Topics in Number Theory - Siksek
Topics in Number Theory - Siksek
Topics in Number Theory - Siksek
Samir Siksek
Samir Siksek, Mathematics Institute, University of War-
wick, Coventry, CV4 7AL, United Kingdom
E-mail address: [email protected]
Contents
6. p-adic Integers 49
7. Hensel’s Lemma Revisited 51
8. The Hasse Principle 53
Chapter 5. Geometry of Numbers 57
1. The Two Squares Theorem 57
2. Areas of Ellipses and Volumes of Ellipsoids 58
3. The Four Squares Theorem 60
4. Proof of Minkowski’s Theorem 62
Chapter 6. Irrationality and Transcendence 65
1. Irrationality: First Steps 65
2. The irrationality of e 66
3. What about Transcendental Numbers? 67
Appendix X. Last Year’s Exam 71
Appendix Y. Mathematical Pornography 75
1. An Integral Equation 76
5
Do you subscribe to the illustrious Warwick tradition
of setting the same exam every year? I am not going to
answer that question, except to point out that this is only the
second time the module is being offered, so there’s no way for
you to guess what I’m going to do, and you’ll just have to work
hard for the exam.
The exam is tomorrow/next week/within six months.
I’m running around like a headless chicken and stressing
all my friends because I can’t do a homework question.
Can I knock on your door and ask you about it? Don’t
worry, I’ve already branched out into agony-aunting. Yes come
and ask, and I promise not to set the dogs on you.
IS THIS IT? I’ve spent two whole quid from my beer
money on your notes and they’re only 70 odd pages. Do
you call that value for money? I’m gutted to see you upset.
I was just about to offer you your two pounds back but I’ve had
a better idea. I’ll go through many sleepless nights to write more
notes and make them available via mathstuff. Do you forgive
me now?
After this year is over, I’m going to devote my life to
drunkeness and antisocial behaviour. This year however
is my last year in mathematics and I want to enjoy it to
the full. Can you pleeeeeease set us lots
of homework? We must be careful. If I give you too
much homework then you’ll suffer severe withdrawl symptoms
once the term is over, and there’s no telling what you might
do to yourself. I simply can’t have that on my conscience. I’ll
therefore limit the homework to one sheet per week. It cuts me
deep to be so hard on you, but sometimes you have to be tough
to be kind.
CHAPTER -1
FAQ
Why is this FAQ upside down? This is to increase the prob-
ability that you will notice it and read it, but you should already
know this if you took Vectors and Matrices with me.
I have/haven’t done MA246 Number Theory. Am I al-
lowed to take this module? Yes. This module has nothing
to do with MA246 Number Theory.
How is this module related to Algebra II? In my humble
opinion, this module should have been a prerequiste to Algebra
II. On the other hand, Algebra II is not needed to follow this
module.
I got massacred in second year algebra. Is it total suicide
for me to take your module? No. This module relies on
common sense, not algebra.
How is this course assessed? 15% for a yet undetermined
number of homework assignments and 85% for the final exam.
Are past exam papers available? This course has been of-
fered only once before. Last year’s paper is at the end of these
notes.
Can we have solutions to last year’s paper? I want you
to answer last year’s paper on your own; it will be good prac-
tice. However, if you email me two weeks before the final exam
solemnly swearing that you have done the paper and only want
to check your answers I’ll be happy to oblige you.
Are we required to know the proofs taken during the
lectures or found in the lecture notes? Yes, theorems, def-
initions, proofs and homework questions. I love bookwork.
-1. FAQ 6
CHAPTER 0
Prologue
1. What’s This?
These are my lecture notes for MA3H1 Topics in Number The-
ory, with the usual Siksek trademarks. Thanks go to Jenny Cooley,
Samantha Pilgram and Vandita (Ditz) Patel for corrections. Please
send comments, misprints and corrections to [email protected].
2. The Queen of Mathematics
Gauss wrote that “mathematics is the queen of sciences and number
theory is the queen of mathematics”. In this module we hope to cover
some fascinating but fairly elementary aspects of the subject, to ensure
maximal enjoyment with minimal prerequisites. Topics covered
should include:
(1) A review of the number theory you met in the first year Foun-
dations module (primes, unique factorisation, greatest com-
mon divisors, modular arithmetic, Chinese Remainder Theo-
rem).
(2) Structure of Z/mZ and Um .
(3) p-adic numbers.
(4) Geometry of Numbers.
(5) Diophantine equations.
(6) The Hasse Principle for ternary quadratic forms.
(7) Counterexamples to the Hasse Principle.
(8) Irrationality and transcendence.
7
CHAPTER 1
Review
You’ve spent the last two or three years thinking about rings, topo-
logical spaces, manifolds, and so on. This chapter reminds you of the
heros of your mathematical childhood: the integers. We review some
of their properties which you have seen before, but perhaps not for a
long time.
1. Divisibility
Definition. Let a, b be integers. We say that a divides b and write
a | b if there exists an integer c such that b = ac.
The following lemma gives easy properties of divisibility; all have
one-line proofs from the definition.
Lemma 1.1. (Easy Properties of Divisibility) For all integers a,
b, c and k:
(1) a | 0;
(2) if a | b then a | kb;
(3) if a | b and a | c then a | (b ± c);
(4) if a | b and b | c then a | c;
(5) if a | b and b | a then a = ±b.
(6) if a | b and b 6= 0 then |a| ≤ |b|;
(7) (±1) | a for all integers a;
(8) if a | (±1) then a = ±1.
Example 1.1. Show that 42 | (7n − 7) for all positive integers n.
