
30°
30°
3. (Right-angled AAA) A triangle has angles
π
6
,
π
4
and
π
2
: find its sides.
Rather than using the second version of the cosine rule, we instead
indicate part of its proof by employing the tan-formula twice,
1
3
= tan
π
6
=
tanh a
sinh b
=
sinh a
cosh a sinh b
1 = tan
π
4
=
tanh b
sinh a
=
sinh b
sinh a cosh b
Multiply together and use hyperbolic Pythagoras,
a
c
b
π
4
π
6
1
3
=
1
cosh a cosh b
=
1
cosh c
= c = cosh
1
3 = ln(
3 +
2) 1.15
Since sinh c =
p
cosh
2
c 1 =
2, the sine-rule yields the other sides:
sinh b = sin
π
4
·
sinh c
sin π
= 1 = b = sinh
1
1 = cosh
1
2 0.88
= cosh a =
cosh c
cosh b
=
r
3
2
= a = cosh
1
r
3
2
0.66
4. (ASA) Solve a triangle with angles
π
6
,
π
3
and a distance cosh
1
3 between them.
Apply the second version of the cosine rule:
cos C = sin A sin B cosh c cos A cos B
=
1
2
·
3
2
·3
3
2
·
1
2
=
3
2
= C =
π
6
The triangle is isosceles, whence a = cosh
1
3 1.76 also.
a
b
cosh
1
3
π
3
C
π
6
The remaining side can be found using multiple methods: here is the cosine rule
cosh b = cosh a cosh c sinh a sinh c cos B = 9
p
3
2
1
2
1
2
= 5
= b = cosh
1
5 = ln(5 +
24) 2.29
74
Hyperbolic Tilings (just for fun!)
Example 4.39.3 can be used to make a regular tiling of hyperbolic space.
Take eight congruent copies of the triangle and arrange
them around the origin as in the picture. Now reflect
the quadrilateral over each of its edges and repeat the
process in all directions. We obtain a regular tiling of
hyperbolic space comprising four-sided figures with six
meeting at every vertex!
In hyperbolic space, many different regular tilings are
possible. Suppose such is to be made using regular m-
sided polygons, n of which are to meet at each vertex:
each polygon comprises 2m copies of the fundamental
right-triangle, whose angles are therefore
π
2
,
π
m
and
π
n
.
Since the angles sum to less than π radians, we see that
there exists a regular tiling of hyperbolic space when-
ever m, n satisfy
π
2
+
π
m
+
π
n
< π (m 2)(n 2) > 4
The first example is m = 4 and n = 6, where the fun-
damental triangle is clear. In the second example four
pentagons meet at each vertex and the interiors of the
polygons have been colored. This was produced using
the tools found here and here: have a play!
The multitude of possible tilings in hyperbolic geome-
try is in contrast to Euclidean geometry, where a regular
tiling requires equality
( m 2)(n 2) = 4
The three solutions (m, n) = (3, 6), (4, 4), (6, 3) cor-
respond to the only tilings of Euclidean geometry by
regular polygons (equilateral triangles, squares and
hexagons). However, all can be scaled to arbitrary side-
lengths. In hyperbolic geometry, there are infinitely
many distinct tilings, but each has a unique side-length.
For related fun, look up M.C. Escher’s Circle Limit art-
works, some of which are based on hyperbolic tilings.
If you want an excuse to play video games while pre-
tending to study geometry, have a look at Hyper Rogue,
which relies on (sometimes irregular) tilings.
π
m
π
n
The fundamental triangle
( m, n) = (4, 6)
( m, n) = (5, 4)
75
Exercises 4.6. 1. Use Definition 4.31 to prove that d(z, 0) = ln
1+
|
z
|
1
|
z
|
.
2. (a) Use an isometry to find angle ABC when A = 0, B =
i
2
, and C =
1+i
2
.
(b) Now compute ACB, and thus find the angle sum and area of the triangle.
3. Associate a M
¨
obius transformation f (z) =
az+b
cz+d
with the matrix
a b
c d
in the obvious way. If g
is another M
¨
obius transformation, prove that the composition f g is associated to the product
of the matrices associated to f , g. Verify
32
that f
1
( z) =
dzb
acz
.
