Cambridge Part 2567
Cambridge Part 2567
Mathematical Tripos
Sky Wilshaw
Part II
University of Cambridge
2020–2022
2
Contents
1 Algebraic Topology 7
3 Graph Theory 15
5 Galois Theory 25
3
4 CONTENTS
Introduction
This book contains notes for the maths courses at Cambridge University. Please note that while
efforts have been made to ensure completeness and correctness, no guarantees can be made; this
is simply a reasonably complete way of collating information about the courses.
This book can be downloaded in PDF form for free at https://thirdsgames.co.uk/maths.html,
and the source code (for the book itself and for the individual courses) can be accessed at
https://github.com/zeramorphic/cambridge-maths-notes.
You are given the right to download the PDF of the book (or its component parts) for private use.
You are permitted to download and modify the source code of the repository (the book and the
course notes it contains), but may not distribute these modifications (including object files such as
PDFs generated from these modifications) to others. However, you are permitted to make public
forks of the repository in order to create pull requests, but this does not grant you permission to
distribute object files created from these forked repositories. It must be clear when visiting it that
such a repository is a fork of https://github.com/zeramorphic/cambridge-maths-notes, and
must include a link to the original repository. (Forks created on GitHub satisfy this requirement,
as the title contains the words ‘forked from’ and then a link to the original repository.)
This project makes use of free software in accordance with their license terms, including the
cobra CLI interface generator
https://pkg.go.dev/mod/github.com/spf13/[email protected]. For more information, read the
licenses of each dependency.
5
6 CONTENTS
Course 1
Algebraic Topology
Contents
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.1.2 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 Homotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.2 Contractible spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
7
8 COURSE 1. ALGEBRAIC TOPOLOGY
1.1 Motivation
1.1.1 Motivation
Topological spaces are difficult to study on their own, and so we will assign algebraic invariants
to these spaces which allow us to reason more easily about these spaces. To a topological space
X, a ‘numerical invariant’ is a number g( X ) ∈ R ∪ {∞} such that X ≃ Y (where ≃ denotes
homeomorphism) implies g( X ) = g(Y ). An example of a numerical invariant is the number
of path-connected components of X. An algebraic invariant is a group G ( X ) assigned to a
topological space X such that X ≃ Y implies G ( X ) ≃ G (Y ), where here ≃ denotes isomorphism.
We will construct two kinds of such invariants: the fundamental group, and invariants related
to homology. The invariants we construct will behave nicely under maps: if f : X → Y is a
continuous map, we induce a homomorphism f ⋆ : G ( X ) → G (Y ). We will prove the following
model theorems.
• If Rn ≃ Rm , then n = m.
The above theorems are easy to prove in the case n = 1 by appealing to path-connectedness and
the intermediate value theorem. Our study allows us to prove similar things about these higher
dimensional cases, among other things.
1.1.2 Notation
• A space is a topological space.
– I = [0, 1] ⊂ R;
1.2 Homotopy
1.2.1 Definition
Proof. It suffices to show that the preimage of a closed set is closed. Let K ⊆ Y be closed. Then
−1
Ki = f −1 (K ) ∩ Ci = f C (K ) is a closed set in Ci and so is closed in X because Ci is closed.
i
Since K = K1 ∪ K2 , K is also closed in X.
Proof. If Y is contractible, and p, q ∈ Y, then c{•},p ∼ c{•},q via H : {•} × I → Y. Then define
γ(t) = H (•, t) is a path from p to q in Y.
Example. If X ≃ Y, X and Y are homotopy equivalent. Note that the definition of homotopy
equivalence is simply the definition of homeomorphism, except that the requirement that f ◦ g
and g ◦ f be equal to the identity is relaxed into simply being homotopic to the identity.
Contents
2.1 Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.2 Rings and algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
11
12 COURSE 2. PROBABILITY AND MEASURE
2.1 Measures
2.1.1 Definitions
Remark. Since n An = ( n Acn )c , any σ-algebra E is closed under countable intersections as well
T S
as under countable unions. Note that B \ A = B ∩ Ac ∈ E , so σ-algebras are closed under set
difference.
Definition. A set E with a σ-algebra E is called a measurable space. The elements of E are
called measurable sets.