Answer. This is easy to do using congruences (have a go). But let us
try to do it from the definition of divisibility using induction on n. It is
obvious for n = 1. Suppose it is true for n = k. That is, suppose that
42 | (7k − 7). In other words, 7k − 7 = 42c for some integer c. Then
7k+1 − 72 = 7 × 42c
so
7k+1 − 7 = 7 × 42c + 42 = 42(7c + 1).
As c is an integer, 7c + 1 is an integer, so 42 | (7k+1 − 7).
9
10 1. REVIEW
2. Ideals
Definition. An ideal in Z is a subset I satisfying the following three
properties:
(i) 0 ∈ I,
(ii) if a, b ∈ I then a + b ∈ I,
(iii) if a ∈ I and r ∈ Z then ra ∈ I.
The principal ideal of Z generated by a is the subset
(a) = aZ = {ka | k ∈ Z}.
In other words, (a) is the set of multiples of a.
It is very easy to check that every principal ideal is an ideal 1.
1Ifyou have done Algebra II then you will know that the converse is not true
in every ring, but is true for Z.
3. GREATEST COMMON DIVISORS 11
d = u1 a1 + · · · + un an ,
with ui ∈ Z. So
d = (u1 k1 + · · · + un kn )c.
Hence c | d. This proves (ii) and completes the proof of Theorem 1.4.
4. Euler’s Lemma
The fact that the gcd can expressed as a linear combination is used
again and again. For example, in the proof of the following crucial
lemma.
Lemma 1.5. (Euler’s Lemma) If u | vw and gcd(u, v) = 1 then u | w.
c | a and c | b ⇐⇒ c | b and c | r.
Lemma 1.6 is the basis for the Euclidean Algorithm for computing
the GCD, which you did in Foundations. Here is an example.
6. PRIMES AND IRREDUCIBLES 13
Example 5.1. To find the greatest common divisor of 1890 and 909
using the Euclidean Algorithm you would write
1890 = 2 × 909 + 72,
909 = 12 × 72 + 45,
72 = 1 × 45 + 27,
45 = 1 × 27 + 18,
27 = 1 × 18 + 9,
18 = 2 × 9 + 0,
therefore
gcd(1890, 909) = gcd(909, 72) = gcd(72, 45) = gcd(45, 27)
= gcd(27, 18) = gcd(18, 9) = gcd(9, 0) = 9.
You also know, or should know, how to use the above to express the
GCD, in this case 9, as a linear combination of 1890 and 909:
9 = 27 − 18
= 27 − (45 − 27) = −45 + 2 × 27
= −45 + 2(72 − 45) = 2 × 72 − 3 × 45
= 2 × 72 − 3(909 − 12 × 72) = −3 × 909 + 38 × 72
= −3 × 909 + 38(1890 − 2 × 909) = 38 × 1890 − 79 × 909.
Proof. Suppose that there are finitely many and let them be
p1 , . . . , pn . Let N = p1 p2 · · · pn + 1. Then N ≥ 2 and so by the
Fundamental Theorem of Arithmetic must have a prime divisor. This
must be one of p1 , . . . , pn ; say it’s pi . Then pi | N and pi | p1 p1 · · · pn .
Hence pi divides N − p1 p1 · · · pn = 1 giving a contradiction.
This proof is a model for many other proofs. For example, we’ll
show later that there are infinitely many primes p ≡ 1 (mod 4), p ≡ 3
(mod 4), p ≡ 1 (mod 3) etc.
7. Coprimality
Definition. We say that integers m1 , m2 , . . . , mn are coprime if
gcd(m1 , m2 , . . . , mn ) = 1.
We say that integers m1 , . . . , mn are pairwise coprime if gcd(mi , mj ) =
1 whenever i 6= j.
Lemma 1.11. Let m1 , . . . , mn be pairwise coprime
Q integers and sup-
pose mi | x for all i. Then M | x where M = mi .
8. ordp
Let p be a prime, and let n be a non-zero integer. We define ordp (n)
by the property
e = ordp (n) if and only if pe | n and pe+1 - n.
In a sense, ordp (n) measures how divisible n is by powers of p. We
define ordp (0) = ∞.
Example 8.1. If n = 23 × 32 × 7, then ord2 (n) = 3, ord3 (n) = 2,
ord7 (n) = 1 and ordp (n) = 0 for all primes p 6= 2, 3, 7.
16 1. REVIEW
where p does not divide a, b, c, d. Here u = ordp (α) and v = ordp (β).
Without loss of generality, we suppose that u ≤ v. Then
a c ad + pv−u bc
α + β = pu + pv−u = pu .
b d bd
Note that p - bd. However, we don’t know if the integer ad + pv−u bc is
divisible by p, so let’s write ad + pv−u bc = pw e, where p - e and w ≥ 0,
and write f = bd. Hence
e
α + β = pu+w
f
and so
ordp (α + β) = u + w ≥ u = min(u, v) = min(ordp (α), ordp (β)).
To complete the proof, suppose that ordp (α) 6= ordp (β), or in other
words, u 6= v. Since we are assuming u ≤ v we have u < v and so
v − u > 0. Now if p | (ad + pv−u bc) then p | ad which contradicts p - a,
d. Hence p - (ad + pv−u bc) which says that w = 0. We obtain the
desired equality
ordp (α + β) = u + w = u = min(u, v) = min(ordp (α), ordp (β)).
9. Congruences
We are still revising the material you have met in the first year
Foundations module.
Definition. Let a, b and m be integers with m positive. We say a
is congruent to b modulo m and write a ≡ b (mod m) if and only if
m | (a − b).