4. (a) A triangle has vertices A =
1
3
, B =
1
2
and C, where C lies in the upper half-plane (positive
imaginary part) such that BAC = 45° and b = d(A, C) = cosh
1
3
Compute a = d(B, C) using the hyperbolic cosine rule.
(b) The isometry f (z) =
1
3
z
1
3
z1
=
13z
z3
moves A to the origin. What is f (B) and therefore f (C)?
(Hint: remember that f is orientation preserving)
(c) Use the inverse of the isometry f to compute the co-ordinates of C. As a sanity-check, use
the cosh distance formula to recover your answer to part (a).
5. Suppose f (z) =
αz
αz1
for some constant α C with
|
α
|
= 1. If
|
z
|
= 1, prove that
|
f (z)
|
= 1.
Argue that the functions f in Theorem 4.33 really do map the interior of the unit disk to itself.
6. Provided α = β, show that the isometry T
β
T
α
which translates α to β (page 71) is the trans-
lation T
γ
where γ =
βα
αβ1
followed by a rotation around the origin.
7. Use the power series cosh x = 1 +
1
2
x
2
+
1
4!
x
4
+ ··· to expand the hyperbolic Pythagorean
theorem cosh c = cosh a cosh b to order 4 (a
4
, a
2
b
2
, etc.). What do you observe?
8. A hyperbolic right-triangle has non-hypotenuse sides a = cosh
1
2 and b = cosh
1
3. Find the
hypotenuse, the angles and the area of the triangle.
9. Given ASA data c = cosh
1
(
2 +
3) , A =
π
4
, B =
π
6
, find the remaining data for the triangle.
10. An equilateral hyperbolic triangle has side-length a and angle A. Prove that cosh a =
cos A
1cos A
.
If A = 45°, what is the side-length?
11. Find the interior angles and side-lengths for the quadrilateral and pentagonal tiles on page 75.
12. Use the hyperbolic cosine rule to prove that the cosh distance formula is valid.
13. Suppose you are given isosceles ASA data: angles A = B and side c between them. Prove that
c cosh
1
(2 csc
2
A 1). What happens when this is equality?
14. (a) Prove the second cosine rule when C =
π
2
(see the trick in Example 4.39.3).
(b) (Hard!) Prove the full version by dropping a perpendicular from B = B
1
+ B
2
and observ-
ing that
cos A
sin B
1
=
cos C
sin B
2
=
cos C
sin(BB
1
)
. . .
32
Since multiplying a, b, c, d by a non-zero scalar doesn’t change f , we see that the group of M
¨
obius transformations is
isomorphic to the projective special linear group PSL
2
(R). The isometries of hyperbolic space form a proper subgroup.
76
The Poincar´e Disk for Differential Geometers (non-examinable)
Most of this last optional section should be accessible to anyone who’s taken basic vector-calculus.
All we really need is the Poincar
´
e disk model with its distance function d(z, w) and a description of
the isometries (Theorems 4.32, 4.33).
Consider the infinitesimally separated points z and z + dz. Map z to
the origin via an isometry
f : ξ 7→
z ξ
zξ 1
Then z + dz is mapped to
P := f (z + dz) =
dz
z( z + dz) 1
=
dz
1
|
z
|
2
z + dz
z + dw
z
0
P
Q
f
where we deleted z dz since it is infinitesimal compared to 1
|
z
|
2
.
Since isometries preserve length and angle, this construction has several consequences.
Infinitesimal distance, arc-length, and geodesics
The hyperbolic distance from z to z + dz is
d( z, z + dz) = d(O, P) = ln
1 +
|
P
|
1
|
P
|
= ln(1 +
|
P
|
) ln(1
|
P
|
) = 2
|
P
|
=
2
|
dz
|
1
|
z( t)
|
2
()
where the approximation ln(1 ±
|
P
|
) = ±
|
P
|
is used since
|
P
|
is infinitesimal. If z(t) parametrizes a
curve in the disk, then the infinitesimal distance formula allows us to compute its arc-length
Z
t
1
t
0
2
|
z
( t)
|
1
|
z( t)
|
2
dt
Example 4.40 (Circles and ‘hyperbolic π’). Suppose that a circle has hyperbolic radius δ. By moving
its center to the origin via an isometry, we may parametrize in the usual manner:
z( t) = r
cos θ
sin θ
, θ [0, 2π) where δ = ln
1 + r
1 r
equivalently r =
e
δ
1
e
δ
+ 1
Its circumference (hyperbolic arc-length) is then
Z
2π
0
2r
1 r
2
dθ =
4πr
1 r
2
= 2π sinh δ = 2π
δ +
1
3!