Definition. A measure µ is a set function µ : E → [0, ∞], suchSthat µ(∅) = 0, and for a
sequence ( An )n∈N such that the An are disjoint, we have µ( n∈N An ) = ∑n∈N µ( An ).
This is the countable additivity property of the measure.
∑ x∈ A µ({ x }). Hence, measures are uniquely defined by the measure of each singleton. This
corresponds to the notion of a probability mass function.
Rings are easier to manage than σ-algebras because there are only finitary operators.
2.1. MEASURES 13
Remark. Rings are closed under symmetric difference A △ B = ( B \ A) ∪ ( A \ B), and are closed
under intersections A ∩ B = A ∪ B \ A △ B. Algebras are rings, because B \ A = B ∩ Ac =
( Bc ∪ A)c . Not all rings are algebras, because rings do not need to include the entire space.
is a σ-algebra (or aS
ring, if the union is finite). Then there exist Bn ∈ E that are disjoint
such that n An = n Bn .
S
Proof. Define A
en = S j≤n A j , then Bn+1 = A en−1 .
en \ A
M = { A ⊆ E | ∀ B ⊆ E. µ⋆ ( A) = µ⋆ ( A ∩ B) + µ⋆ ( A ∩ Bc )}
Graph Theory
Contents
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.1.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
15
16 COURSE 3. GRAPH THEORY
3.1 Introduction
3.1.1 Definitions
We use the notation [n] for {1, . . . , n}. For a set X and k ∈ N, we define X (k) = {Y ⊆ X | |Y | = k }.
Definition. A graph is a pair (V, E), where V is a set of vertices and E is a set of edges where
E ⊆ V (2) . We use the notation V ( G ) to denote the set of vertices and E( G ) to denote the
set of edges, where G = (V, E) is a graph. We define | G | = |V ( G )|, and e( G ) = | E( G )|.
Example. The complete graph on n vertices, denoted Kn , is the graph with V = [n] and E = V (2) .
Note that we sometimes use juxtaposition of names of vertices to denote an edge between them,
so 13 represents the edge {1, 3}.
Remark. Edges are undirected. There are no edges from a vertex to itself. Edges between vertices
are unique if they exist. Most of the graphs covered in this course are finite.
Example. The empty graph on n vertices, denoted K n , is the graph with vertex set V = [n] and
E = ∅.
Example. The path of length n, denoted Pn , is the graph with V = [n + 1] and E = {{1, 2}, . . . , {n, n + 1}}.
Example. The cycle of length n, denoted Cn , is the graph with V = [n] and E = {{1, 2}, . . . , {n − 1, n}, {n, 1}}.
NG ( x ) = {y ∈ V ( G ) | { x, y} ∈ E( G )}
If y is a neighbour of x, we write x ∼ y.
Proposition. If W is a x–y walk for x ̸= y, W contains a x–y path, where ‘contains’ denotes
a subsequence.
Proof. Let W ′ be the minimal x–y walk in W. This is a path, because if there were a repeated
vertex, we could find a shorter path by eliminating the detour.
18 COURSE 3. GRAPH THEORY
Course 4
Contents
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.1 Exposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.2 Basic definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.3 Revisiting Numbers and Sets . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.4 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Rewrite systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2.2 Relation to languages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.2.3 Grammars . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
19
20 COURSE 4. AUTOMATA AND FORMAL LANGUAGES
4.1 Introduction
4.1.1 Exposition
Computation, or computability, is central to modern mathematics. However, we very rarely
think about the precise definition of what it means for something to be ‘computable’. There is
an important difference between existence and algorithmic access to a witness. Contrast the
statements ‘every polynomial of order n has a root’, and ‘there is an algorithm that, given a
polynomial of order n, we can find a root’. In many cases, there is an existence proof but no
algorithm to construct the relevant object.
In 1900, Hilbert gave a talk in Paris known as Mathematical Problems, in which he described a
list of 100 problems to be worked on in the coming 100 years. One of these problems, the tenth,
relates to an algorithm to determine whether solutions of Diophantine equations, those in Z[ X ],
exist. In 1928, Ackermann wrote the book Grundzüge der theoretischen Logik, in which he described
the famous Entscheidungsproblem: given a formula φ, determine whether φ is a tautology (true
regardless of how the variables are interpreted).