Lemma 1.16. Congruence modulo a fixed positive integer m is an
equivalence relation:
• Reflexive: a ≡ a (mod m) for all integers a;
• Symmetric: if a ≡ b (mod m) then b ≡ a (mod m);
• Transitive: if a ≡ b (mod m) and b ≡ c (mod m) then a ≡ c
(mod m).
The equivalence classes are represented by 0, 1, . . . , m − 1. In other
words, every integer is congruent to precisely one of 0, 1, . . . , m − 1
modulo m.
Lemma 1.17. (a) If a ≡ b (mod m) and c ≡ d (mod m) then
a + c ≡ b + d (mod m) and ac ≡ bd (mod m).
(b) If a ≡ b (mod m) and d | m then a ≡ b (mod d).
18 1. REVIEW
Example 9.2. Let us find the inverse of 502 modulo 2001. One way
of doing this is to try all the numbers b = 0, 1, · · · , 2000 and see which
one satisfies 502b ≡ 1 (mod 2001). Using Euclid’s algorithm is much
faster!
2001 = 3 × 502 + 495,
502 = 1 × 495 + 7,
495 = 70 × 7 + 5,
7 = 1 × 5 + 2,
5 = 2 × 2 + 1.
Therefore gcd(502, 2001) = 1. Moreover,
1=5−2×2
= 5 − 2(7 − 5) = −2 × 7 + 3 × 5
= −2 × 7 + 3(495 − 70 × 7) = 3 × 495 − 212 × 7
= 3 × 495 − 212(502 − 495) = −212 × 502 + 215 × 495
= −212 × 502 + 215(2001 − 3 × 502) = 215 × 2001 − 857 × 502.
Reducing 215 × 2001 − 857 × 502 = 1 modulo 2001 we obtain −857 ×
502 ≡ 1 (mod 2001), so the inverse of 502 is −857 ≡ 2001−857 ≡ 1144
(mod 2001).
Proof. Suppose first that {a1 , . . . , aϕ(m) } and {b1 , . . . , bϕ(m) } are
reduced residue systems modulo m. By part (b) of Lemma 1.20 we
have
ϕ(m) ϕ(m)
Y Y
ai ≡ bi (mod m).
i=1 i=1
Now let {a1 , . . . , aϕ(m) } be any reduced residue system and observe that
{ca1 , ca2 , . . . , caϕ(m) } is also a reduced residue system by Lemma 1.22.
Hence
ϕ(m) ϕ(m)
Y Y
ai ≡ cai (mod m).
i=1 i=1
ϕ(m)
Q
We may rewrite this as A ≡ c A (mod m) where A = ai . Clearly
gcd(A, m) = 1, and by part (d) of Lemma 1.17 we obtain cϕ(m) ≡ 1
(mod m).
Proof. Let p be a prime. Note that the only integer in the set
{0, 1, . . . , p − 1} that is not coprime with p is 0. Hence, by definition
of ϕ, we have ϕ(p) = p − 1. Now (i) follows from Euler’s Theorem.
Let us prove (ii). If p - a then (ii) follows from (i) on multiplying both
sides by a. If p | a then (ii) is obvious since both sides are congruent
to 0 modulo p.
1. Euler’s ϕ Revisited
With the help of the Chinese Remainder Theorem we will derive a
convenient formula for ϕ. For this we have to revisit reduced residue
systems.
Lemma 2.1. If gcd(m1 , m2 ) = 1 then ϕ(m1 m2 ) = ϕ(m1 )ϕ(m2 ).
Proof. For a positive integer m define
U (m) = {a | 0 ≤ a ≤ m − 1 and gcd(a, m) = 1}.
Note that ϕ(m) = #U (m). Now let m1 , m2 be coprime and write
M = m1 m2 . We will shortly define a bijection
f : U (m1 ) × U (m2 ) → U (M ).
You know if two finite sets are related by a bijection then they have
the same number of elements. Assuming the existence of the bijection
f we obtain
2. Orders Modulo m
Definition. Let gcd(a, m) = 1. We define the order of a modulo m to
be the least positive integer d such that ad ≡ 1 (mod m).
Lemma 2.3. Suppose
au ≡ av ≡ 1 (mod m).
and let w = gcd(u, v). Then aw ≡ 1 (mod m).
Proof. By Euclid’s Algorithm, there are r, s such that w = ru +
sv. So that
aw = (au )r (av )s ≡ 1 (mod m).
Theorem 2.4. Let gcd(a, m) = 1, and let d be the order of a modulo
m.
(i) If ae ≡ 1 (mod m) then d | e.
(ii) d | ϕ(m). In particular, if m = p is prime then d | (p − 1).
3. PRIMITIVE ROOTS 25
3. Primitive Roots
Lemma 2.6. Let p be a prime and X an indeterminate. Then
X p−1 − 1 ≡ (X − 1)(X − 2) · · · (X − (p − 1)) (mod p).
Proof. By Fermat’s Little Theorem, ap−1 ≡ 1 (mod p) for a =
1, 2, . . . , p−1. So X p−1 −1 must have a = 1, 2, · · · , p−1 as roots modulo
p. Thus, modulo p, the polynomial (X − 1)(X − 2) · · · (X − (p − 1)) is
a factor of X p−1 − 1. But both are monic of degree p − 1, so they must
be the same modulo p.
Lemma 2.7. Let p be a prime. If n | (p − 1) then xn ≡ 1 (mod p) has
exactly n incongruent solutions modulo p.
