Lecture Notes
Lecture Notes
Lecture Notes
These notes are modified from previous versions (due to Neil Dummigan,
Alan Lauder and Roger Heath-Brown) and have been recently revised by
me. They draw mainly upon “A Classical Introduction to Modern Num-
ber Theory”, by Ireland and Rosen, and “Algebraic Number Theory”, by
Stewart and Tall. While I take full responsibility for their current contents,
considerable thanks are clearly due to Neil, Alan and Roger.
Lectures will omit some of the non-examinable proofs, and some of the
examples in Section 9. However these should prove helpful for examination
revision.
Useful texts
Algebraic Number Theory and Fermat’s Last Theorem, I. Stewart and D.
Tall, Third Edition. We shall frequently cite this as “S&T”.
Older editions under the name “Algebraic Number Theory” will also suf-
fice.
Other useful but more advanced references:
A Classical Introduction to Modern Number Theory, (Chapter 12) K. Ireland
and M. Rosen
Algebraic Number Theory, A. Frohlich and M.J. Taylor
A Course in Computational Algebraic Number Theory, H. Cohen.
1
1 Introduction
1.1 Motivation
Consider “Fermat’s Last Theorem” which asserts that xn + y n = z n has no
solution x, y, z ∈ Z (x, y, z all nonzero) if n ∈ N, n > 3. It is sufficient to
prove this for n = 4 and n = p > 3 prime [since any n > 3 is divisible by 4
or some prime p > 3; if n = 4k, then any solution to xn + y n = z n would give
(xk )4 + (y k )4 = (z k )4 ; similarly if n = pk, then any solution to xn + y n = z n
would give (xk )p + (y k )p = (z k )p ].
Fermat himself proved the result for n = 4 after which it remained to
prove it for n = p > 3 prime. Let ζp = exp(2πi/p) ∈ C and let K := Q(ζp ).
Factoring the left hand side in Z[ζp ] we see that
(x + y)(x + ζp y) . . . (x + ζpp−1 y) = z p .
2
i = 0, . . . , n − 1, and p2 does not divide a0 . Then, apart from constant
factors, f (t) is irreducible over Z, and hence irreducible over Q.
Definition 1.3. A number field (or algebraic number field) is a finite ex-
tension K of Q. The index [K : Q] is the degree of the number field.
σi : K → C (i = 1, . . . , n).
The elements σi (θ) are the distinct zeros in C of the minimal polynomial mθ
of θ over Q.
3
Lemma 1.7. Let G be a free abelian group of rank n with Z-basis {x1 , . . . , xn }.
Suppose (aij ) is an n × n matrix with integer entries. Let
X
yi = aij xj (1 6 i 6 n).
j
Then the elements {y1 , . . . , yn } form a Z-basis for G if and only if (aij ) is
unimodular.
Proof. NE. See S&T, page 28, Lemma 1.15.
Theorem 1.8. Let G be a free abelian group of rank n, and H a subgroup.
Then G/H is finite if and only if H has rank Pn. Moreover, if G and H have
Z-bases x1 , . . . , xn and y1 , . . . , yn with yi = j aij xj we have
#G/H = | det(aij )|.
Proof. NE. See S&T, page 30, Theorem 1.17.
4
Definition 2.2. Let w = {w1 , . . . , wn } be an n-tuple of elements of K, where
n = [K : Q].
• The determinant is ∆(w) := det(σi (wj )), i.e., the determinant of the
n × n matrix whose (i, j)th entry is σi (wj ).
*Warning*: S&T and some other books use ∆ where we write ∆2 (!).
Proof. We have
1 α1 α12 . . . α1n−1
1 α2 α22 . . . α2n−1
∆(v) = .. ..
. .
1 αn αn2 . . . αnn−1
5
(This is a so-called van der Monde determinant.) We can view this as a
polynomial of degree n(n − 1)/2 in α1 , . . . , αn . Since it vanishes when we
set αi = αj the polynomial is divisible by αi − αj for all i < j. There
are n(n − 1)/2 of these factors. Hence, on checking that the coefficient of
α2 α32 . . . αnn−1 is +1 we deduce that
Y
∆(w) = (αj − αi ).
i<j
6
√
Example 3.3 Let K = Q( d), where d ∈ Z, d 6= ±1, with d squarefree
2
(i.e. there is
√ no prime p for which p |d). Then
√ [K : Q] = 2, and K has√a
Q-basis {1, d}. If√a, b ∈ Q then α = a + b d ∈ K. Since σ1 (α) = a + b d
and σ2 (α) = a − b d we have TrK/Q (α) = 2a and NormK/Q (α) = a2 − db2 .
Moreover mα (x) = x2 − 2ax + (a2 − db2 ) (if b 6= 0). Hence α ∈ OK if and
only if 2a, a2 − db2 ∈ Z.
Suppose α ∈ OK . Then (2a)2 − d(2b)2 ∈ Z, giving d(2b)2 ∈ Z. Writing
2b = u/v (u, v ∈ Z) we have du2 v −2 ∈ Z, so that v 2 |du2 . Since d is squarefree
this implies v|u, giving 2b ∈ Z. Write 2a = A, 2b = B, with A, B ∈ Z.
Then a2 − db2 ∈ Z, so that A2 ≡ dB 2 mod 4. Now observe that any square
is congruent to 0 or 1 modulo 4.
7
Example 3.5
1. Let R be a field and M a vector space over R. Then M is an R-module.
2. If R = Z and M is any additive abelian group then M is an R-module.
8
Then M N is finitely generatedy (by {vi wj : 1 6 i 6 d, 1 6 j 6 e}) and in
K. Moreover,
However NormK/Q (α) and NormK/Q (α−1 ) are in Z, so both are ±1.
Conversely, suppose that NormK/Q (α) = ±1. Let α1 , . . . , αn be the K/Q-
conjugates, with α = α1 , say. Then α1 . . . αn = ±1, so that α(α2 . . . αn ) =
±1. Hence α−1 = ±(α2 . . . αn ), which by Theorem 3.7 lies in OL . However
we know that α−1 lies in K, and so α−1 ∈ OL ∩ K = OK .
