Alexandru Gica - Rational Points On Elliptic Curves PDF
Alexandru Gica - Rational Points On Elliptic Curves PDF
Alexandru Gica - Rational Points On Elliptic Curves PDF
Alexandru Gica1
April 8, 2006
1
Notes, LATEXimplementation and additional comments by Mihai Fulger
2
Contents
1 Lecture I 5
1.1 Projective Geometry . . . . . . . . . . . . . . . . . . . . . . . 5
1.1.1 Equivalent definitions . . . . . . . . . . . . . . . . . . 5
1.1.2 The Geometry of the Projective Plane . . . . . . . . . 6
1.2 The Group Law of a Cubic . . . . . . . . . . . . . . . . . . . 7
1.2.1 The choice of origin . . . . . . . . . . . . . . . . . . . 10
1.3 The Weierstrass Normal Form . . . . . . . . . . . . . . . . . . 10
1.3.1 Explicit formulas for the group operation on C(Q) . . 11
1.3.2 The existence of a Weierstrass Normal Form . . . . . 14
2 Nagell-Lutzs Theorem 21
2.1 Discriminants and Resultants . . . . . . . . . . . . . . . . . . 21
2.2 The Nagell-Lutz Theorems . . . . . . . . . . . . . . . . . . . 25
3 Lecture III 31
3.1 Torsion points on rational elliptic curves . . . . . . . . . . . . 31
3.2 Test Paper . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.1 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 35
4 A Theorem of Gauss 39
5 Mordell-Weils Theorem 43
6 Lecture VI 53
6.1 The Weak Mordell-Weil Theorem . . . . . . . . . . . . . . 53
6.2 Computing the rank of elliptic curves . . . . . . . . . . . . . . 57
7 Lecture VII 63
7.1 Eulers Equation . . . . . . . . . . . . . . . . . . . . . . . . . 63
7.2 Computing the rank of nonsingular elliptic curves . . . . . . . 65
7.2.1 The curves Cp . . . . . . . . . . . . . . . . . . . . . . 67
7.3 Stories and conjectures . . . . . . . . . . . . . . . . . . . . . . 69
7.3.1 Congruent numbers . . . . . . . . . . . . . . . . . . . 69
7.3.2 The Birch and Swinnerton-Dyer Conjecture . . . . . . 70
3
4 CONTENTS
8 Lecture VIII 73
8.1 Algebraic Number Theory Prerequisites . . . . . . . . . . . . 73
8.2 Completing the proof of Mordell-Weils Theorem . . . . . . . 76
8.3 C17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
9 Lecture IX 89
9.1 Test Paper . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
9.1.1 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . 90
Lecture I
PnK = An+1
K \ {(0, 0, . . . , 0)}/
5
6 CHAPTER 1. LECTURE I
Exercise 1.2.5. Let C be the complex cubic defined by the degree 3 homo-
geneous polynomial F . Then C is smooth if and only if F is irreducible.
In 1.1.5 we have defined the hyperplane at infinity of PnK for any field K
by working in the first variable X0 . It is clear that nothing changes if we
work with any other variable. In particular, when working in P2 , we will be
working with points at infinity with respect to the last variable, Z.
The points at infinity of C are given by F (X, Y, 0) = 0 X 3 = 0
X = 0. This means that the only point at infinity of C is O = [0 : 1 : 0]
1.3. THE WEIERSTRASS NORMAL FORM 11
x(2P ) = 2 a 2x1
. (1.3.2)
y(2P ) = 3 + a + 2x1
We have seen that the inverse of P (x, y) C is P 0 (x, y). This also
holds in C(Q).
Remark 1.3.9. These formulas can be used to give complete proofs for
Theorem 1.2.12 and Proposition 1.2.13 without any prior knowledge of Al-
gebraic Geometry, but only if the affine cubic has the particular form y 2 =
x3 + ax2 + bx + c i.e. the cubic is an elliptic curve.
Proof: Let v = 36 xy 1
x+y and u = 12 x+y . Then solving this system in
36+v 3
3
x, y gives x = 36+v and y = 36v + 36v
u u . (x, y) C(Q) 6u 6u =
v 2 = u3 4322 . This proves (C(Q)) C 0 (Q). It iseasy to check
that : C 0 (Q) C(Q) defined by (u, v) = 36+v 36v
6u , 6u is an inverse
for .