Proof. Let p − 1 = nd. Recall the factorization
X p−1 − 1 = X nd − 1 = (X n − 1)(X n(d−1) + X n(d−2) + · · · + 1).
26 2. MULTIPLICATIVE STRUCTURE MODULO m
Quadratic Reciprocity
Proof. (i) follows straightaway from the definition, and (iii) fol-
lows from (ii). Let’s prove (ii). Let a be an integer. If p | a then
a
= 0 ≡ a(p−1)/2 (mod p).
p
Quadratic Reciprocity.
94 2 47
= by Proposition 3.3
257 257 257
47
= using the second supplement
257
257
= since 257 ≡ 1 (mod 4)
47
22
= 257 ≡ 22 (mod 47)
47
2 11
=
47 47
11
=
47
47
=− 11 ≡ 47 ≡ 3 (mod 4)
11
3
=−
11
11
= 3 ≡ 11 ≡ 3 (mod 4)
3
2
= 11 ≡ 2 (mod 3)
3
= −1 using the second supplement.
The right-hand sides of the last two equations are identical except for a
minus sign for each term in the product. But there are (#Sp )(#Sq ) =
(p − 1) (q − 1)
terms in the product. Thus
2 2
q p (p−1) (q−1)
= (−1) 2 2 ,
p q
36 3. QUADRATIC RECIPROCITY
p-adic Numbers
1. Congruences Modulo pm
In quadratic reciprocity we studied congruences of the form x2 ≡ a
(mod p). We now turn our attention to situations where p is replaced
by a power of p.
We shall need the following lemma whose proof is an easy exercise,
but try out a few examples first to convince yourself that it is true.
Lemma 4.1. Let f (X) ∈ Z[X] and let n > 0 be an integer. Then
f (n) (X)/n! has integer coefficients.
Next is Hensel’s Lemma which is the main result of this section.
Theorem 4.2. (Hensel’s Lemma) Let f (X) ∈ Z[X]. Let p be a
prime and m ≥ 1. Suppose a ∈ Z satisfies
f (a) ≡ 0 (mod pm ), f 0 (a) 6≡ 0 (mod p).
Then there exists some b ∈ Z such that
(3) b ≡ a (mod pm ), f (b) ≡ 0 (mod pm+1 ).
We say that we lift a to a solution modulo pm+1 .
Proof of Hensel’s Lemma. By Taylor’s Theorem
f (2) (a) 2 f (n) (a) n
f (a + x) = f (a) + f 0 (a)x + x + ··· + x
2! n!
where n is the degree of f (note that all higher derivatives vanish). We
want b to satisfy two conditions, one of them that b ≡ a (mod pm ).
Let us write b = a + pm y where the integer y will be determined later.
Then
f (b) = f (a) + pm f 0 (a)y + p2m (integer).
Since f (a) ≡ 0 (mod pm ) we have f (a) = pm c where c is an integer.
Thus
f (b) = pm (c + f 0 (a)y) + p2m (integer).
Note that pm+1 | p2m . To make f (b) ≡ 0 (mod pm+1 ) it is enough to
choose y so that p | (c + f 0 (a)y). In otherwords, we want y so that
f 0 (a)y ≡ −c (mod p). But f 0 (a) 6≡ 0 (mod p) and so is invertible
39
40 4. p-ADIC NUMBERS
or equivalently
7(1 + 6y) ≡ 0 (mod 72 )
m solutions to x2 ≡ 2 (mod 7m )
1 ±3
2 ±(3 + 7)
3 ±(3 + 7 + 2 × 72 )
4 ±(3 + 7 + 2 × 72 + 6 × 73 )
5 ±(3 + 7 + 2 × 72 + 6 × 73 + 74 )
We are writing solutions as a series in powers of 7 with coefficients
between 0 and 6. This suggests very much an analogy with decimal
expansions. We immediately begin to wonder if the series converges
in any sense. Of course it does not converge in the sense of 1st year
analysis as the powers of 7 are tending to infinity. However we will
change our notion of large and small to make it converge.
where P is the set of all primes. Of course only finitely many of the
exponents ordp (α) are non-zero, so the product makes sense.
Definition. Let p be a prime and α a non-zero rational number. We
define the p-adic absolute value of α to be
|α|p = p− ordp (α) .
We define |0|p = 0 which is consistent with our convention that ordp (0) =
+∞.
Example 2.1. Let α = −50/27. Then
2−1 p = 2
33
p=3
|α|p = −2
5 p=5
1 p 6= 2, 3, 5.
Q
Now evaluate p∈P |α|p . What do you notice.
42 4. p-ADIC NUMBERS
Here it is easy to add two elements in the disc so that you leave the
disc. The triangle inequality for the usual absolute value will tell you
that if |α| ≤ C and |β| ≤ C then |α + β| ≤ 2C, so you can see that the
ultrametric inequality is much stronger than the triangle inequality.
Theorem 4.4. (The Product Formula) Let α be a non-zero rational
number. Then Y
|α| |α|p = 1,
p∈P
where P is the set of primes.
Proof. Prove this using (4). Notice that all but finitely many
terms in the product are 1, so the product makes sense.
3. Convergence
Definition. We say that the series of rational numbers {an }∞
n=1 con-
verges p-adically to a ∈ Q if
lim |an − a|p = 0.
n→∞
only if it converges; but here we are talking about real numbers, not
just rational numbers. For example, you know that the sequence
n
1
an = 1 +
n
is a Cauchy sequence of rational numbers that converges to e which is
not rational but real. But what is a real number? The best way to
define real numbers is to say that a real number is simply a Cauchy
sequence of rational numbers! Think about it. This motivates our next
definition.