9
Note 3.13 Let v = {v1 , . . . , vn } and w = {w1 , . . . , wn } be any two Q-bases
of K. Let M = hv1 , . . . , vn iZ and N = hw1 , . . . , wn iZ , as Z-submodules of
K. Suppose v, w ⊆ OK , then ∆(v)2 , ∆(w)2 ∈ Z. (Recall that ∆(v)2 =
det(TrP
K/Q (vi vj )).) Suppose N ⊆ M . Then there exist cij ∈ Z such that
wi = nj=1 cij vj . Let C = (cij ). Then by Theorem 1.8 we have
Theorem 3.15 (Integral Basis Theorem). The ring OK has an integral basis
(that is, a Z-basis).
10
Let w be a new Q-basis for K obtained from v by replacing vj by α. Then
w ⊆ OK . The change of basis matrix
1 0 ... 0
0 1 ... 0
. .
. ..
.
C=
c1 . . . c2 . . . cn
. ..
..
.
0 0 ... 0 1
11
2. If [OK : M ] = m, then |∆(M )2 | = m2 |∆(OK )2 |. If ∆(M )2 is squarefree
then m = 1 and OK = M . Otherwise (and if OK 6= M ), by Proposi-
tion 3.16, there exist p prime with p2 |∆(M )2 and c1 , . . . , cn ∈ Z, not
all divisible by p, such that p1 (c1 w1 + . . . + cn wn ) ∈ OK .
√ √
Example√ 3.17 K = Q( d), d squarefree. Start with Q-basis {1, d}.
Then {1, d} ⊆ OK and
√ 2
√
1 − d
2 √
∆({1, d}) =
= 4d.
1 + d
√
Since d is squarefree the only prime p such that p2 |∆({1, d})2 is p = 2.
12
√
• Case 1: d ≡ 1 mod 4. We find 21 (1 + d) ∈ OK (This element has
minimal polynomial x2 − x + (1 − d)/4 ∈ Z[x]). In this case we find
1 √ 1
∆({1, (1 + d)})2 = 2 4d = d.
2 2
√
• Case 2: d 6≡ 1 mod 4. Then 21 (1 + d) 6∈ OK since x2 − x + 1−d 6∈ Z[x].
1 1
√ 4
The only other cases to check are 2 , 2 d, which are not in OK . Since
we did not find√any “α”, we conclude
√ that 2 does not divide the index
m = [OK : h1, diZ ]. Hence {1, d} is an integral basis.
4 Cyclotomic fields
None of the proofs in this section are examinable!
Let p > 2 be a prime and ζp := e2πi/p , so that ζpp = 1. Let K = Q(ζp ), a
cyclotomic field. Clearly ζ := ζp satisfies
xp − 1
f (x) = = xp−1 + xp−2 + · · · + x + 1.
x−1
Lemma 4.1. f (x) is irreducible in Q[x].
(x + 1)p − 1
g(x) = = xp−1 + pxp−2 + · · · + p.
(x + 1) − 1
Since p divides all the coefficients apart from the first, but p2 does not divide
the final coefficient p, the polynomial g(x) is irreducible over Z by Eisenstein’s
criterion and so over Q by Gauss’ Lemma.
Corollary 4.2. [K : Q] = p − 1.
13
Note 4.3
Qp−1
1. NormK/Q (1 − ζ) = i=1 (1 − ζ i ) = f (1) = p
(x − 1)pxp−1 − (xp − 1)
f ′ (x) =
(x − 1)2
and so
−pζ p−1
f ′ (ζ) = .
1−ζ
Hence from Note 4.3 above,
NormK/Q (−p)NormK/Q (ζ)p−1 (−p)p−1 1p−1
′
NormK/Q (f (ζ)) = = = pp−2
NormK/Q (1 − ζ) p
as required.
Theorem 4.5. The set {1, ζ, . . . , ζ p−2 } is an integral basis for OK .
Proof. Let θ = ζ − 1. Certainly we have Z[θ] = Z[ζ]. We shall show that
{1, θ, . . . , θp−2 } is an integral basis.
By Lemma 4.4 and Note 3.13 we see that
14
Hence p is the only prime whose square divides ∆(Z[θ])2 . It follows that p is
the only prime which may divide [OK : Z[θ]]. If OK 6= Z[θ] then there exists
α ∈ OK such that
p−2
1X j
α= cj θ ,
p j=0
with cj ∈ Z not all divisible by p. Let r be minimal such that p does not
divide cr . We may assume cj = 0 for j < r by subtracting integer multiples
of the basis elements. Now αθp−2−r ∈ OK , since α and θ are in OK . Write
1
θp−2−r α = (cr θp−2 + cr+1 θp−1 + · · · + cp−2 θ2p−4−r ). (4.1)
p
Then
p(p − 1) p−3
θp−1 = −pθp−2 − θ − ··· − p
2
and so p−1 θp−1 ∈ OK . Hence by subtracting multiples of this from both sides
of (4.1) we see that p−1 cr θp−2 ∈ OK . However
p−1
cp−1
cr p−2 cr
NormK/Q θ = p p−2
= r ,
p p p
since NormK/Q (θ) = p and NormK/Q (cr /p) = (cr /p)p−1 . This, finally, con-
tradicts the fact that NormK/Q (α) ∈ Z for all α ∈ OK , since p does not
divide cr .
Definition 5.1.
15
3. A nonzero, non-unit element α ∈ R is irreducible if (α = βγ ⇒ β or
γ is a unit). We write β|α if there exists γ ∈ R such that α = βγ.
1. α = β1 . . . βn
R a ED ⇒ R a PID ⇒ R a UFD.
16
5.2 Some applications of unique factorisation
First, a useful lemma:
Lemma 5.5. Let OK be the ring of integers in a number field K, and α, β ∈
OK . Then
1. α is a unit (in OK ) if and only if NormK/Q (α) = ±1.
2. If α and β are associates (in OK ) then NormK/Q (α) = ±NormK/Q (β).
3. If NormK/Q (α) is a rational prime, i.e. a prime number in Z, then α
is irreducible in OK .
Proof. 1. Proposition 3.10.
2. We have α = uβ with u a unit, and so:
NormK/Q (α) = NormK/Q (u)NormK/Q (β) = ±NormK/Q (β), by part 1.
3. Let α = γδ. Then NormK/Q (α) = p = NormK/Q (γ)NormK/Q (δ) for
some prime p ∈ Z. The result now follows from 1.
The converses of 2 and 3 are false (see later the proof of Proposition 5.8).
17
Now Z is a UFD and neither a+bi or c+di has norm ±1, giving p = a2 +b2 =
(a + bi)(a − bi). This yields the existence part of the theorem.