In both cases, Hilbert expected that solutions to these questions exist. Positive solutions to such
problems do not require a definition of words like ‘algorithm’ or ‘procedure’, because we can
agree on what an algorithm is when we see an example. However, to disprove such statements, we
need to rigorously define what an algorithm is, in order to rule all possible algorithms out.
Theorem (Cantor’s theorem). Let X be infinite. Then its power set P ( X ) is uncountable.
Proof. A simple diagonalisation argument shows there is no surjection from the naturals to the
power set P ( X ).
Proof. We construct a surjection from X ⋆ to Fin( X ); then by composition with the surjection
obtained in the first proposition we construct a surjection N → Fin( X ). Consider the forgetful
function f : X ⋆ → Fin( X ), mapping ( x1 , . . . , xn ) to { x1 , . . . , xn }. Since X is countable, π : N → X
is surjective, hence for x ∈ X, π −1 ( x ) ⊆ N is a nonempty set of naturals. Therefore, let n x be the
least element of π −1 ( x ). Then, given F ∈ Fin( X ), consider the set {n x | x ∈ F }, order it in the
usual way, and represent this as a sequence. This is a sequence of naturals with | F | elements, and
its π-image is exactly F.
4.1.4 Notation
We will use the following notational conventions.
• The natural numbers N are defined as {0, 1, 2, . . . }.
• We use the standard set-theoretic construction of naturals as Von Neumann ordinals, n =
{0, 1, . . . , n − 1}. Therefore, a natural is the set of all lower naturals.
• X n is the set of sequence of X-strings of length n, defined as X n = n → X, treating n as a
set as above.
• We write |α| = domain(α) for the length of a sequence.
• X 0 = 0 → X is a type with only one element ε, which is the empty sequence.
• We can write X ⋆ = Xn .
S
n ∈N
Definition. Let Ω be a finite set of symbols, and let Ω⋆ be the set of Ω-strings. We call
elements of Ω⋆ × Ω⋆ rewrite rules or production rules. Such elements (α, β) are written
α → β.
Definition. A pair R = (Ω, P) is called a rewrite system if P is a finite set of rewrite rules.
R
Definition. If R = (Ω, P) is a rewrite system, and σ, τ ∈ Ω⋆ , we write σ −
→1 τ, pronounced
‘σ is rewritten to τ in one step’ or ‘R producees τ from σ in one step’, if there exist
α, β, γ, δ ∈ Ω⋆ such that σ = αγβ, τ = αδβ, and γ → δ ∈ P.
R R R
The relation −
→ is the reflexive and transitive closure of −
→1 . The sequence σ0 −
→1
R R
n1 −
σ →1 σnois called a R-derivation of length n of σn from σ0 . We write D( R, σ) =
→1 . . . −
R
τ ∈ Ω⋆ | σ −→ τ for the set of strings that can be rewritten, produced, or derived from
σ.
There is a difference between a sentence being grammatical and being meaningful. One notable
example is the grammatically correct ‘colourless green ideas sleep furiously’ that does not have
meaning, to contrast with ‘furiously sleep ideas green colourless’ which is neither grammatically
correct or meaningful. We can use grammar to distinguish these two sentences, but we cannot
distinguish algebraically whether a sentence has meaning.
Example. Consider the following generative grammar of rewrite rules for English.
S → NP VP
NP → Adj NP
NP → Noun
VP → Verb
VP → Verb Adv
This rewrite system allows us to derive the sentence ‘colourless green ideas sleep furiously’ from
S.
4.2.3 Grammars
Definition. Let Σ be an alphabet of letters or terminal symbols, and let V be a set of variables
or nonterminal symbols, such that Σ, V are nonempty and disjoint. Let Ω = Σ ∪ V. a, b, c, . . .
refer to letters and A, B, C, . . . refer to variables. Elements of W = Σ⋆ ⊆ Ω⋆ are called
words. u, v, w, . . . refer to words. We denote W+ = Σ⋆ \ {ε} for the set of nonempty words.
A subset of W is called a language.
Note that there are uncountably many languages over any nonempty alphabet.