Definition. A p-adic number α is a p-adically Cauchy sequence {an }∞
n=1
of rational numbers. We write Qp for the set of p-adic numbers. We
identify Q as a subset of Qp via the map
(6) Q → Qp , a 7→ {a}∞
n=1 .
Let’s go back to the reals for a moment to make sure that our
definition makes sense. We said that a real number is simply a Cauchy
sequence of rationals. So e is just the sequence (1 + 1/n)n . But there
are other sequences converging to e. For example, take the partial sums
of the series
1 1 1
1 + + + + ··· .
1! 2! 3!
So to say that a real number is a Cauchy sequence seems an ambiguous
way to define real numbers. However, the ambiguity disappears as soon
as we adopt the convention that two Cauchy sequences define the same
real number if their difference is a null sequence. We do the same in
the p-adic setting.
Definition. We say that two p-adic numbers {an } and {bn } are equal
if the difference {an − bn } is p-adically null.
Example 3.6. Via the identification (6) we think of 0 ∈ Q to be the
same as the zero sequence {0} in Qp . Now the {pn } and {0} are both
p-adically null sequences and we have that
0 = {0} = {pn } = any null sequence of rationl numbers.
Lemma 4.6. Suppose that the sequence of rational numbers {an } con-
verges p-adically to a ∈ Q. Then in Qp
lim an = a = {an }∞
n=1 .
n→∞
4. Operations on Qp
Of course Qp would not be very interesting if it was a set with no
additional structure. In fact we can define addition and multiplication
on Qp in a natural way:
{an } + {bn } = {an + bn }
and
{an } · {bn } = {an bn }.
One must check that these operations are well-defined. For a start
we want to make sure that the sequences {an + bn } and {an bn } are
p-adically Cauchy so that we are staying in Qp . We also want to
check that if {an } and {a0n } differ by a p-adically null sequence and
if {bn } and {b0n } differ by a p-adically null sequence then {an + bn } and
{a0n + b0n } differ by a p-adically null sequence and {an bn } and {a0n b0n }
differ by a p-adically null sequence. These we’ll leave as relatively easy
exercises. We also want to check that the usual properties of addition
4. OPERATIONS ON Qp 47
Proof. Note that the convergence we’re talking about in the sec-
ond sentence of the lemma is covergence with respect to the usual
absolute value. Now certainly |an |p is in the set {0} ∪ {pr : r ∈ Z},
and it’s easy to see that any Cauchy subsequence of {0} ∪ {pr : r ∈ Z}
must actually converge to some element of this set. Thus all we have
to show is that {|an |p } is Cauchy with respect to the usual absolute
value. Now it is an easy exercise to check that
|a|p − |b|p ≤ |a − b|p .
Hence
0 ≤ |am |p − |an |p ≤ |am − an |p .
As {an } is p-adically Cauchy, limm,n→∞ |am − an |p = 0. Hence by the
Sandwich Theorem,
lim |am |p − |an |p = 0.
m,n→∞
This shows that the sequence {|an |p } is Cauchy with respect to the
usual absolute value and completes the proof.
The above lemma allows us to define division. If {an } and {bn } are
elements of Qp and {bn } = 6 0 (i.e. non-null) then there is some N such
that for n ≥ N , bn 6= 0 and we define {cn } = {an }/{bn } by assigning
cn randomly for n < N and letting cn = an /bn for n ≥ N . Note that
{cn } · {bn } agrees with {an } except for finitely many terms and so their
difference is null; in other words {cn } · {bn } = {an } in Qp .
Theorem 4.9. Qp is a field containing Q as a subfield.
Hence
lim |bn |p = lim |an |p ,
n→∞ n→∞
5. Convergence of Series
The ultrametric inequality has a dramatic effect of making the con-
vergence of series very easy to check.
Theorem 4.11. Let p be a prime. The series ∞
P
j=1 aj converges p-
adically if and only if limj→∞ |aj |p = 0.
We know that with the usual absolute value the theorem is true
only in the left to right direction. The famous counterexample being
the harmonic series
1 1 1
1 + + + + ··· ,
2 3 4
which diverges even though limj→∞ 1/j = 0. Working p-adically, we
don’t need any of the complicated convergence tests of first-year analysis—
the theorem makes it all very easy!
Proof of Theorem 4.11. Suppose that limn→∞ |an |p =P0. All
we have to do is to show the the sequence of partial sums sn = nj=1 aj
is Cauchy. A Cauchy sequence converges to some element of Qp (which
happens to equal the sequence itself). Now suppose m > n. Then
|sm − sn |p = |an+1 + an+2 + · · · + am |p = max |aj |p ,
n+1≤j≤m
by the ultrametric inequality. For any > 0, there is some N such that
if j ≥ N then |aj |p < . Hence if m, n ≥ N then |sm − sn |p < , proving
that the sequence {sn } is p-adically Cauchy.
6. p-adic Integers
Definition. The set of p-adic integers Zp is defined by
Zp = {α ∈ Qp : |α|p ≤ 1}.
50 4. p-ADIC NUMBERS
Consider ordp (an ). If there are infinitely many n such that ordp (an ) ≤
−1 then there are infinitely many n such that |an |p ≥ p and this con-
tradicts the above. Hence there is some N such that ordp (an ) ≥ 0 for
all n ≥ N . So we can write
un
an =
vn
where un , vn ∈ Z, with p - vn . Since p - vn , we know that vn is
invertible modulo pn . Let vn wn ≡ 1 (mod pn ), where wn ∈ Z and
write bn = un wn ∈ Z. Then an = un /vn ≡ un wn = bn (mod pn ) and so
|an − bn |p ≤ p−n . This completes the proof.