If a + bi = αβ in Z[i] then, taking norms, we find that
p = Norm(α)Norm(β).
Thus α or β must be a unit. Hence a + bi is irreducible in Z[i], and similarly
for a − bi. Thus p = (a + bi)(a − bi) is the unique factorisation of p into
irreducibles.
If also p = e2 + f 2 = (e + f i)(e − f i), then e + f i is an associate of either
a + bi or a − bi, so that e + f i is one of a + bi, −(a + bi), i(a + bi), −i(a + bi),
or a − bi, −(a − bi), i(a − bi), −i(a − bi). It follows that {a2 , b2 } = {e2 , f 2 },
which proves uniqueness.
√ √
Application (2). Take K = Q( −2) so that O √ K = Z[ −2]. This is a
UFD (Problem Sheet 2). We have NormK/Q (a + b −2) = a + 2b2 , so that2
18
More theorems of Fermat
(a) := {ra : r ∈ R}
19
Note 6.2 It is easy to check that:
1. IJ is an ideal of R,
Comment. We shall prove later (Theorem 6.26) that any nonzero proper
ideal A of OK can be written as a product of prime ideals A = P1 P2 . . . Pr
and this factorisation is unique up to the order of the factors.
20
Proof. (⇐): Suppose OK is a PID. Then for any nonzero I ⊆ OK , there
exists α ∈ Ok such that I = (α). Then I(1) = OK (α), so I ∼ OK .
(⇒): Suppose hK = 1. Then for all I ✁ OK there exist α, β ∈ OK such
that
I(α) = OK (β). (6.1)
Now the right hand side is just (β). Since β ∈ (β) from Note 6.2 (3), we see
that β = iα for some i ∈ I. Hence β/α ∈ I ⊆ OK . We claim I = (β/α).
Certainly (β/α) ⊆ I. Also, a ∈ I =⇒ aα ∈ I(α) = (β), so aα = rβ, for
some r ∈ OK , giving: a = rβ/α, and so a ∈ (β/α); hence I ⊆ (β/α).
Lemma 6.9. Let I ⊆ OK be a nonzero ideal. Then I ∩ Z 6= {0}.
Proof. Choose any nonzero α ∈ I. Suppose that αd + ad−1 αd−1 + · · · + a0 = 0
(all ai ∈ Z) with a0 6= 0. Then a0 = −α(a1 + · · · + αd−1 ) ∈ I ∩ Z.
Lemma 6.10. Let I ⊆ OK be a nonzero ideal. Then OK /I is a finite ring.
Proof. Choose any nonzero a ∈ I ∩ Z. Then OK ⊇ I ⊇ (a). The map from
OK /(a) to OK /I which takes α + (a) to α + I is well-defined and onto. It
therefore suffices to show that OK /(a) is finite. Let w = {w1 , . . . , wn } be an
integral basis for OK . Then OK /(a) is isomorphic as an additive group to
⊕ni=1 (Z/(a))wi ∼
= (Z/(a))n , where n := [K : Q]. So #OK /(a) = an < ∞.
Definition 6.11. The norm of I is defined as N (I) := #OK /I.
Proposition 6.12. Let σ : K → K be an automorphism. Then I =
σ σ
(α1 , . . . , αn ) and , αnσ ) have the same
√ I = (α1 , . . . √ √ norm.
[So, for example,
in OQ( 7) = Z[ 7], N (3, 1 + 7) = N (3, 1 − 7) .]
√
21
Lemma 6.14 (Hurwitz). Let K be a number field with [K : Q] = n. Then
there exists a positive integer M , depending only on the choice of integral
basis for OK , such that for any γ ∈ K, there exist w ∈ OK and 1 6 t 6 M ,
t ∈ Z with
NormK/Q (tγ − w) < 1.
B := {(x1 , . . . , xn ) ∈ Rn : 0 6 xi < 1} .
Partition B into mn subcubes of side 1/m, and consider the points φ({kγ}),
for 0 6 k 6 mn . There are mn + 1 such points and only mn available
22
subcubes. Hence, by the “Pigeon-hole principle”, there are two points lying
in the same subcube. Suppose these correspond to k = h and l, with h > l.
Letting t = h − l, we have 1 6 t 6 mn = M . It follows that tγ = w + δ
where w := [hγ] − [lγ] ∈ OK and δ := {hγ} − {lγ} with
23
Proof. Let {α1 , . P
. . , αn } be a Z-basis for I. Since I = JI there exist bij ∈ J
n
such that αi = j=1 bij αj . Hence det(bij − δij ) = 0, and expanding this
determinant out, one sees that all terms lie in J, except the product of the
1’s in the identity matrix. Hence 1 ∈ J and so J = (1) = OK .
Lemma 6.17. If I is a nonzero ideal of OK , and w ∈ K with wI ⊆ I, then
w ∈ OK .
Proof. Take M = I in Lemma 3.6.
Lemma 6.18. If I, J are nonzero ideals in OK , and w ∈ OK is such that
(w)I = JI, then (w) = J.
Proof. Choose an arbitrary β ∈ J. Then (w)I ⊇ (β)I, so that {β/w}I ⊆ I.
By Lemma 6.17 we therefore have β/w ∈ OK , and so β ∈ (w). Since β was
arbitrary we deduce that J ⊆ (w), giving that w−1 J is an ideal in OK . We
then have I = (w−1 J)I and so by Lemma 6.16, we obtain w−1 J = OK , so
that J = (w).
Proposition 6.19. For any nonzero ideal I ⊆ OK , there exists k such that
1 6 k 6 hK and I k is principal.
Proof. Among the hK + 1 ideals {I i : 1 6 i 6 hK + 1} some two must be
equivalent. Suppose that I i ∼ I j with j > i. Then (α)I i = (β)I j for some
α, β ∈ OK . Let k = j − i and J = I k . Then (α)I i = (β)I i J ⊆ (β)I i , so that
{α/β}I i ⊆ I i . By Lemma 6.17 we have α/β ∈ OK . Also (α/β)I i = JI i and
so, by Lemma 6.18, (α/β) = J. It follows that J = I k is principal.
Proposition 6.20. The ideal classes form a group CK . It is called the class
group of K and its order is the class number hK .
Proof. Given two ideal classes [I], [J] we define the product [I] · [J] := [IJ].
This is well-defined (easy). The element [OK ] acts as an identity, and asso-
ciativity is easily verified. Thus it remains to show the existence of inverses.