Remark 1.3.13. In the conditions of the example above, can be extended
to a group morphism : C(Q) C 0 (Q) by setting (O) = O0 , where
O([1 : 1 : 0]) and O0 ([0 : 1 : 0]) are the only points at infinity of C(Q) and
C 0 (Q) respectively. The proof that is a group morphism is of geometric
z
nature. ([x : y : z]) = (12 x+y , 36 xy
x+y ) is a projective transformation
0
which sends C to C , O to O and lines to lines.
Example 1.3.14. Find the rational solutions of y 2 + 432 = x3 .
Proof: The previous example proves that there is bijective correspon-
dence between the rational solutions of y 2 = x3 432 and the rational
solutions of u3 + v 3 = 1. This is a well known particular case of Fermats
Last Theorem and thus it has no nontrivial solutions. The trivial solutions
are (u, v) {(1, 0), (0, 1)}. The corresponding solutions of y 2 = x3 432 are
(1, 0) and (0, 1). (1, 0) = (12, 36) and (0, 1) = (12, 36).
Proposition 1.3.22. Let l1 and l2 be lines in the projective plane and let
P1 l1 and P2 l2 . There exists a projective transformation sending l1 to
l2 and P1 to P2 .
F F F
R(P, Q) = (P )x(Q) + (P )y(Q) + (P )z(Q),
X Y Z
where x(Q), y(Q), z(Q) are fixed representatives of Q. Based on this we give
the following definition:
not all the partials at (0, 1, 0) vanish and a7 6= 0. Multiplying the coefficients
of the homogeneous equation defining C by a nonzero constant does not
change its solutions therefore we can assume a7 = 1. The equation of C is
now: Y 2 Z + a4 XY Z + a8 Y Z 2 + (a0 X 3 + a2 X 2 Z + a5 XZ 2 + a9 Z 3 ). It has
the form
Nagell-Lutzs Theorem
21
22 CHAPTER 2. NAGELL-LUTZS THEOREM
the (n + m) (n + m) matrix:
an an1 . . . a0 0 . . .
0 an . . . a1 a0 0
0 0 an . . . a1 a0
..
.
Res(f, g) = det
bm bm1
.
. . . b0 0 . . .
0 bm . . . b 1 b0 0
0 0 bm . . . b1 b0
..
.
We now give some formulas for the resultant and for the discriminant of
a polynomial.
F (X1 , . . . , Xn , Y1 , . . . , Ym )(X) = an (X X1 ) . . . (X Xn ),
G(X1 , . . . , Xn , Y1 , . . . , Ym )(X) = bm (X Y1 ) . . . (X Ym ).
Clearly f = F (1 , . . . , n , 1 , . . . , m ) and g = G(1 , . . . , n , 1 , . . . , m ) in
[X1 , . . . , Xn , Y1 , . . . , Ym ].
A = A[X1 , . . . , Xn , Y1 , . . . , Ym ] is an integral domain, so Res(F, G) is a
well defined element of A and it is clear from the definitions that
Let V = ni=1 m
Q Q
j=1 (Xi Yj ) A. By expanding F and G we see that
the variable Xi appears in the matrix of Res(F, G) only on the first n lines
and on these lines it appears in degree 1, hence the degree of Res(F, G) in
Xi is n for all i = 1, n. Similarly we prove that the degree of Res(F, G) in
each Yj is m. From these we conclude that Res(F, G) = V with A.
Notice that = Res(F, G)(1, . . . , 1, 0 . . . , 0) = Res(an (X 1)n , bm X m ).
Computing Res(an (X 1)n , bm X m ) = am n
n bm is a good exercise.
Qn Qm
Corollary 2.1.7. Res(f, g) = am
n i=1 g(i ) = (1)n+m bnm j=1 f (j ).
Proof: Use f 0 (i ) =
Q
j6=i (i j ).
These formulas are often useful for computations and could have just as
well been used to define both the resultant and the discriminant. However
they are not quite useful in proving more theoretical results. For example
dont make clear why Res(f, g) and f are not just elements of but also
of A.
The following is a very important property of the resultant of two poly-
nomials.
m2 u3 1 u2 1 u1
= + a + b t + c.
n2 p2r v 3 p3t v 2 p2t vp
2 3
u 3t 2
If t 0, then nm2 p2r = v 3 p + a uv2 p2t + b uv pt + c, and it is easy to see
that this implies p|v which is a contradiction.