7. HENSEL’S LEMMA REVISITED 51
is an easy exercise.
52 4. p-ADIC NUMBERS
c
Proof. Suppose b = p2r c where r ∈ Z and = 1. All we have
p
to show is that c is a square in Zp . Let f (X) = X − c. Since pc = 1,
2
these classes of polynomials we say that the Hasse Principle holds. But
it is false for many other classes of polynomials and for those we say
that the Hasse principle fails. Here is a counterexample to the Hasse
principle for polynomials in 1 variable.
Example 8.1. Let f (X) = (X 2 − 2)(X 2 − 17)(X 2 − 34). Show that
f (X) = 0 is a counterexample to the Hasse principle.
Answer: Basically we are asked to show that f (X) = 0 has solutions
in Zp for all primes p and in R, √ but has
√ no solutions
√ in Z. It clearly
has solutions in R, which are ± 2, ± 17, ± 34, and clearly it has
no solutions in Z as none of these roots are integral.
Now 17 ≡ 1 (mod 8) and so by Corollary 4.18 17 = α2 for some
α ∈ Z2 . Then f (α) = 0, so f (X) = 0 has a solution in Z2 . Also
2
17
= 1, so 2 is a square in Z17 by Corollary 4.15, and so f (X) = 0
has a solution in Z17 . Suppose that p 6= 2, 17. We want to show that
f (X) = 0 has a solution in Zp . Equivalently, we want to show that
at least one of 2, 17, 34 is a square in Zp . Suppose that 2, 17 are not
squares in Zp . By Corollary 4.15,
2 17
= −1, = −1.
p p
But multiplying we obtain
34 2 17
= = 1.
p p p
x4 ≡ 17 (mod p).
8. THE HASSE PRINCIPLE 55
Geometry of Numbers
However
√ x2 + y 2 + z 2 < r2 − w2 is a ball (sphere) in xyz-space of radius
r2 − w2 , so
ZZZ
4π 2
dxdydz = (r − w2 )3/2 .
2 2 2 2
x +y +z <r −w 2 3
60 5. GEOMETRY OF NUMBERS
Hence
Z w=r Z w=r
4π 2 2 3/2 8π
V (Br ) = (r − w ) dw = (r2 − w2 )3/2 dw.
3 w=−r 3 w=0
You immediately say to yourself that this needs are trigonometric sub-
stitution, and you’re right: let w = r sin θ, so dw = r cos θdθ, and
so
8π θ=π/2 3 8πr4 π/2
Z Z
2 3/2
V (Br ) = r (1 − sin θ) r cos θdθ = cos4 θdθ.
3 θ=0 3 0
We need to integrate cos4 , and one way of doing this is using multiple-
angle formulae. See your Vectors and Matrices lecture notes if you
haven’t ceremoniously incinerated them at the end of your first year.
But just in case, here is how it works. Write
eiθ + e−iθ
cos θ = .
2
Taking fourth powers we get
1 4iθ
cos4 θ = e + 4e2iθ + 6 + 4e−2iθ + e−4iθ
16
which we can rewrite as
1 1 3
cos4 θ = cos 4θ + cos 2θ + .
8 2 8
Hence Z π/2
3π
cos4 θdθ = .
0 16
We deduce that the volume of the ball of radius r in 4-space is
π 2 r4
(9) V (Br ) = .
2
3. The Four Squares Theorem
Theorem 5.3. Every positive integer n can be written as the sum of
four integer squares.
This is a statement that your non-mathematical parents would un-
derstand. If they ask you what you’ve learned in three or four years
on a maths degree you can mention this, and they’ll be very impressed
and think that your education has been worthwhile. Most of your other
modules give you statements that are pure gobbledygook to the unini-
tiated. Galois Theory gives a few statements that your parents might
understand but they’re all negative: you can’t solve a quintic, or con-
struct a heptagon, or trisect an angle. Number Theory gives positive
3. THE FOUR SQUARES THEOREM 61
assertions that broaden your horizons, and expand the frontiers of your
knowledge . . .
If you’ve survived reading the previous paragraph without vomiting
then you have strong constitution and is ready for the proof of the Four
Squares Theorem.
Proof of the Four Squares Theorem. First we prove the state-
ment of the theorem for primes. If p = 2 then we can write p =
12 + 12 + 02 + 02 , so assume that p is an odd prime. By one of the
exercises on the early homework assignments—unassessed due to lack
of foresight on my part—you know that there integers a, b such that
a2 + b 2 + 1 ≡ 0 (mod p).
Let
Λ = {(x, y, z, w) ∈ Z4 : x ≡ az+bw (mod p), y ≡ bz−aw (mod p)}.
This common-sensically is a sublattice of Z4 of index p2 .
We also take
C = {(x, y, z, w) ∈ R4 : x2 + y 2 + z 2 + w2 < 2p}.
√
This is a ball of radius 2p, so is convex and symmetric and by (9) we
have
π2 p
V (C) = ( 2p)4 = 2π 2 p2 > 24 p2 .
2
Hence the hypotheses of Minkowski are satisfied. So we have a point
(x, y, z, w) common to both Λ and C that is not (0, 0, 0, 0). As (x, y, z, w)
is in Λ, the coordinates are integers and
x2 + y 2 + z 2 + w2 ≡ (az + bw)2 + (bz − aw)2 + z 2 + w2
= (a2 + b2 + 1)(z 2 + w2 ) ≡ 0 (mod p).