G R
D( G, σ) = D( R, σ); σ−
→(1) τ ⇐⇒ σ −
→ (1) τ
L( G ) = D( G, S) ∩ W
• P1 = {S → aSa, S → a}
• P2 = {S → Saa, S → a}
• P3 = {S → aaS, S → aaSaa, S → a}
This notion is called equivalence of grammars.
Course 5
Galois Theory
Contents
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5.1.2 Solving quadratics, cubics and quartics . . . . . . . . . . . . . . . . . . 26
5.2 Polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2.2 Symmetric polynomials . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
25
26 COURSE 5. GALOIS THEORY
5.1 Introduction
5.1.1 Introduction
Galois theory concerns itself with solving polynomial equations of higher degree, and discussing
how the symmetries of these polynomials relate to their solubility. The modern interpretation of
Galois theory is more interested in the fields that particular polynomials generate, rather than
their particular solutions; this naturally extends to studying symmetries of fields.
Cubics were solved much later, in the early 16th century, by the Italian mathematician del Ferro.
Consider the cubic X 3 + aX 2 + bX + c, written as ( X − x1 )( X − x2 )( X − x3 ). Multiplying, we
find
x1 + x2 + x3 = − a; x1 x2 + x2 x3 + x3 x1 = b; x1 x2 x3 = − c
2πi
where ω = e 3 . The u, v are known as Lagrange resolvents. Applying a cyclic permutation
to x1 , x2 , x3 in u or v, we find u 7→ ωu and v 7→ ωv. Hence, the cubes of u and v are invariant
under cyclic permutations of x1 , x2 , x3 . Under a permutation x2 7→ x3 , x3 7→ x2 , u and v swap.
Hence, u3 + v3 and u3 v3 are invariant under all permutations of roots. A general fact that we will
prove later is that such invariant expressions can be written in terms of the coefficients of the
polynomial. In this case, we have
u3 + v3 = −27c; u3 v3 = −27b2
Now, u3 and v3 are the roots of the quadratic Y 2 + 27cY − 27b2 . This then provides a formula for
the root x1 . This process is known as Cardano’s formula.
5.2 Polynomials
5.2.1 Definitions
In this course, ring means a commutative nonzero ring. If R is a ring, R[ X ] denotes the ring of
polynomials with elements ∑in=0 ai X i , and the usual operations of addition and multiplication. A
polynomial f ∈ R[ X ] can be interpreted as a function f : R → R, given by x 7→ ∑in=0 ai xi . It is,
however, important to distinguish the polynomial and its associated function; the polynomial
is not in general uniquely determined by the function. For example, let R = Z⧸pZ, so for all
a ∈ R, we have a p = a, and hence X p and X are different polynomials yet represent the same
function.
Recall from Groups, Rings and Modules that if R = K is a field, K [ X ] is a Euclidean domain
(and hence is a unique factorisation domain, a Noetherian ring, a principal ideal domain, and an
integral domain). Hence, there is a division algorithm: for polynomials f , g ∈ K [ X ], there exists
a unique q, r ∈ K [ X ] such that f = gq + r and deg r < deg g, where we denote deg 0 = −∞. If
g = X − a is linear, f = ( X − a)q + r where r = f ( a) ∈ K; this is the familiar remainder theorem.
Note that every polynomial f ∈ K [ X ] is a product of irreducible polynomials since K [ X ] is a
unique factorisation domain, and there are greatest common divisors which can be computed
using Euclid’s algorithm in the usual way.
Proof. If f has no roots, the proof is complete. If f has a root a, consider f = ( X − a)q + f (0) =
( X − a)q. For a root b of f , either b = a or q(b) = 0. By induction, q has at most deg q roots, since
deg q < deg f . Then deg q + 1 ≤ deg f as required.
Note that constant polynomials are symmetric, and the property of symmetry is closed un-
der addition and multiplication. Hence, the set of symmetric polynomials is a subring of
R [ X1 , . . . , X n ] .
Proposition. Let f σ ( X ) = f ( Xσ(1) , . . . , Xσ(n) ). This gives an action (on the right) of the
group Sn on R[ X1 , . . . , Xn ]. A polynomial f ∈ R[ X1 , . . . , Xn ] is symmetric if and only if f
is fixed under the action of Sn ; in other words, f σ = f for all σ ∈ Sn .