δ
3
+
1
4!
δ
5
+ ···
> 2πδ
where we used the Maclaurin series to compare.
A hyperbolic circle has a larger circumference : diameter ratio than for a Euclidean circle (π). More-
over, this ratio is not constant: one might say that the hyperbolic version of π is a function (
π sinh δ
δ
)!
77
Our arc-length integral approach also allows us to show that hyperbolic lines are really what we
want them to be: lines of shortest distance between points.
Theorem 4.41. The geodesics—paths of minimal length between two points—in the Poincar
´
e disk
are precisely the hyperbolic lines.
Following the comments on page 69, the distance function really does define the concept of hyper-
bolic line.
Proof. First suppose b lies on the positive x-axis. Parametrize a curve from 0 to b via
z( t) = x(t) + iy(t) where 0 t 1, z(0) = 0, z(1) = b
Its arc-length satisfies
Z
1
0
2
|
z
( t)
|
1
|
z( t)
|
2
dt =
Z
1
0
2
p
x
2
+ y
2
1 x
2
y
2
dt
Z
1
0
2
|
x
|
1 x
2
dt
Z
1
0
2x
( t)
1 x(t)
2
dt =
Z
b
0
2 dx
1 x
2
= ln
1 + b
1 b
= d(0, b)
where we have equality if and only if y(t) 0 and x(t) is increasing. The length-minimizing path is
therefore along the x-axis.
More generally, given points A, B, apply an isometry f such that f (A) = 0 and f (B) = b lies on the
positive x-axis. The geodesic from A to B is therefore the image of the segment 0b under the inverse
isometry f
1
. By the properties of M
¨
obius transforms, this is an arc of a Euclidean circle through A, B
intersecting the unit circle at right-angles, our original definition of a hyperbolic line.
Area Computation
If dx and idy are infinitesimal horizontal and vertical changes in z = x + iy, then the area of the
infinitesimal rectangle spanned by z z + dx and z z + idy is the area element
dA =
2 dx
1
|
z
|
2
2 dy
1
|
z
|
2
=
4 dx dy
(1 x
2
y
2
)
2
The area of a region R in the Poincar
´
e disk is therefore given by the double integral
ZZ
R
4 dxdy
(1 x
2
y
2
)
2
=
ZZ
R
4r dr dθ
(1 r
2
)
2
=
ZZ
R
sinh δ dδ dθ
where the last expression is written in polar co-ordinates using the hyperbolic distance δ. In this way
the measure of area also depends on the distance function.
Example (4.40, cont). The area of a hyperbolic circle with hyperbolic radius δ is
Z
2π
0
Z
δ
0
sinh δ dδ dθ = 2π(cosh δ 1) = π
δ
2
+
2
4!
δ
4
+
2
6!
δ
6
+ ···
> πδ
2
Again, this is larger than you’d expect in Euclidean geometry.
78
Angle Measure and the First Fundamental Form
If we repeat the distance translation (, page 77) for a second infinitesimal segment z z + dw, it can
be checked that the angle between the original segments is precisely that between the infinitesimal
vectors dz and dw. This is precisely the conformality observation in Theorem 4.32 and moreover
shows how the distance function determines the angle measure.
If you’ve studied differential geometry, then a more formal way to think about this is to use the
first fundamental form or metric: essentially the dot product of infinitesimal tangent vectors. For the
Poincar
´
e disk model, () says that this is
I =
4( dx
2
+ dy
2
)
(1 x
2
y
2
)
2
=
4( dr
2
+ r
2
dθ
2
)
(1 r
2
)
2
Since this is a scalar multiple of the standard Euclidean metric dx
2
+ dy
2
, angle measures are identi-
cal.
33
It also gels with the fact that arc-length is the integral
Z
q
I
z
( t), z
( t)
dt
Using this language, two of the major theorems of introductory differential geometry quickly put a
couple of remaining issues to bed.