Therefore t 1 and then
m2 1
2 2r
= 3 3t (u3 + au2 pt v + bup2t v 2 + cv 3 p3t ).
n p v p
Proposition 2.2.5. Let C(Q) be the smooth rational elliptic curve given by
y 2 = f (x) with f monic in Z[X]. Then for all (x, y) C(Q), there exist
m, n, e Z, e > 0, gcd(m, e) = gcd(n, e) = 1 such that y = em3 and x = en2 .
28 CHAPTER 2. NAGELL-LUTZS THEOREM
Since vp (ti ) and vp (si ) 3 for i = 1, 2 we easily get that the denomina-
tor of is prime to p, hence vp () = vp (t22 +t1 t2 +t21 +as2 (t2 +t1 )+bs21 ) 2.
ds ds ds ds
= 3t2 + 2ats + at2 + bs2 + 2bts + 3cs2
dt dt dt dt
ds 3t2 + 2ats + bs2
= = .
dt 1 at2 2bts 3cs2
Just as in the case t1 6= t2 , vp () 2.
2.2. THE NAGELL-LUTZ THEOREMS 29
= s1 t1 vp () min{vp (s1 ), vp ( t1 )} 3.
The intersection of P1 P2 and C 0 is given by
.
g(x)f (x) + h(x)f 0 (x)..y y|f .This completes the proof of the Nagell-Lutz
Theorem.
Proof: Let P (x, y) be a torsion point with o(P ) > 2. We have proved
0 (x)2 f 0 (x)2 4f (x)(a+2x)
that x(P ), y(P ) and x(2P ) = f 4y 2 a2x = 4f (x) are integers.
0 2
Let (x) = f (x) 4(a 2x)f (x). Now x(2P ) Z implies 4f |. To prove
the strong version of Nagell-Lutzs Theorem, we use a stronger version of
Remark 2.2.3, but unfortunately we do not prove it.
Remark 2.2.8. In the context above, there exist g, h Z[X] such that
g + hf = f .
Exercise 2.2.11. For an arbitrary prime number p, find the torsion points
of C(Q) : y 2 = x3 + px.
Lecture III
Remark 3.1.2. Just like working over C we can define what it means for
a curve over p to be nonsingular.
Proposition 3.1.4. C(Zp ) has a natural group structure with neutral ele-
ment O = [0 : 1 : 0] in C(Zp ) P2Zp . This group structure is induced by the
group structure of C.
Proof: This can be done similarly to 1.2.12 and 1.2.13. Another way
to prove it is to reduce mod p the algebraic formulas in 1.3.1 and 1.3.2 for
P + Q and 2P respectively. If p divides the denominators of the fractions
in these equations, just set P + Q or 2P as [0 : 1 : 0]. Notice that because
p 6 |2f , p 6= 2, 2P is not always the point at infinity of C(Zp ).
31
32 CHAPTER 3. LECTURE III
Proof: f = 35 = 243.
5 6 |235 so M injects in C(Z5 ). By Lagranges Theorem, |M| | |C(Z5 )|.
Because 3 6 |5 1, the function Z5 Z5 : x x3 is bijective. This means
that over Z25 , the equation y 2 = x3 + 3 has 5 solutions (for each y = 0, 4, x
is uniquely defined). The group C(Z5 ) also contains the point at infinity O,
so |C(Z5 )| = 6.
7 6 | 2 35 so M also injects in C(Z7 ) and |M| | |C(Z7 )|. The solutions
over Z7 of y 2 = x3 + 3 are
(1, 2), (1, 5), (2, 2), (2, 5), (3, 3), (3, 4), (4, 2), (4, 5), (5, 3), (5, 4), (6, 3), (6, 4).
Exercise 3.1.8. Let (tn )n1 be the sequence of rational numbers defined by
t1 = t2 = t3 = t4 = t5 = 1 and
for every n 1.
Prove that all the terms of (tn )n1 are in fact integers.
3.1. TORSION POINTS ON RATIONAL ELLIPTIC CURVES 33
The next page contains the solutions to these problems. If you want to
try to solve them yourself, do not turn the page.