However, as (x, y, z, w) is a non-zero point of C,
0 < x2 + y 2 + z 2 + w2 < 2p,
so x2 + y 2 + z 2 + w2 is an integer strictly between 0 and 2p that is
divisible by p. The inescapable conclusion is x2 + y 2 + z 2 + w2 = p.
This proves the theorem for primes. To complete the proof we need
the identity
(10) (a2 + b2 + c2 + d2 )(x2 + y 2 + z 2 + w2 ) =
(ax−by−cz−dw)2 +(ay+bx+cw−dz)2 +(az−bw+cx+dy)2 +(aw+bz−cy+dx)2
Now if n > 1 is a positive integer then you can write as a product
of primes and use the identity repeatedly to write n as a sum of four
squares.
62 5. GEOMETRY OF NUMBERS
Notes:
• The Four Squares Theorem was proved by Joseph Louis La-
grange in 1770, though the theorem appears–without proof–
in the Arithmetica of Diophantus (probably written around
250AD). We have followed Davenport’s proof of the Four Squares
Theorem (1941).
• Another fascinating question is, in how many ways can we
write a positive integer n as the sum of four squares? This
was answered in 1834 by Carl Jacobi. He showed that this
number is eight times the sum of the divisors of n if n is odd,
and 24 times the sum of the odd divisors of n if n is even.
Jacobi’s theorem has remarkable proof using modular forms.
• Where does identity in (10) come from? You are surely familiar
with the multiplicative property of norms of Gaussian integers.
If α = a + bi ∈ Z[i] then the norm of α is defined by N (α) =
a2 + b2 , and you know N (αβ) = N (α)N (β). The identity in
(10) is the corresponding identity for quaternion norms.
where
z + W = {z + w : w ∈ W }.
Thus we can rewrite (11) as
XZ
V (S) = χS (z + w) dw.
z∈Zn w∈W
2
Interchanging the summation and integration signs we obtain
Z !
X
V (S) = χS (z + w) dw.
w∈W z∈Z
P
Write f (w) = z∈Z χS (z + w), and recall that V (S) > m is a hypoth-
esis of the theorem. Hence
Z
f (w)dw > m.
w∈W
But W has volume P 1. Hence there is some point w ∈ W such that
f (w) > m; i.e. z∈Z χS (z + w) > m for that particular w. But the
χS (z+w) are ones and zeros, so there are m+1 distinct z0 , . . . , zm ∈ Zn
such that χS (zi + w) = 1. Write xi = zi + w, so the xi are distinct.
Now note that χS (xi ) = 1, so by definition of χS , the xi are in S.
Finally
xj − xi = (zj + w) − (zi + w) = zj − zi ∈ Zn ,
which completes the proof.
Here is the statement of Minkowski again, with proof.
Theorem 5.5. (Minkowski’s Theorem) Let Λ be a sublattice of Zn
of index m. Let C be a convex symmetric subset of Rn having volume
V (C) satisfying
V (C) > 2n m.
Then C and Λ have a common point other than 0.
1For example in R2 , we write (1.7, 5.9) = (1, 5) + (0.7, 0.9) where we note that
(1, 5) ∈ Z2 and (0.7, 0.9) ∈ W .
2To justify interchanging integration with infinite summation one needs rather
delicate theorems in Lebesgue Integration. Fortunately/unfortunately for you, I’ve
forgotten my Lebesgue and so I can’t tell you about it. But beware, Analysis
lecturers with no sense of humour don’t like to see this sort of thing without justi-
fication; they would regard my lecture notes as mathematical pornography.
64 5. GEOMETRY OF NUMBERS
Proof. Let
1 1
S= C= x:x∈C .
2 2
The volume of S is
1
V (S) = V (C) > m.
2n
By Blichfeldt’s Theorem, there are m+1 distinct points x0 , . . . , xm ∈ S
such that
xj − xi ∈ Zn , for 0 ≤ i, j ≤ m.
n
Let yj = xj − x0 ∈ Z for j = 0, . . . , m. These are m + 1 distinct
points yj in Zn and Λ has m cosets in Zn . So two distinct yi , yj lie in
the same coset of Λ. Thus, xj − xi = yj − yi is a non-zero element of
Λ. Now we can write xj = c/2 and xi = c0 /2 where c and c0 are in C.
Hence
c − c0
2
is a non-zero element of Λ. Now C is symmetric so, −c0 ∈ C as well
as c ∈ C. Finally C is convex and (c − c0 )/2 is the mid-point between
c and −c0 , so it must be in C as well as being a non-zero element of
Λ. This is the point whose existence is asserted in the statement of the
theorem.
CHAPTER 6
2. The irrationality of e
So far the only irrational numbers we’ve seen are roots of polynomi-
als. It is natural to wonder about the irrationality of naturally occuring
numbers such as e = exp(1). In fact Euler proved that e is irrational.
Theorem 6.4. (Euler) e = exp(1) is irrational.
Proof. The proof starts with the familiar power series expansion
∞
X xn
exp(x) = .
n=0
n!
Thus
1 1
e=1+1+ + + ··· .
2! 3!
Suppose that e is rational, and write e = a/b where a, b are positive
coprime integers. Now
1 1
(b − 1)!a = b!e = b! 1 + 1 + + + · · ·
2! 3!
1 1 1 1
= b! 1 + 1 + + · · · + + b! + + ··· .
2! b! (b + 1)! (b + 2)!