Gauss’ Theorem Egregium The first fundamental form determines the Gaussian curvature K. In this
case K = 1 is constant and negative, as you should easily be able to verify if you’ve studied
differential geometry.
Gauss–Bonnet Theorem The angle-sum Σ
of a geodesic triangle in a space with Gaussian curva-
ture K satisfies
Σ
π =
ZZ
K
This establishes our earlier assertion that Area = π Σ
(Corollary 4.30).
33
Recall that the angle ψ between vectors u, v satisfies u ·v =
|
u
||
v
|
cos ψ. For infinitesimal vectors we use I = λ
2
(dx
2
+
dy
2
) instead of the dot product, where λ =
2
1r
2
. The resulting angle is the same as if we use the Euclidean metric
dx
2
+ dy
2
, since factors of λ
2
cancel on both sides.
79
5 Fractal Geometry
5.1 Natural Geometry, Self-similarity and Fractal Dimension
Classical geometry typically considers objects (lines, curves, spheres, etc.) which seem flatter and
less interesting as one zooms in: a differentiable curve at small scales looks like a line segment!
By contrast, real-world objects tend to exhibit greater detail at smaller scales. A seemingly spherical
orange is dimpled on closer inspection. Is its surface area that of a sphere, or is it greater due to the
dimples? What if we zoom in further? Under a microscope, the dimples are seen to have minute
cracks and fissures. With modern technology, we can see almost to the molecular level; what does
surface area even mean at such a scale?
The Length of a Coastline In 1967 Benoit Mandelbrot asked a related question in a now-famous
paper, How Long Is the Coast of Britain? Statistical Self-Similarity and Fractional Dimension. His essential
point was that the question has no simple answer:
34
Should one measure by walking along the mean
high tide line? But where is this? Do we ‘walk’ round every pebble? Round every grain of sand?
Every molecule? As one shrinks the scale, the measured length becomes absurdly large. We sketch
Mandelbrot’s approach.
35
Given a ruler of length R, measure how many N are required to trace round the coastline when
laid end-to-end.
Plotting log N against log(1/R) for several sizes of ruler seems to give a straight line!
log N log k + D log(1/R) = log(kR
D
) = N kR
D
The number D is Mandelbrot’s fractal dimension of the coastline.
Mandelbrot’s fractal dimension is purely empirical, though it does seem to capture something about
the ‘bumpiness’ of a coastline: the bumpier, the greater its fractal dimension. For mainland Britain
with its smooth east and rugged west coasts, D 1.25. Given its many fjords, Norway has a far
rougher coastline and a higher fractal dimension D 1.52.
Example 5.1. As a sanity check, consider a smooth circular ‘coastline.’
Approximate the circumference using N rulers of length R: clearly
R = 2 sin
π
N
As N , the small angle approximation for sine applies,
R
2π
N
= N 2πR
1
where the approximation improves as N . The fractal dimension
of a circle is therefore 1.
1
R
2π
N
34
The official answer from the Ordnance Survey (the UK government mapping office) is, ‘It depends.’ The all-knowing
CIA states 7723 miles, though offers no evidence as to why.
35
For more detail see the Fractal Foundation’s website. Mandelbrot coined the word fractal, though he didn’t invent the
concept from nothing. Rather he applied earlier ideas of Hausdorff, Minkowski and others, and observed how the natural
world contains many examples of fractal structures.
80
Our goal is to describe self-similar objects and thus create a new notion of dimension related to
Mandelbrot’s. To begin consideration of self-similarity, we first consider some of the standard objects
of pre-fractal geometry.
Segment A segment can be viewed as N copies of itself scaled by a factor
r =
1
N
.
Square A square comprises N copies of itself scaled by a factor r =
1
N
.
Cube A cube comprises N copies of itself scaled by a factor r =
1
3
N
.
In each case observe that N =
1
r
D
where D is the usual dimension of the
object (1, 2 or 3). Inspired by this, we make a loose definition.
Definition 5.2. A geometric figure is self-similar if it may be subdivided into N similar copies of
itself, each scaled by a magnification factor r < 1. The fractal dimension of such a figure is
D := log
1/r
N =
log N
log( 1/r)
=
log N
log r
Example 5.3. The botanical pictures below offer some evidence for non-integer fractal dimension
and that self-similarity is a natural phenomenon. The ‘tree’ comprises N = 3 copies of itself, each
scaled by a factor of r = 0.4. Its fractal dimension is D =
log 3
log 0.4
1.199.