1
Working time 2 hours. This test only counts as extra for the final grade
3.2. TEST PAPER 35
3.2.1 Solutions
Solution to 3.2.1 For y = 0 we get the solutions (2, 0), (0, 0), (2, 0) which
are all elements of order 2 in M. Together with O, M contains at least 4
elements.
f = 44 . 3 6 |2f . The solutions mod 3 of y 2 = x3 4x = x3 x = 0 are
(0, 0), (1, 0), (2, 0). Together with O we find that C(Z3 ) has 4 elements. Like
we did before, we find |M||4. Since M has at least 4 elements, |M| = 4.
M is an 2-torsion abelian group with 4 elements (i.e. 2x = O x M),
therefore M ' Z2 Z2 .
Q. These roots must be integers and we easily get that b must be a square
different from 0. From the condition p4 6 |b for any prime p, b must be the
square of a square free integer.
If M is cyclic of order 4, let P (x, y) be a generator of this group. Notice
that y 6= 0 because otherwise P would be of order 2. This implies x 6= 0.
Because Z4 has only one element of order 2, namely 2, the only element of
M of order 2 is (0, 0). Therefore 2P = (0, 0) P P = (0, 0). We get the
equations:
2 2
x(P P ) = 3x2y+b 2x = 0
y(P P ) = 3x2 +b x(P P ) + (y 3x2 +b x) = 0
2y 2y
3x2 +b 2 2x = 0
2y
3x2 +b
y 2y x =0
( 2
3x2 +b
2x = 0
2y .
2y 2 = 3x3 + bx
Because P C(Q), we have y 2 = x3 + bx. Together with 2y 2 = 3x3 + bx
this gives x3 = bx. Since x 6= 0, x2 = b y 2 = x3 + bx = 2bx.
So we have arrived to solving
x2 = b
y 2 = 2x3 .
y 2 = 2bx
From the condition p4 6 |b for any prime p, x must be a square free integer.
It is easy to see that y 2 = 2x3 x = 2u2 , y = 4u3 for some u Z. Since x
must be square free, u = 1 b = x2 = 4. Hence M is cyclic of order 4 if
and only if b = 4. A generator for M is then P (2, 4).
A Sherlock Holmes type argument gives that M ' Z2 if and only if b
is not a square and b 6= 4. These in the conditions of the problem of course.
Notice that the second equation is a consequence of the first, hence the
2 2
system is equivalent to 3x
2y 2x = x x (3x3 4y 2 ) = 0.
If x = 0, then y 2 = c and so c is a square.
If 3x3 = 4y 2 , then x = 12u2 and y = 36u3 for some u Z. The condition
c = y 2 x3 implies u6 |C u = 1 c = 432.
We are now able to characterize M in the conditions of the problem:
Z6 , if c = 1
Z3 , if (c is a square and c 6= 1) or c = 432
M'
Z2 , if c is a cube and c 6= 1
trivial = O, otherwise
38 CHAPTER 3. LECTURE III
Chapter 4
A Theorem of Gauss
Proof: We quickly rid the case p 2(mod3). In this case, the appli-
cation Zp Zp : x x3 is bijective, hence for each y Zp , there is a
unique x such that x3 + y 3 = 1. Together with the unique point at infinity,
|C(Zp )| = p + 1.
Assume p 1(mod3). Recall that we have identified C(Zp ) to its pro-
jective closure C(Zp ). If C(Zp ) is given by x3 + y 3 = 1, then C(Zp ) is given
via homogenization by x3 + y 3 = z 3 . A substitution z z proves that
|C(Zp )| is the number of solutions in P2Zp of x3 + y 3 + z 3 = 0.
(Zp , ) is a cyclic group generated by an element u. let R = {a3 |a Zp }.
Notice that R = {u3i |i Z} and that it is a subgroup of Zp . Let S = uR and
T = u2 R = uS. Because p 1(mod3), R, S and T are disjoint and actually
form a partition of Zp . To see this, let m = p1 m
3 . In Zp , 1, u and u
2m are
39
40 CHAPTER 4. A THEOREM OF GAUSS
1 )(X P
Let F (X) = (X P 2 )(X 3 ). We set to find its coefficients.
i p1 i
1 + 2 + 3 = iZp = ( i=0 ) 1 = 1.
X p1
X
s+t
2 3 = = N (x)x ,
sS x=0
tT
Let a, b, c N such that |ST R| = ma, |ST S| = mb, |ST T | = mc. Then
|C(Zp )| = 9a and 2 3 = a1 + b2 + c3 .