Let [I] be the class of I, and [OK ] = [(1)] the identity. However, given
[I] ∈ CK , if I k is principal, then [I k−1 ] is an inverse of [I].
24
Proof. Let k be such that Ak = (α) is principal. Multiplying by Ak−1 , we
get (α)B = (α)C, and so B = C.
Definition 6.22. Let A, B ⊆ OK be nonzero ideals. We write B|A if there
exists an ideal C ⊆ OK such that A = BC.
Proposition 6.23. Let A, B be nonzero ideals in OK . Then B ⊇ A if and
only if there exists an ideal C such that A = BC, i.e., B|A.
So to contain is to divide!
Proof. Let k > 1 be such that B k = (β) is principal. If B ⊇ A then we
have B k−1 A ⊆ B k = (β). Let C := {1/β}B k−1 A, so that C ⊆ OK is an
ideal. Then BC = B{1/β}B k−1 A = A. Hence B|A. Conversely, if B|A then
A = BC ′ , for some C ′ ; furthermore BC ′ ⊆ B, since B is an ideal. Hence
B ⊇ A.
Lemma 6.24. Let A, B be nonzero ideals, and P a prime ideal of OK such
that P |AB. Then either P |A or P |B.
Proof. Suppose that P |AB and P does not divide A. We must show that
P |B. Now P ⊇ AB but P 6⊇ A, so there exists α ∈ A with α 6∈ P . For any
β ∈ B we will have αβ ∈ P , since P ⊇ AB. However P is a prime ideal,
so if αβ ∈ P one of α or β must belong to P . In our case we conclude that
β ∈ P . Hence P ⊇ B, so that P |B by Proposition 6.23.
Note 6.25 In general, for any ring, every maximal ideal is prime. In the
case of rings OK the converse is true for nonzero ideals. To prove this, note
that if P is a nonzero prime ideal of OK then OK /P is a finite integral
domain. Any finite integral domain is a field, and hence OK /P is a field. It
then follows that P is maximal.
This following key theorem is due to Dedekind — as is most of the theory
of ideals in number fields.
25
Proof. Assume not every ideal A (nonzero and proper) has a prime factori-
sation. Let A be such an ideal with N (A) minimal. There exists a maximal
(hence prime) ideal P1 containing A. So Proposition 6.23 shows that there
is an ideal C with A = P1 C.
If A = C then P1 C = C and P1 = OK , by Lemma 6.16. This is clearly
impossible. Hence A ⊆ C, and by the definition of the norm (Definition
6.11) we have N (A) = N (C)[C : A] > N (C). Hence, by our minimality
assumption for A, one can factor C into prime ideals as C = P2 . . . Pr (or
C = OK and A = P1 ). Therefore A = P1 . . . Pr , a contradiction. Hence every
nonzero proper ideal has a prime factorisation.
Suppose
A = P1 P2 . . . Pr = Q1 Q2 . . . Qs .
Now P1 |Q1 . . . Qs . Let k be minimal such that P1 |Q1 . . . Qk . If k = 1 then
P1 |Q1 . If k > 1 then P1 |(Q1 . . . Qk−1 )Qk , but P1 does not divide Q1 . . . Qk−1 .
Since P1 is prime, we must have P1 |Qk . We therefore have P1 |Qk (so P1 ⊇ Qk )
in either case. Since Qk is maximal this implies that P1 = Qk . Without loss
of generality we take k = 1 and then, by the cancellation lemma 6.21, we
have P2 . . . Pr = Q2 . . . Qs . We may now repeat the process until every Pi
has been shown to equal some Qj .
Note that the prime ideals which occur in the factorisation of A are those
which contain A.
Note also that if u ∈ OK is a unit, then (u) = OK and so (u)I = I for any
ideal I ⊆ R; that is to say, ideals “absorb” units. Thus “unique factorisation
of ideals” is simpler to describe than “unique factorisation of elements”. If
OK is a PID then the theorem implies directly that it is a UFD. However, in
general OK will not be a PID, that is to say, not all ideals will be principal.
Note 6.27
At this point we explain how to multiply ideals in practice. It is a fact,
which we will not prove here, that every ideal can be written with at most 2
generators. We shall write (α, β) for the ideal
(α, β) = {αa + βb : a, b ∈ OK }.
26
clearly contains the four elements αγ, αδ, βγ, βδ, giving
Moreover any term µi νi in the sum above is of the shape (αa + βb)(γc + δd) ∈
(αγ, αδ, βγ, βδ), so that
Xn
(α, β)(γ, δ) = { µi νi : µi ∈ (α, β), νi ∈ (γ, δ)} ⊆ (αγ, αδ, βγ, βδ).
1
On the other hand 11 is the highest common factor of 121 and 22, over Z,
so that one can solve 11 = 121m + 22n over Z. It follows that
√ √
(11) ⊆ (121, 22) ⊆ (121, 33 − 11 −13, 33 + 11 −13, 22).
27
6.5 Multiplicativity of the Norm
Definition 6.28. Let A, B be ideals. We define
A + B := {a + b : a ∈ A, b ∈ B},
OK /(A ∩ B) ∼
= OK /A ⊕ OK /B
28
⇔ P i+1 |(α)P i B ⇔ P |B(α) ⇔ P |(α).
Hence ker θ = P .
It now suffices to show that θ is surjective. However
OK /P ∼
= OK / ker θ ∼
= imθ = P i /P i+1 .
= (#OK /P )e = N (P )e .
Corollary 6.33. If AQ= i Piei , (Pi being distinct nonzero prime ideals),
Q
then we have N (A) = N (Pi )ei .
Proof. Use the corollary above and Lemma 6.30.
From the Unique Factorisation Theorem 6.26 and this last corollary we
deduce:
Proposition 6.34. If A, B are nonzero ideals then N (AB) = N (A)N (B).
Note that if N (I) = p, a rational prime, then I is automatically prime.
The converse is not true, but we shall soon see that every prime ideal P does
have N (P ) = pk for some rational prime p and integer k.
√
Example 6.35 What happens in Z[ −5]? Recall that
√ √
6 = 2 × 3 = [1 − −5] × [1 + −5].
29
In terms of ideals we write this as
√ √
(6) = (2)(3) = (1 − −5)(1 + −5).
√ √ √
= (2, 1 + −5), P2 = (2, 1 − −5), Q1 = (3, 1 + −5) and Q2 =
Let P1 √
(3, 1 − −5) where (α, β) := {rα + sβ : r, s ∈ OK }. Now
demonstrating that we have the same factorisation into ideals, even though
the factorisations into irreducibles are different.