The fern has N = 7 and r = 0.3 for a fractal dimension D =
log 7
log 0.3
1.616.
Tree fractal D 1.199 Fern fractal D 1.616
The pictures illustrate the interpretation of fractal dimension. Both objects seem to occupy more
space than mere lines, but neither has positive area. Moreover, the fern seems to occupy more space
than the tree. The ‘trunk’ and ‘branches’ in the first picture aren’t part of the fractal, and are drawn
only to give the picture a skeleton.
81
Example 5.4 (Cantors Middle-third Set). This famous example dates from the late 1800s.
36
Starting with the unit interval C
0
= [0, 1], define a se-
quence of sets (C
n
) where C
n+1
is obtained by deleting the
open ‘middle-third’ of each interval in C
n
; for instance
C
1
=
0,
1
3
2
3
, 1
Cantor’s set is essentially the limit of this sequence:
C :=
\
n=0
C
n
0
1
3
2
3
1
C
0
C
1
C
2
C
3
.
.
.
Cantor’s set has several strange properties. . .
Zero length If the length of a set is the sum of the lengths of its disjoint sub-intervals, then
length(C
n
) =
2
3
n
since we delete
1
3
of the remaining set at each step. It follows that
n N
0
, length(C)
2
3
n
= length(C) = 0
Otherwise said, C contains no subintervals.
Uncountable There exists a bijection between C and the original interval [0, 1]!
Self-similarity Abusing notation somewhat,
C
n+1
=
1
3
C
n
1
3
C
n
+
2
3
where we mean that C
n+1
consists of two copies of C
n
, each shrunk by a factor of
1
3
and one
shifted
2
3
to the right. The upshot is that the Cantor set itself satisfies
C =
1
3
C
1
3
C +
2
3
Being similar to two disjoint subsets of itself, its fractal dimension is D =
log 2
log 3
0.631. The
above image links to an animation showing how the full set may be doubled to produce itself.
The Cantor set has many generalizations. Look up the Sierpi
´
nski triangle (D =
log 3
log 2
1.585) and
carpet (Examples 5.8, D =
log 8
log 3
1.893), and the Menger sponge (D =
log 20
log 3
2.727).
36
Henry Smith discovered this set in 1874 while investigating integrability, in which context the ‘length’ of a set was later
formalized using measure theory. Cantor’s description in 1883 was more focused on topological properties. Self-similarity
was less of a concern at the time.
82
Example 5.5 (The Koch Curve and Snowflake). The Koch curve is another generalization of the
Cantor set, produced as the limit of a sequence of curves.
Let K
0
be a segment of length 1.
Replace the middle third of K
0
with the other two sides of
an equilateral triangle to create K
1
.
Replacing the middle third of each segment in K
1
as before
to create K
2
.
Repeat this process ad infinitum.
The curve is drawn, along with the Koch snowflake obtained by
arranging three copies around an equilateral triangle.
The relation to the Cantor set should be obvious in the construc-
tion. Indeed if K
0
= [0, 1], then the intersection of this with the
Koch curve is the Cantor set!
The Koch curve is self-similar in that it comprises N = 4 copies
of itself shrunk by a factor of r =
1
3
. Its fractal dimension is
therefore
log 4
log 3
1.2619.
We may also consider the curve’s length. Let s
n
be the number
of segments in K
n
, each having length t
n
. Also let
n
= t
n
s
n
be
the length of the curve K
n
. We easily see that
s
n
= 4
n
, t
n
=
1
3
n
=
n
=
4
3
n
from which the Koch curve is infinitely long!
Koch Curve
Koch Snowflake
Self-similarity
Exercises 5.1. 1. By removing a constant middle fraction of each interval, construct a fractal analo-
gous to the Cantor set but with dimension
1
2
.
2. Prove that the area inside the n
th
iteration of the construction of the Koch snowflake is
A
n
=
1 +
3
5
1
4
9
n
3
4
The area inside the complete snowflake is therefore
8
5
that of the original triangle.