Similarly, 1 2 = p1 x
P
x=1 N1 (x) , where N1 (x) = |RS{x}|. Also,
|RSR|, if x R
mN1 (x) = |RSS|, if x S ,
|RST |, if x T
1 + 2 + 3 = 3 + 3(1 + 2 + 3 ) = 0.
1 2 + 2 3 + 3 1 = 3 + 6(1 + 2 + 3 ) + 9(1 2 + 2 3 + 3 1 ) =
3 6 9m = 3(3m + 1) = 3p.
1 2 3 = 1 + 3(1 + 2 + 3 ) + 9(1 2 + 2 3 + 3 1 ) + 271 2 3 =
139m+9(km+a) = 29m+9km+3(3a) = 29m+9km+3(m+k) =
(3k 2) + (9km 6m) = (3k 2)(1 + 3m) = p A.
We have proved G(X) = X 3 3pX pA. Lets compute G , the dis-
criminant of G. Recall the following formulas for the discriminant of a
polynomial:
p|AB1 + A1 B
(AB1 BA1 )(AB1 +BA1 ) or . Assume that p|AB1 BA1 .
p|AB1 BA1
The other case is treated in perfect analogy.
(4p)2 = 16p2 = (A21 + 27B12 )(A2 + 27B 2 ) = (AA1 + 27BB1 )2 + 27(AB1
2 2
AB1 BA1
BA1 )2 p|AA1 + 27BB1 16 = AA1 +27BB p
1
+ 27 p . If
2 2
AB1 6= BA1 , then 27 AB1 BA p
1
27. This and AA1 +27BB p
1
0
imply 16 27 which is absurd. So AB1 = BA1 A2 B12 = B 2 A21
(4p 27B 2 )B12 = B 2 (4p 27B12 ) 4pB12 27(BB1 )2 = 4pB 2 27(BB1 )2
4pB12 = 4pB 2 B12 = B 2 4p 27B12 = 4p 27B 2 A21 = A2 A1 =
A. Since A A1 1(mod3), we finally have the proof for A = A1 .
Chapter 5
Mordell-Weils Theorem
In this lecture we begin the proof of the Mordell-Weil Theorem. The theorem
states that the abelian group C(Q) associated to an elliptic curve is finitely
generated. Restricted by an elementary proof we will first prove the theorem
in the particular case it has points of order 2. Later, in Lecture VIII, we
will give a complete proof using some facts of Algebraic Number Theory.
The main steps of the proof of Mordell-Weils Theorem are given by the
following four lemmas. The first three are elementary and will also be used
in the proof of the general Mordell-Weil Theorem in Lecture VIII. To prove
43
44 CHAPTER 5. MORDELL-WEILS THEOREM
(x)
(x) ,
where and are polynomials in x of degrees 4 and 3 respectively. If
and had common roots, then f and f 0 would have had common roots
contradicting the smoothness of C. The coefficients of and are integers
and only depend on C(Q).
To finish the proof of 5.0.16 we need the following lemma:
Lemma 5.0.17. Let , Z[X] be two polynomials with no common com-
plex roots. Let d = max{deg(), deg()}. Then:
1. There exists a nonzero integer R such that gcd(nd ( m d m
n ), n ( n ))|R
for all m, n Z, n 6= 0 and gcd(m, n) = 1.
( m )
2. There exist c2 , c3 R+ such that dh( m
n )c2 h( ( m
n m
) ) dh( n )+c3
n
for rational numbers m m
n such that ( n ) 6= 0.
Assume the above lemma proved. Applied to our setting, it gives the
existence of a constant c02 such that 4h(P ) c02 h(2P ) for every P (x, y)
C(Q) of order different from 1 or 2. Let c2 = max{4h(P ) h(2P )|P
C(Q) is of order 1 or 2}. Choosing c2 = max{c02 , c2 } we have proved 5.0.16.