30
Definition 7.1. The integer ei is called the ramification index of Pi . If
ei > 1 we say that Pi is ramified. If some ei > 1 we say that p ramifies in
K. The integer fi is called the degree of Pi .
Note that pfi = #OK /Pi and that OK /Pi is isomorphic to the finite field
with pfi elements.
Theorem 7.2 (Dedekind). Suppose that K = Q(α) with α ∈ OK having
minimal polynomial m(x) ∈ Z[x] of degree n. If p does not divide [OK : Z[α]]
and m̄(x) := m(x) mod p ∈ Fp [x] factorises as
r
Y
m̄(x) = ḡi (x)ei
i=1
Proof. Suppose that p does not divide the index [OK : Z[α]]. Consider the
natural map Z[α] → OK /pOK . An element γ of the kernel must have the
form pβ for β ∈ OK . Since p does not divide the index [OK : Z[α]] we must
have β ∈ Z[α]. The kernel is thus precisely pZ[α] and we get an injection
Z[α]/pZ[α] ֒→ OK /pOK . Indeed this must be an isomorphism of rings since
both sides have order pn . Now consider the ring homomorphism from Z[x]
to Z[α]/pZ[α] taking g(x) to g(α) + pZ[α]. This has kernel
giving
Z[α]/pZ[α] ∼
= Z[x]/(p, m(x)).
Finally consider the homomorphism from Z[x] to Fp [x]/(m̄(x)), sending g(x)
to ḡ(x) + (m̄(x)). The kernel of this map is
31
Thus Z[x]/(p, m(x)) ∼
= Fp [x]/(m̄(x)), and composing our various maps we
obtain
OK /pOK ∼
= Z[α]/pZ[α] ∼
= Z[x]/(p, m(x)) ∼
= Fp [x]/(m̄(x)).
32
for some algebraic integer γ. We now have
√
(p) = P := (p, −5 2 + 5) = (p)
is inert.
Case 2: −5
p
= 1. Then
x2 + 5 ≡ (x − a)(x + a) mod p
√
where a 6≡ −a√mod p. In this case (p) = P1 P2 where P1 = (p, −5 − a)
and P2 = (p, −5√+ a). e.g. √ x2 + 5 ≡ x2 − 1 ≡ (x − 1)(x + 1) mod 3,
so that (3) = (3, −5 − 1)(3, −5 + 1). (Note that for case 2 we have
p ≡ 1, 3, 7, 9 mod 20 by quadratic reciprocity.)
33
8 Minkowski: computation of the class group
8.1 Minkowski’s convex body theorem
Let {v1 , . . . , vn } be any basis for Rn . Let L = { ni=1 ai vi : ai ∈ Z} be
P
n
the
Pnlattice generated by the vi . It is an additive subgroup of R . Let D =
{ i=1 ai vi : ai ∈ [0, 1)}. We call D a fundamental domain for L. Every
v ∈ Rn canP be expressed uniquely as v = u + w with u ∈ L and w ∈ D.
If vi = nj=1 aij ej where {e1 , . . . , en } is the “standard basis” for Rn , then
we define Vol(D) := | det(aij )|; this is sometimes denoted Vol(L). We also
have Vol(D)2 = det(vi · vj ), being the determinant of matrix (aij )(aij )t . One
can easily check that Vol(D) is independent of the choice of Z-basis for the
lattice L.
Lemma 8.1 (Blichfeldt). Let L be a lattice in Rn , and let S be a bounded,
measurable subset of Rn such that Vol(S) > Vol(L). Then there exist x, y ∈ S
with x 6= y and such that x − y ∈ L.
Proof. (Non-examinable)
Let D be a fundamental domain for L. When a ∈ L write S(a) = (S −a)∩
D. Then S is the
Pdisjoint union of the sets S(a)+a as a runs over L. It follows
that Vol(S) = a∈L Vol(S(a)). However Vol(S) > Vol(D) and S(a) ⊆ D.
Thus some S(b) and S(c) with b 6= c must overlap. Let v ∈ S(b) ∩ S(c). Then
x = v + b ∈ S and y = v + c ∈ S, and x − y = b − c ∈ L.
Definition 8.2. We say S ⊆ Rn is convex if
x, y ∈ S, 0 6 λ 6 1 ⇒ λx + (1 − λ)y ∈ S.
We say S is symmetric (about the origin) if
x ∈ S ⇒ −x ∈ S.
Theorem 8.3 (Minkowski’s Convex Body Theorem). Let L be a lattice in
Rn . Let S be a bounded measurable subset of Rn which is convex and sym-
metric. If Vol(S) > 2n Vol(L) then there exists v ∈ L − {0} with v ∈ S.
Proof. (Non-examinable)
We have Vol( 12 S) = 2−n Vol(S) > Vol(L). Thus Blichfeldt’s result tells us
that there exist x, y ∈ 21 S such that x − y ∈ L − {0}. Now 2x ∈ S and, by
symmetry, −2y ∈ S. Using convexity we then find that 12 (2x + (−2y)) ∈ S,
that is to say, x − y ∈ S.
34
Note 8.4 If S is closed, and therefore compact, then it is enough to have
Vol(S) > 2n Vol(L).
Example 8.5 We give another proof that if p ≡ 1 mod 4 then there exist
p = x2 + y 2 .
x, y ∈ Z such that
We know that −1 p
= 1, so there is an s such that s2 ≡ −1 mod p. If
p = x2 + y 2 then x2 + y 2 ≡ 0 mod p and so (x/y)2 ≡ −1 mod p. Hence x ≡
±sy mod p. We will search for a “small” integer solution to x ≡ sy mod p.
Such points form a lattice L in R2 . We have
x ≡ sy mod p ⇔ x = sy + pz, with z ∈ Z ⇔ (x, y) = y(s, 1) + z(p, 0).
Hence {(s, 1), (p, 0}} is a basis for L, and
s p
Vol(L) = det
= p.
1 0
√
Let C be the disc x2 + y 2 < 2p, with radius 2p. The set C is clearly convex
and symmetric about the origin, and
p
Vol(C) = π( 2p)2 = 2πp > 22 p = 22 Vol(L).