3. Suppose r(t), t [0, 1] is a regular (smooth) curve in the plane.
(a) Use the arc-length formula L =
R
1
0
|
r
( t)
|
dt together with Riemann sums and the linear
approximation r(t + ϵ) r(t) + ϵr
( t) with ϵ =
1
N
to argue that
L
N1
k=0
r
k + 1
N
r
k
N
()
(b) Suppose the curve is parametrized such that each segment on the right side of () has the
same length R. Prove that L NR.
Any regular curve thus has fractal dimension 1 in the sense stated by Mandelbrot (pg. 80).
83
5.2 Contraction Mappings & Iterated Function Systems
Thus far we have only dealt with fractals where the whole consists of pieces scaled by the same
factor. In general we can mix up scaling factors. To do this it is helpful to borrow some language
from topology.
Definition 5.6. A contraction mapping is a function S on a subset of R
n
such that c [0, 1) with
|
S(x) S(y)
|
c
|
x y
|
A contraction mapping therefore moves points closer together. It should be clear that every con-
traction mapping is continuous. The main idea of this section is that fractals may be generated by
repeatedly applying contraction mappings to an initial shape. We have already seen a example:
Example (5.4, mk. II). Consider the following functions S
1
, S
2
: R R
S
1
(x) =
x
3
S
2
(x) =
x
3
+
2
3
These are certainly contraction mappings
x, y R,
|
S
1
(x) S
1
( y)
|
=
|
S
2
(x) S
2
( y)
|
=
1
3
|
x y
|
with scale factor c =
1
3
. More importantly, these functions define the Cantor set: at each stage of its
construction, we have
C
n+1
:= S
1
(C
n
) S
2
(C
n
)
As the limit of this process, the self-similarity of the Cantor set can be expressed in the same manner:
C = S
1
( C) S
2
( C).
Surprisingly, it barely seems to matter what initial set C
0
we choose. For example, we could start
with the singleton set C
0
= {0}, from which
C
1
= {0,
2
3
}, C
2
= {0,
2
9
,
2
3
,
8
9
}, C
3
= {0,
2
27
,
2
9
,
8
27
,
2
3
,
20
27
,
8
9
,
26
27
}, . . .
We draw the first few iterations below. In the second picture, we start with a very different initial set
C
0
= [0.2, 0.5] [0.6, 0.7]. Iterating this also appears to produce the Cantor set!
0
1
3
2
3
1
C
0
C
1
C
2
C
3
C
4
C
5
C
6
.
.
.
0
1
3
2
3
1
0
1
3
2
3
1
C
0
C
1
C
2
C
3
C
4
C
5
C
6
.
.
.
0
1
3
2
3
1
84
Iterated Function Systems
It certainly seems as if the Cantor set might be generated by the contraction maps S
1
, S
2
indepen-
dently of the initial data C
0
. The following result shows in what sense this is the case, though it relies
on some heavy lifting from topology. If you’ve done some analysis, then several of the concepts will
be familiar. We summarize the discussion without proof.
A subset of R
m
is compact if it is closed (contains its boundary points) and bounded (all points lie
within some ball centered at the origin).
The set of all compact subsets of R
m
is a metric space H. This means that the distance d(X, Y)
between two compact sets X, Y H may sensibly be defined, though it is a little tricky. . .
37
Since H is a metric space, we can discuss convergent sequences (K
n
) of compact sets
lim
n
K
n
= K lim
n
d(K
n
, K) = 0
It also makes sense to speak of Cauchy sequences in H. Moreover, H is complete in that every
Cauchy sequence (K
n
) H converges to some K H.
The Banach Fixed Point Theorem now applies.
If S : H H is a contraction mapping on a complete metric space H, then S has a
unique fixed point (some F H such that S(F) = F). Moreover, if F
0
H is any
initial value, then the sequence defined iteratively by F
k+1
:= S(F
k
) converges to F.
This powerful result has applications throughout mathematics.
Theorem 5.7. Let S
1
, . . . , S
n
be contraction mappings on R
m
with ratios c
1
, . . . , c
n
. Define
S : H H by S(D) =
n
[
i=1
S
i
(D)
1. S is a contraction mapping on H, with contraction ratio c = max{c
i
}.
2. S has a unique fixed set F H given by F = lim
k
S
k
(E) for any non-empty E H.
Part 1 is not difficult to prove if you’re willing to work with the definition of the Hausdorff metric
(try it if you’re comfortable with analysis!). Part 2 is Banach’s theorem.