Proof of 5.0.17: (1) Using that h(q) = h( 1q ) for any nonzero rational
number q, we can assume d = deg() deg() = e. Because and
have no common complex roots, there exist 1 and 1 in Q[X] such that
1 + 1 = 1. There exists a positive integer A such that A1 , A Z[X]
and A depends only on and . Let D = max{deg(1 ), deg(1 )}. Then
m m m m
nd ( ) nD A1 ( ) + nd ( ) nD A1 ( ) = nd+D A.
n n n n
Let k = gcd(nd ( m d m
n ), n ( n )). It is easy to see that the previous equality
implies k|And+D . Let (x) = a0 xd + a1 xd1 + . . . + ad .
k|nd ( m
n ) k|An
d+D1 (nd ( m )) = And+D1 (a md + a md1 n +
n 0 1
. . . + ad nd ). For i > 0, k|And+D k|And+D1+i mdi ai . From this
we conclude k|Aa0 nd+D1 md . Therefore k|gcd(Aa0 nd+D1 md , Ann+D ) =
And+D1 gcd(a0 md , n) = And+D1 gcd(a0 , n)|Aa0 nd+D1 . Repeating this
argument we prove k|Aad+D 0 . A solution to the problem is R = Aa0d+D .
(2) We are only interested in the first inequality i.e. in the existence of
c2 . Proving the existence of c3 is done in analogy to Lemma 2.
nd ( m )
The previous point proves that H( nd ( mn
)
) R1 max{|nd ( m d m
n )|, |n ( n )|}.
n
This is because what is cancelled to make the the denominator and numer-
nd ( m )
ator of nd ( m
n
)
coprime divides R and therefore is less or equal to R.
n
By the well known inequality max{|a|, |b|} 21 (a + b) for positive reals
nd ( m ) 1 m
a, b, we have H( nd ( m
n
)
) d
2R (|n ( n )| + |nd ( m
n )|).
n
varphi( m )
H( n
) |nd ( m d m
|( m m
( m
n
) n )| + |n ( n )| n )| + |( n )|
= k1 > 0
H( mn)
d 2R max{|m|d , |n|d } 2R max{| m d
n | , 1}
46 CHAPTER 5. MORDELL-WEILS THEOREM
For the beginning, assume we have proved Lemma 4 and lets see how the
proof of Mordell-Weils Theorem follows. We will use the following descent
argument:
4. |G : 2G| < .
Proof of 5.0.18: As I have said, we will first prove the Weak Mordell-
Weil Theorem only in the case C(Q) has a point of order 2. The equation
of C(Q) was given by y 2 = x3 + ax2 + bx + c = f (x) with a, b, c Z
and f 6= 0. We have seen that the points P (x, y) of order 2 of C(Q)
are given by y = 0 f (x) = 0. x Q and f (x) = 0 imply that x is
a rational number, integral over Z, therefore x is an integer. Assume x0
is an integer root of f . Up to the substitution x x x0 we can as-
sume that x0 = 0 c = 0. The condition for non-singularity of C is
0 6= f = 4a3 c + a2 b2 + 18abc 4b3 27c2 = a2 b2 4b3 = b2 (a2 4b).
y 2 (x2 b)2 y6 y4 2 y2
= 2a + (a 4b)
x4 x6 x4 x2
If y = 0, the equality is obviously true. If y 6= 0, then the equality is
equivalent to:
2
y0
2 a x 0 = a x = xb . y(P + T ) = x(P + T ) = xby2 .
x0
by 2 by b2
( b) 2 2)
In conclusion (P + T ) = ( xb2 , x2 bx22 ) = ( xy 2 , y(bx
x2
) = (P ).
x x2
We have used b 6= 0 as implied by b2 (a2 4b) = f 6= 0.
It is clear from the definition that (P ) = (P ). We now wish to
prove (P1 ) + (P2 ) = (P1 + P2 ) for all P1 , P2 C(Q). This is equivalent
to (P1 ) + (P2 ) (P1 + P2 ) = O (P1 ) + (P2 ) + (P1 P2 ) = O. Thus
it suffices to prove the following problem: Given P1 , P2 , P3 three collinear
points on C(Q), prove that (P1 ) + (P2 ) + (P3 ) = 0. The collinearity of
P1 , P2 , P3 is equivalent to P1 + P2 + P3 = O. By the preceding remarks it is
enough to consider Pj 6= O, T j = 1, 3. Assume P1 , P2 , P3 are 3 distinct.
Let y = x + be the equation of the line l passing through the three
collinear points P1 , P2 and P3 . We prove that y = x + with
b
=
and
2 a + b2
=
is the equation of a line passing through (P1 ), (P2 ) and (P3 ). Notice that
6= 0 otherwise the line l contains T and by Bezout P1 , P2 or P3 is T which
contradicts our choice for the three collinear points. Let P (x, y) = (P1 ).