Hence by Minkowski’s Theorem there exists a nonzero v ∈ L such that
v ∈ C. Suppose that v = (x, y). Since v ∈ L we have x ≡ sy mod p, and
hence x2 + y 2 ≡ 0 mod p. However v ∈ C implies x2 + y 2 < 2p, so that
x2 + y 2 = 0 or p. Finally, since v 6= 0 we must have x2 + y 2 = p.
35
where ∆p 2
:= ∆2 (K). Thus det(A) 6= 0, and σ(OK ) is a lattice in Rn of
volume |∆2 |/2s . P
If I is an ideal of OK , with basis w = {w1 , . . . , wn } then wi = j cij vj
and
N (I) = [OK : I] = | det(cij )|
by Theorem 1.8. Moreover, ∆2 (w) = det2 (cij )∆2 (v) by Lemma 2.4, and
so ∆2 (w) = N (I)2 ∆2 (v). We can now replace the basis v in the previous
calculation by w, to deduce that
p p p
|∆2 (w)| N (I) |∆2 (v)| N (I) |∆2 |
Vol(σ(I)) = = = .
2s 2s 2s
Lemma 8.7. For t > 0 let
( r s
)
X X
Rt := (x1 , . . . , xr , z1 , . . . , zs ) ∈ Rr × Cs : |xi | + 2 |zi | 6 t .
i=1 i=1
Then
36
By Minkowski’s theorem (compact version), there exists a nonzero α ∈ I
such that σ(α) ∈ Rt . Hence
r
X r+s
X p
|σi (α)| + 2 ℜ(σi (α))2 + ℑ(σi (α))2 6 t.
i=1 i=r+1
Pn
This means that i=1 |σi (α)| 6 t and so
n
1X t
|σi (α)| 6 .
n i=1 n
n
!1/n n
!
Y 1 X t
|σi (α)| 6 |σi (α)| 6 ,
i=1
n i=1
n
t n
giving |NormK/Q (α)| 6 n
= cK N (I).
Theorem 8.9. Any ideal class c ∈ CK contains an ideal J such that N (J) 6
cK , that is to say s
4 n! p 2
N (J) 6 |∆ (K)|.
π nn
Proof. Let I be any ideal in the inverse class c−1 . We now know there
exists a nonzero α ∈ I such that |NormK/Q (α)| 6 cK N (I). Since (α) ⊆ I
we have I|(α), and so there exists an ideal J such that IJ = (α). The
relations I ∈ c−1 and IJ = (α) imply that [J] = c and J ∈ c. Moreover
N (I)N (J) = N (IJ) = |NormK/Q (α)| 6 cK N (I), and so N (J) 6 cK .
37
Proof. Since N (J) > 1 for any ideal J ⊆ OK , we must have
s n
4 n! p 2 4 n! p 2
16 |∆ (K)| 6 |∆ (K)|.
π nn π nn
n n
Let bn := π4 nn! . It will suffice to show that bn > 1 for all n > 2. Now
b2 = π 2 /8 > 1. Moreover
n
bn+1 π 1 π 1 π
= 1+ = 1 + n + . . . > > 1.
bn 4 n 4 n 2
Hence bn > 1 for all n > 2.
Note 9.1 The class group is abelian. Let c be any ideal class. Then
there exists J ∈ c with N (J) 6 cK . Write J as a product of prime ideals,
J = P1 . . . Ps , say. By the multiplicativity of the norm, N (Pi ) 6 cK for each
i. Moreover c = [J] = [P1 . . . Ps ] = [P1 ] . . . [Ps ]. Hence c is in the group
generated by ideal classes of prime ideals of norm at most cK . Thus the class
group itself is generated by classes of prime ideals in OK of norm at most
cK .
In order to find a suitable set of generators we observe that prime ideals
of norm 6 cK are factors of ideals (p) where p ∈ N is prime and p 6 cK .
Using Dedekind’s Theorem 7.2, we can factor all such primes p into prime
ideals, to give a complete set of generators.
To determine the class group it remains to find any relations satisfied by
the classes of these prime ideals. Some such relations can be found from the
prime factorisations of the ideals (p), since these are principal, and others can
be obtained by factoring principal ideals (α) generated by elements α ∈ OK
of small norm.
To show that the set of relations found is complete one needs to show that
appropriate combinations of the generators are not principal. In general this
can be awkward, but for complex quadratic fields one can prove that an ideal
I is non-principal by finding all elements α ∈ OK with NormK/Q (α) = N (I),
and checking whether or not I = (α). If K is complex quadratic there will
only be finitely many possible α with NormK/Q (α) = N (I) to check.
38
√ √
Example 9.2 Let K = Q( −5), so that OK = Z[ −5]. We know from
Proposition 5.8 that OK is not a PID, so that hK > 1. We have n = 2, s =
1, r = 0, and ∆2 (K) = −20. Thus
√
2! 4 √
4 5
cK = 2 20 = < 3.
2 π π
It follows that every ideal class contains an ideal of norm at most 2, and
that CK is generated by classes√ of prime ideals of norm at most 2. However
(2) = P22 where P2 = (2, 1 + −5) with N (P2 ) = 2. Hence [P2 ] generates
CK . Moreover P22 = (2), giving [P2 ]2 = [(2)] = [OK ], which is the identity in
CK . Hence CK is cyclic of order 2, and hK = 2.
√ √
Example 9.3 Next consider K = Q( −6), for which OK = Z[ −6], with
n = 2, r = 0, s = 1 and ∆2 (K) = −24. In this case
√
2! 4 √
4 6
cK = 2 24 = ≈ 3.1.
2 π π
The ideal class group CK is generated by classes of prime ideals P such that
N (P ) 6 cK , which means that N (P ) = 2 or 3. √
Now x2 + 6 ≡ x2 mod 2, and so (2) = P22 where√P2 := (2, −6). Similarly
x2 +6 ≡ x2 mod 3, so that (3) = P32 with P3 := (3, −6). We have N (P2 ) = 2
and N (P3 ) = 3. (Indeed e = 2, f = 1 in both cases.) It follows that CK
is generated by [P2 ] and [P3 ], but we need to see if there are any relations
satisfied by these classes. √
If P2 is principal then P2 = (x + y −6) with x, y ∈ Z. Taking norms
this gives 2 = |x2 + 6y 2 |, which is impossible. Similarly P3 is not principal,
so that [P2 ], [P3 ] 6= [OK ] in CK .
Since P22 = (2) we have√[P2 ]2 = [O√ K ], and similarly
√ [P3 ]2 = [OK ].