The upshot is this: if we take any non-empty compact set E and repeatedly apply contraction map-
pings, the process will converge to a limit which is independent of E! We call the limit set F for fractal.
Such fractals are often called attractors: being limit-sets, they ‘attract’ data towards themselves.
37
This is the Hausdorff metric. Given Y H, and x R
n
, define d
Y
(x) = inf
yY
||
x y
||
to be the distance from x to the
‘nearest’ point of Y. Define d
X
(y) similarly. The Hausdorff distance between X and Y is then
d(X, Y) := max
(
sup
xX
d
Y
(x), sup
yY
d
X
(y)
)
Roughly speaking, find x X which is as far away d
Y
(x) as possible from anything in Y, and find y Y similarly; d(X, Y)
is the larger of these distances.
85
Examples 5.8. 1. (Cantor set Ex. 5.4) Theorem 5.7 shows that we may let C
0
be any closed bounded
subset of R. Repeatedly applying the contraction mappings S
1
and S
2
will always result in the
same set C.
A nice application is that one can easily find all sorts of interesting points in the Cantor set. For
instance, suppose x, y R are a pair such that y = S
1
(x) and x = S
2
( y): otherwise said
y =
1
3
x and x =
1
3
( y + 2)
Since E = {x, y} is a compact set satisfying E S(E), it follows that E lim S
k
(E) = C, from
which x, y both lie in the Cantor set! However, we can easily solve to see that (x, y) = (
3
4
,
1
4
).
This seems paradoxical:
1
4
does not lie at the end of any deleted interval (denominators of the
form 3
n
) but yet the Cantor set contains no intervals. How does
1
4
end up in there?!
2. (Koch curve, Ex. 5.5) Define four mappings S
i
: R
2
R
2
,
each with scale factor c =
1
3
.
Mapping Effect
S
1
(x, y) =
x
3
,
y
3
Scale
1
3
S
2
(x, y) =
1
6
x
3
6
y +
1
3
,
3
6
x +
1
6
y
Scale
1
3
, rotate 60°, translate
S
3
(x, y) =
1
6
x +
3
6
y +
1
2
,
3
6
x
1
6
y +
3
6
Scale
1
3
, rotate 60°, translate
S
4
(x, y) =
x
3
+
2
3
,
y
3
Scale
1
3
, translate
Applied to the Koch curve, the image of each map corresponds by color. The picture links to a
series of animated constructions of the curve starting with different initial sets E.
3. (Sierpi
´
nski carpet) Eight contraction mappings produce this fractal,
each reducing the whole by a (length-scale) factor of
1
3
.
As with the Koch curve, the image links to several alternative con-
structions using different initial starting sets.
4. (A Fractal Fern) This is built from three contraction mappings:
S
1
: Scale by
3
4
, rotate clockwise, and translate by (0,
1
4
)
S
2
: Scale by
1
4
, rotate 60° counter-clockwise, and translate by (0,
1
4
)
S
3
: Scale by
1
4
, rotate 60° clockwise, and translate by (0,
1
4
)
86
Fractal Dimension Revisited
Since Theorem 5.7 permits several different contraction factors, we need a new
approach to computing fractal dimension. We ask how many disks of a given
radius ϵ are required to cover a set. In the picture, the unit square requires four
disks of radius ε = 0.4. For smaller ε, we will plainly need more disks. . .
Definition 5.9. Let A be a compact subset of R
m
.
1. If ε > 0, the closed ε-ball centered at x A consists of the points at most a distance ε from x:
B
ϵ
(x) = {y R
m
: d(x, y) ε}
2. The minimal ε-covering number for A is
N(A, ε) = min
(
M : x
1
, . . . , x
M
A with A
M
[
n=1
B
ϵ
(x
n
)
)
3. Given a compact set A R
m
, its fractal dimension is the limit
D = lim
ε0
log N(A, ε)
log( 1/ε)
We don’t claim to prove that D must exist, though a simple example should at least convince you
that the definition is reasonable!
Example 5.10. Let A = [0, 1] be the interval of length 1. It is not hard to see that
ε
1
2
N(ε) = 1, and
1
4
ε <
1
2
N(ε) = 2
etc. More generally, N and ϵ are related via
1
2N
ϵ <
1
2(N 1)
The dimension of the line (1) may therefore be recovered via the squeeze theorem
D = lim
ϵ0
log N
log( 1/ε)
= 1
Thankfully an easier-to-use modification is available using boxes.