28 COURSE 5. GALOIS THEORY
s r ( X1 , . . . , X n ) = ∑ Xi 1 Xi 2 · · · Xi r
i1 <···<ir
For instance,
s 2 ( X1 , X2 , X3 ) = X1 X2 + X1 X3 + X2 X3
It is clear that these are symmetric polynomials.
I
Definition. A monomial is an expression of the form X I = X11 · · · XnIn for I ∈ Nn . The
(total) degree of a monomial is ∑in=1 Ii . A term is a scalar multiple of a monomial. A
polynomial is uniquely characterised by a sum of terms. The total degree of a polynomial
is the maximum total degree of its terms.
Monomials are equipped with a lexicographic ordering, where we say monomials X I > X J
if either I1 > J1 or I1 = J1 and for some r ∈ {1, . . . , n − 1} we have I1 = J1 , . . . , Ir =
Jr , Ir+1 > Jr+1 . This is a total order.
Proof. Let d be the total degree of a symmetric polynomial f . Let X I be the largest (in lexicographic
order) monomial which occurs in f with coefficient c. Since f is symmetric, any permutation of
the Xi yields another monomial that occurs in f . Hence, I1 ≥ I2 ≥ · · · ≥ In , because otherwise the
rearranged monomial that satisfies this will be a strictly larger monomial in f . We can therefore
write
I −I
X I = X11 2 ( X1 X2 ) I2 − I3 · · · ( X1 . . . Xn ) In
Consider
I − I2 I2 − I3
g = s11 s2 · · · snIn
By construction, the largest monomial in g is X I . Since g is symmetric, cg is symmetric. By
induction, we may assume f − cg is expressible as a sum of symmetric polynomials as it has total
degree no larger than d, its leading monomial term is smaller than X I , and there are only finitely
many monomials of degree at most d. Hence f is also expressible as a sum of polynomials as
required.
5.2. POLYNOMIALS 29
Now we prove uniqueness by induction on n. Let G ∈ R[Y1 , . . . , Yn ] such that G (s1,n , . . . , sn,n ) = 0.
We want to show that G is the zero polynomial. If n = 1, the result is trivial as s1,1 = X1 . If
G = Ynk H with Yn not dividing H, then skn,n H (s1,n , . . . , sn,n ) = 0. Since sn,n = X1 . . . Xn , it is
not a zero divisor in R[ X1 , . . . , Xn ]. Hence H (s1,n , . . . , sn,n ) = 0. Without loss of generality, we
can assume that G is not divisible by Yn . Now, replacing Xn with zero, sk,n is mapped to sk,n−1
for k ̸= n, and sn,n is mapped to zero. Hence, G (s1,n−1 , . . . , sn−1,n−1 , 0) = 0. By induction,
G (Y1 , . . . , Yn−1 , 0) = 0. Hence Yn | G, contradicting our assumption.
= −3 ∑ Xi X j X k
i< j<k
= −3s3
Hence f = s1 s2 − 3s3 .
Proof. It suffices to consider R = Z (or, for example, R = R). Consider the generating function
n n
F(T ) = ∏(1 − Xi T ) = ∑ (−1)r sr Tr
i =1 r =0
d
dT ( f g ) f ′ g + f g′ f′ g′
= = +
fg fg f g
30 COURSE 5. GALOIS THEORY
F′ (T ) d
dT ∏in=1 (1 − Xi T )
=
F(T ) ∏in=1 (1 − Xi T )
n d
dT (1 − Xi T )
= ∑ 1 − Xi T
i =1
n
Xi
=−∑
i =1
1 − Xi T
−1 n ∞ r r
T i∑ ∑ Xi T
=
=1 r =1
−1 ∞
T r∑
= pr T r
=1
Hence,
− TF ′ ( T ) = s1 T − 2s2 T 2 + · · · + (−1)n−1 nsn T n
but also
∞
− TF ′ ( T ) = F ( T ) ∑ pr Tr = (s0 − s1 T + · · · + (−1)n sn T n ) p1 T + p2 T 2 + . . .
r =1
Equating the coefficients of powers of T, we find the identity as required by the theorem.
D ( X1 , . . . , X n ) = ∆ ( X1 , . . . , X n ) 2
where
∆ ( X1 , . . . , X n ) = ∏ ( Xi − X j )
i< j