We next √ observe that −6 = −6.3 − 2. −6√∈ P2 P3 . We also have
NormK/Q ( −6) = 6, and we therefore deduce that ( −6) = P2 P3 . It follows
that [P2 ][P3 ] = [OK ]. Thus [P3 ] = [P2 ]−1 = [P2 ], and CK must be cyclic of
order 2, generated by [P2 ], and hK = 2.
39
y12 + 6 = 3x31 . Hence 3|y1 , and on writing y1 = 3y2 we obtain 3y22 + 2 =
x31 . However 3y22 + 2 ≡ 2 or 5 mod 9 while x31 ≡ 0, 1 or 8 mod 9. This
contradiction shows that we must have y coprime to 3.
It follows that hcf(y, 6) = 1, and hence that√ hcf(x, 6) =√1.
We now use the ideal factorisation (y + 3 −6)(y − 3 −6) = (x)3 . We
proceed to show that the factors
√ on the left
√ are coprime. √If a prime ideal P
divides
√ both factors then 6 −6 = √ {y + 3 −6} − {y − 3 −6} ∈ P , and so
P |(6 −6) = P23 P33 . (Recall
√ that ( −6) = P2 P3 .) Thus P can only be P2
or P3 . However P |(y + 3 −6) implies P |(x)3 , and on taking norms we find
that N (P )|x6 , which is impossible,
√ since hcf(x,
√ 6) = 1.
It follows that (y + 3 −6) and (y − 3 −6) are coprime as ideals of OK .
By unique factorisation of ideals we have
√
(y + 3 −6) = I 3
for some ideal I. Since I 3 is principal we have [I]3 = [OK ], the identity
in CK . However we know from above that hK = 2 (giving [I]2 = [OK ] by
Lagrange’s Theorem), and so we must have [I] = [OK ]. Thus I is principal,
so that I = (α) for some √ α ∈ OK . √
It follows that (y + 3 −6) = (α)3 = (α3 ), giving y + 3 −6 = uα3 with
u a unit. (Recall√ that if (α) = (β) then α = uβ for some unit u ∈ OK .)
For K = Q( −6) the only units in OK are u = ±1, and for both of these
we have u = u3 . It follows that
√ √
y + 3 −6 = {uα}3 = {a + b −6}3 ,
√
say. Equating the coefficient of −6 on both sides gives 3 = b{3a2 − 6b2 },
and so 1 = b{a2 − 2b2 }. Hence b = −1 and a2 = 1, giving y = a3 − 18b2 a =
a{a2 − 18b2 } = ±17. With these y the only possible x is 7, so that the
complete solution is x = 7, y = ±17.
√ √
Example 9.5 Let K = Q( −163), so that OK = Z[ 12 (1 + −163)] and
2√
cK = 163 ≈ 8.13 < 9.
π
Thus the class group CK is generated by the classes of prime ideals dividing
(2), (3), (5) and (7), so we proceed to factor
√ (2), (3), (5) and (7) in OK .
The minimal polynomial of 21 {1 + −163} is x2 − x + 41. However we
find that x2 − x + 41 ≡ x2 + x + 1 mod 2, which is irreducible. Thus (2) is
inert, so that the only prime ideal above 2 is (2), which is principal.
40
For p = 3, 5 and 7 it is enough to consider the factorisation of
√the polyno-
2
mial x + 163 mod p, since p does not divide the index [OK : Z[ −163]] = 2.
Thus the only relevant prime ideals are all principal; hence CK is trivial
and hK = 1. It follows that OK is a UFD. However, it is not a Euclidean
domain. (For this non-examinable fact see S&T, Theorem 4.18)
Note: it is known that there are only finitely many imaginary quadratic
fields K with hK = 1 (the proof of this is hard!). On the other hand it is
conjectured that OK is a UFD for infinitely many real quadratic fields.
√
Proposition 9.6. The fact that hK = 1 for K = Q( −163) implies that
n2 + n + 41 is prime for 0 6 n 6 39.
Proof. Suppose n2 + n + 41 is not prime for some n < 40. Now n2 + n + 41 <
412 , and so n2 + n + 41 must have a prime factor q < 41.
Now
√ √
2 1 1
q|n + n + 41 = n + 1 + −163 n+ 1 − −163 .
2 2
41
√
• n2 + n + 17 is prime for 0 6 n 6 15 (consider Q( −67)).
√
• n2 + n + 11 is prime for 0 6 n 6 9 (consider Q( −43)).
√
• n2 + n + 5 is prime for 0 6 n 6 3 (consider Q( −19)).
√
• n2 + n + 3 is prime for 0 6 n 6 1 (consider Q( −11)).
2 √
cK = 116 ≈ 6.9 < 7.
π
Thus CK is generated by the classes of prime ideals dividing (2), (3) and (5).
We need to factor (2), (3), (5) in OK , using Theorem 7.2.
√
• x2 + 29 ≡ (x + 1)2 mod 2, so that (2) = P22 where P2 := (2, −29 + 1)
is a prime ideal of norm 2.
• x2 + 29 ≡√x2 − 1 ≡ (x + 1)(x − 1) √
mod 5, so that (5) = P5 P5′ with
P5 := (5, −29 + 1) and P5′ := (5, −29 − 1) being distinct prime
ideals of norm 5.
42
order 3 since there are no solutions to x2 + 29y 2 = ±27. We shall come back
to [P3 ] later. √
Turning to√[P5 ], note that 32 +29×22 = 125, so that N ((3+2 −29)) = 53 .
3 2 ′ ′2 ′3
Hence
√ (3 + 2 −29) must √ be one of P5 , P5 P5 , P5 P5 or P5 . However√ 2+
2 −29 ∈ P5 , giving 3+2 √ −29 ∈
6 P 5 . Hence P5 does not divide (3+2 −29).
′3
It follows
√ that (33 + 2 −29) = P5 , and, taking conjugates, we also have
(3−2 −29) = P5 . Hence [P5 ] has order dividing 3. Since P5 is not principal,
it must have order exactly 3. √ √
Finally we note that 30 = {1 + −29}{1 − −29}. Thus
√ √
(2)(3)(5) = (1 + −29)(1 − −29).