Theorem 5.11 (Box-counting). Let A be compact and cover R
m
by boxes of side length
1
2
n
. Let N
n
(A)
be the number of boxes intersecting A. Then
D = lim
n
log N
n
(A)
log 2
n
87
We finish with a formula satisfied by the dimension of an iterated function system (Theorem 5.7).
Theorem 5.12. Let {S
n
}
M
n=1
be an iterated function system with attractor (limiting fractal) F and
where each contraction S
n
has scale factor c
n
(0, 1). At each stage of the construction, suppose
portions of the fractal generated by each contraction map meet only at boundary points. Then the
fractal dimension is the unique D satisfying
M
n=1
c
D
n
= 1
Examples 5.13. 1. If all scale-factors are identical c
n
= r, we recover Definition 5.2,
Mr
D
= 1 = D =
log M
log r
=
log M
log( 1/r)
2. The fractal fern (Examples 5.8) is generated by three contraction maps with scale factors
3
4
,
1
4
,
1
4
.
Its dimension is the solution to the equation
3
4
D
+
1
4
D
+
1
4
D
= 1 = D 1.3267
3. Numerical approximation is usually required to solve for D, though sometimes an exact solu-
tion is possible. For instance, if c
1
= c
2
=
1
2
and c
3
= c
4
= c
5
=
1
4
, then
2
1
2
D
+ 3
1
4
D
= 1
Writing α =
1
2
D
yields the quadratic equation
2α + 3α
2
= 1 = α =
1
3
= D = log
2
3 1.584
Other methods of creating fractals
The contraction mapping approach is one of many ways
to create fractals. Two other famous examples are the logis-
tic map (related to numerical approximations to non-linear
differential equations) and the Mandelbrot set (pictured).
The Mandelbrot set arises from a construction in the com-
plex plane. For a given c C, we iterate the function
f
c
( z) = z
2
+ c
If f ( f ( f (··· f (c) ···))) remains bounded, no matter how
many times f is applied, then c lies in the Mandelbrot set.
Much better pictures and some trippy videos can be found
online. . .
1
i
i
88
Exercises 5.2. 1. Let S
1
(x) =
1
3
x and S
2
(x) =
1
3
x +
2
3
be the contraction mappings defining the
Cantor set and suppose x, y, z R satisfy
y = S
1
(x), z = S
2
( y), x = S
2
( z)
Show that x, y, z lie in the Cantor set, and find their values.
2. The construction of a Cantor-type set starts by removing the open intervals (0.1, 0.2) and
(0.6, 0.8) from the unit interval.
(a) Sketch the first three iterations of this fractal.
(b) This construction may be described using three contraction mappings; what are they?
(c) State an equation satisfied by the dimension D of the set and use a computer algebra
package to estimate its value.
3. A variation on the Koch curve is constructed using the following contraction mappings. Each
is built by first scaling the whole picture by a factor c, rotating the picture through an angle
counter-clockwise, and then translating the picture by adding a constant. The resulting fractal
is drawn.
map scale rotate translate (add (x, y) )
S
1
1
2
0 0
S
2
1
4
90° (
1
2
, 0)
S
3
1
4
0 (
1
2
,
1
4
)
S
4
1
4
90° (
3
4
,
1
4
)
S
5
1
4
0 (
3
4
, 0)
(a) Suppose you start with the straight line segment from (0, 0) to (1, 0). Draw the first two
iterations of the fractal’s construction.
(b) The dimension of the fractal is the unique solution D to the equation
1
2
D
+
1
4
D
+
1
4
D
+
1
4
D
+
1
4
D
= 1
By observing that
1
4
=
1
2
2
, convert this to a quadratic equation in the variable α :=
1
2
D
.
Hence compute the dimension of the fractal.
(c) The dimension computed in part (b) is larger than the dimension
log 4
log 3
of the Koch curve.
Explain what this means.
4. Verify the details of Example 5.10, including the computation of the limit.
5. In Theorem 5.12, prove that D exists and is unique.
(Hint: You’ll need the intermediate value theorem from calculus)
89