Write
1 1
α = b! 1 + 1 + + · · · +
2! b!
and note that α is an integer. Thus (b − 1)!a − α is an integer. Write
β = (b − 1)!a − α ∈ Z. We see that
1 1
β = b! + + ··· .
(b + 1)! (b + 2)!
3. WHAT ABOUT TRANSCENDENTAL NUMBERS? 67
f (x) = a0 + a1 x + · · · + ad xd .
Then
pd
p p
f = a0 + a1 + · · · ad d
q q q
N
= d
q
1) Let p be a prime.
a) What does it mean for an integer g to have order d modulo p?
[2]
b) Show that if g has order d modulo p and if g m ≡ 1 (mod p)
then d | m. [6]
c) Suppose g1 and g2 respectively have orders d1 , d2 modulo p.
Suppose moreover that gcd(d1 , d2 ) = 1. Show that g1 g2 has
order d1 d2 modulo p. [6]
d) What does it mean for g to be a primitive root modulo p? [2]
e) Show that p must have a primitive root. You may assume that
e
if q e is a prime power dividing p − 1 then xq ≡ 1 (mod p) has
precisely q e incongruent solutions modulo p. [6]
f) Find a primitive root for 149. You may use the following
observations: [3]
149 = 22 ×37+1, 537 ≡ 444 ≡ 1 (mod 149), 442 6≡ 1 (mod 149).
71
72 X. LAST YEAR’S EXAM
2)
a) Let a be an integer and p an odd prime. Show that
a
≡ a(p−1)/2 (mod p).
p
You may assume standard facts about primitive roots. [7]
b) State without proof the two supplements to the law of qua-
dratic reciprocity. [4]
c) Let x be an even integer. Show that every prime divisor p of
x4 + 1 satisfies
−1 2
= = 1,
p p
and hence p ≡ 1 (mod 8). Hint: You might find it helpful to
observe that x4 + 1 = (x2 + 1)2 − 2x2 . [7]
d) Deduce that there are infinitely many primes p ≡ 1 (mod 8).
[7]
3)
a) State Blichfeldt’s Theorem and Minkowski’s Theorem. [6]
b) Give a proof of Minkowski’s Theorem assuming Blichfeldt’s
Theorem. [6]
c) Let a, b > 0. Show that the area of the ellipse
x2 y 2
+ 2 <1
a2 b
is πab. You may assume the formula for the area of a circle.
[6]
d) Suppose λ and N are coprime positive integers satisfying
λ2 ≡ 2 (mod N ).
Show that there are integers x, y such that [7]
2 2
x − 2y = ±N.
Hint: In Minkowski’s Theorem, take the convex symmetric
set to be
C = {(x, y) ∈ R2 : x2 + 2y 2 < 2N }.
X. LAST YEAR’S EXAM 73
4)
a) Let f (X) ∈ Z[X], a ∈ Z and n a positive integer. Show that
f (n) (a)/n! is an integer. [4]
b) Let f (X) ∈ Z[X]. Let p be a prime and m ≥ 1. Suppose
a ∈ Z satisfies
f (a) ≡ 0 (mod pm ), f 0 (a) 6≡ 0 (mod p).
Show that there exists some b ∈ Z such that [8]
b ≡ a (mod pm ), f (b) ≡ 0 (mod pm+1 ).
c) Solve the following simultaneous system of congruences [4]
2 3 2
x ≡3 (mod 5 ), x ≡6 (mod 7).
d) Solve the following simultaneous system of congruences [9]
3 3
y ≡3 (mod 5 ), y≡1 (mod 4).
5) Let p be a prime.
a) Let α be a rational number. Define ordp (α) and |α|p . [2]
b) Let α, β be rational numbers. Prove that
ordp (α + β) ≥ min{ordp (α), ordp (β)},
and [8]
|α + β|p ≤ max{|α|p , |β|p }.
c) Prove that the series of rational numbers ∞
P
n=1 an converges
in Qp if and only if limn→∞ |an |p = 0. You may assume that a
sequence converges in Qp if and only if it is p-adically Cauchy.
[7]
d) State—with proof—for which primes p do the following series
converge in Qp ?
(i) 1 + (21/2)2 + (21/2)4 + (21/2)8 + · · · . [4]
(ii) 11 + 22 + 33 + 44 + · · · . [4]
APPENDIX Y
Mathematical Pornography
Hence
y
x = x1/i = 1 + = 1 + y.
i
Rearranging we get
1 1
(15) log(1 + t) = log x = y = (x − 1) = (−1 + (1 + t) ) .
Now by the Binomail Theorem
( − 1) 2 ( − 1)( − 2) 3 ( − 1)( − 2)( − 3) 4
(1+t) = 1+t+ t+ t+ t +· · ·
2! 3! 4!
However, by (14) we can eliminate all higher powers of . Thus
− 2 2! 3 −3! 4
(1 + t) = 1 + t + t + t + t + ···
2! 3! 4!
= 1 + t − t2 + t3 − t4 + · · ·
2 3 4
Substituting into (15) we obtain
1 2 3 4
log(1 + t) = t − t + t − t + · · ·
2 3 4
t2 t3 t4
= t − + − + ···
2 3 4
1. An Integral Equation
R Here
R x is an beautiful example I found on mathoverflow.org. Let
“ = 0 ”. We want to solve the integral equation
Z
f − f = 1.
Hence
Z −1
f= 1− 1
Z ZZ ZZZ
= 1+ + + +··· 1
Z x Z xZ x Z xZ xZ x
=1+ 1+ 1+ 1 + ···
0 0 0 0 0 0
x2 x3
=1+x+ + + ···
2! 3!
= ex .