Now (2)(3)(5)√= P22 P3 P3′ P5 P5′ . So, in order to have the correct norm, we
see that (1 ± −29) must be one of P2 P3 P5 , P2 P3′ P5 , P2 P3 P5′ or P2 P3′ P5′ . It
follows that at least one of these products is principal, and so one or other
(and hence both) of [P3 ] and [P3′ ] = [P3 ]−1 is in the group generated by [P2 ]
and [P5 ].
We conclude that CK is an abelian group generated by an element of
order 2 and√an element of order 3. Thus it is cyclic of order 6. (In fact
Norm(2√ ± 5 −29)6 = 729 = √ 36 , and by the argument above we find that
(2 + 5 −29) = P3 and (2 − 5 −29) = P3′ 6 .)
√
Example 9.8 [Paper B9 2005] Let K = Q( −37). Given that hK = 2,
prove there are no integral solutions of the equation y 2 = x3 − 37.
Suppose that x, y ∈ Z are such that y 2 + 37 = x3 . Then as ideals we have
√ √
(y + −37)(y − −37) = (x)3 .
√ √
We claim that (y + −37) and (y − −37)√are coprime ideals. For suppose
that
√ a prime ideal P divides √ both. Then y ± −37 ∈ P , so that the difference
2 −37√∈ P . Hence P |(2 −37), and since P is prime we conclude that P |(2)
or P |( −37). √
Since OK = Z[ −37], we may factor (p) = (2) and (p) = (37) in OK
by using the decomposition of X 2 + 37 modulo p. √We have X 2 + 37 ≡
(X + 1)2 mod 2, giving (2) = P22 , where P2 := (2, 1 + −37) is a prime
√ ideal
2 2
of norm 2. Similarly√X + 37 ≡ X mod 37 and hence (37) = (37, −37)2 =
2
P37 , where P37 := ( −37) is prime of norm 37. √ √
It follows that if P is a common factor of (y√+ −37) and (y − −37)
then P = P2 or P37 . In either case, since P |(y + −37), we have P |(x)3 and
43
taking norms we get 2|x6 or 37|x6 respectively. Hence either 2|x or 37|x, as
appropriate.
Suppose firstly that P = P37 . Then 37|x, and since x3 = y 2 + 37 we
must also have 37|y. Thus 372 divides x3 − y 2 = 37, which is impossible.
Alternatively if P = P2 , so that 2|x, we will have 8|x3 . The equation y 2 +37 =
x3 then implies 2
√ that y + 1 ≡ √ 0 mod 4, which is impossible.
Thus (y + −37) and (y − −37) are coprime ideals as claimed. However
their product is (x)3 , which is a cube. Hence by unique factorisation of ideals,
each of the two factors is a cube. In particular,
√
(y + −37) = I 3
The second equation implies that b = ±1 and 3a2 −37 = ±1. Hence 3a2 = 38
or 36, both of which are impossible.
Hence there are no solutions in integers.
10 The equation x3 + y 3 = z 3
In this section we will establish “Fermat’s Last Theorem” for cubes, that
x3 + y 3 = z 3 has no nontrivial√(x, y, z all nonzero) solutions in Z.
We shall work in K = Q( −3). It is convenient to write
√
ω = (−1 + −3)/2,
44
(i) We have ω 3 = 1. Moreover the set of units of OK is {±1, ±ω, ±ω 2 }.
(ii) The ring OK is a UFD.
√
(iii) The element λ := −3 is prime, with norm 3. Moreover we have
λ = ω(1 − ω) = (−ω 2 )(1 − ω 2 ).
Proof. (i) To find the unit group we note that
NormK/Q (a + bω) = a2 − ab + b2 , a, b ∈ Z.
Thus if NormK/Q (a + bω) = 1 then (2a − b)2 + 3b2 = 4, giving solu-
tions (a, b) = ±(1, 0), ±(0, 1) and ±(1, 1), which produce the six units
specified in the lemma.
(ii) See Problem sheet 2.
(iii) Trivial.
45
Lemma 10.3. If α3 + β 3 = γ 3 with α, β, γ ∈ Z[ω], then λ divides at least
one of α, β or γ.
so that n 6= 1.
Lemma 10.5. Under the conditions of the previous lemma each of the ele-
ments α + β, α + ωβ and α + ω 2 β is divisible by λ. Moreover the quotients
α + β α + ωβ α + ω 2 β
, ,
λ λ λ
are coprime in pairs.
Proof. We have
α + β ≡ α + ωβ ≡ α + ω 2 β mod λ.
46
It follows that all three factors are divisible by λ.
Moreover if δ divides both α + β and α + ωβ then it divides
(α + ωβ) − (α + β) = (ω − 1)β
and also
(α + ωβ) − ω(α + β) = (1 − ω)α.
Hence δ|ω − 1, since α and β are coprime. Similarly if δ divides both α + β
and α + ω 2 β then δ|ω 2 − 1, while if δ divides both α + ωβ and α + ω 2 β then
δ|ω 2 − ω. It follows in all three cases that δ|λ, since ω − 1, ω 2 − 1 and ω 2 − ω
are each associates of λ. The second assertion of the lemma then follows.
α3 + β 3 = µλ3n γ 3
α3 + β 3 = µλ3n γ 3 ,
α + ω2β
3(n−1) 3 α+β α + ωβ
µλ γ =
λ λ λ
with coprime factors on the right, belonging to Z[ω]. Since the factors are
coprime there is one factor, (α + ω j β)/λ say, which is divisible by λ3(n−1) .
Write ν = ω j β; then:
α + ω2ν
3 α+ν α + ων
µγ =
λ3n−2 λ λ
47
We now use the fact that Z[ω] is a UFD. We have three coprime factors
whose product is a unit times a cube, and we deduce that each factor must
be a unit times a cube, say
for appropriate units µ′ and µ′′ . Moreover γ2 and γ3 are coprime, since
(α + ων)/λ and (α + ω 2 ν)/λ were coprime; and λ does not divide γ1 since it
did not divide γ.
After Lemma 10.4 we know that n > 2, so that n − 1 > 1 and
48
Proof. Any such solution must also give a solution in Z[ω]. Remove any
common factor from x, y, z, which means they must be coprime in pairs
(since any common factor of two of x, y, z would also divide the remaining
variable). By Lemma 10.3, at least one of x, y, z must be a multiple of λ,
and indeed only one, since the variables are coprime in pairs. We extract
the largest possible power of λ from this variable, λn say, and use µ = 1
(and replace some of x, y, z with −x, −y, −z, as needed) to put the equation
into the form described in Theorem 10.6, which we have shown to have no
solution.
49