It is enough to prove y = x + .
b y12 2 a + b2
x + = + =
x21
(x2 b)f 0 (x) 4xf (x) = (x2 b)(3x2 + 2ax + b) 4x(x3 + ax2 + bx) =
3x4 +2ax3 2bx2 2abxb2 4x4 4ax3 4bx2 = x4 2ax3 6bx2 2abxb2 .
52 CHAPTER 5. MORDELL-WEILS THEOREM
Chapter 6
Lecture VI
The following group theory theorem will provide the finishing argument
in our proof. However matching the conditions in the hypothesis of the
theorem to 6.1.1 will take some time.
53
54 CHAPTER 6. LECTURE VI
1. = 2 1A and = 2 1B .
(B)
B (B)
(A)
Then:
1. is a group homomorphism.
2. ker = Im.
[)x(Q)
(P )(Q)(R) = x(P [ b = x(P
\ )x(Q)b = b b = 1.
We have proved that is a group homomorphism.
(2). ((O)) = (O) = 1 = (O) = ((T )). Let P (x, y) C(Q),
P 6= O, T . If y 6= 0 x 6= 0, then
y 2
x((P )) = ((P )) = 1.
4x2
If y = 0, then P 6= T x 6= 0. Then 0 = y 2 = x3 + ax2 + bx x2 + ax + b =
2 2 4b)
= (2a) 4(a
2
0 (x a2 )2 = a 4
4
b
4 = 4b bb = 4b
b = 1. We have used
that x Q and a2 4b, b 6= 0 (C is nonsingular) to make sure we get elements
of Q.
y = 0 (P ) = T (P ) = b = 1.
We have proved Im ker .
We now prove the reverse inclusion. It is clear that O Im. If
T ker , then b = (T ) = 1 b = d2 for some rational (hence integer)
d. The equation x2 + ax + b = 0 has discriminant a2 4b = 16b = 16d2 ,
hence it has an integer solution x. P (x, 0) is then a point on C(Q) such that
(P ) = T . Therefore
T ker T Im.
Let P (x, y) ker , P 6= O, T . Then x 6= 0 and 1 = (P ) = x x = u2 for
2 integer, u 6=0. We want to prove that P Im i.e.
some rational, hence
y 2 y
(x, y) = (P ) = 4x 2 , (x b) 8x2 for some P C(Q). Let
a y
x1 = 2(u2 + ), y1 = 2ux1 ,
4 u
56 CHAPTER 6. LECTURE VI
a y
x2 = 2(u2 ), y2 = 2ux2 .
4 u
We prove that P1 (x1 , y1 ), P2 (x2 , y2 ) C(Q) and (P1 ) = (P2 ) = P
Im.
2x
2 2 x3 +ax2 + a y 2
We have x1 x2 = 4((x a4 )2 yx ) = 4((x + a2 )2 yx ) = 4 4x
4
=
2
a 4b = b 6= 0 x1 , x2 6= 0.
We want to prove yi = x3i + ax2i + bxi i = 1, 2. Since xi 6= 0, this is
2
equivalent to xyii = xi + a + xbi = a + (xi + x1xxi 2 ) = a + (x1 + x2 ) = 4u2
which holds by the definitionof yi . Therefore P1 , P2 C(Q).
y 2 yi
(Pi ) = 4xi2 , (x2i b) 8x 2 .
i i
yi2 (2uxi )2
= = u2 = x.
4x2i 4x2i
yi (x2i b) yi (x2i x1 x2 ) 2ux1 (x21 x1 x2 ) u(x1 x2 )
2 = 2 = 2 = = y.
8xi 8xi 8x1 4
We have also proved (Pi ) = P (x, y). This proves the reverse inclusion
ker Im and we have ker = Im.
(3) By 2.2.5, if P (x, y) C(Q) and P 6= O, then there exist m, n, e Z,
e 6= 0, gcd(m, e) = gcd(n, e) = 1 such that x = em2 and y = en3 .
Assume first P 6= O, T . Then m 6= 0, otherwise P = T , and
m
c
(P ) = = m.
e2
2 3 2
y 2 = x3 + ax2 + bx ne6 = m e6
+ am e4
+ b em2 n2 = m3 + am2 e2 + bme4 =
m(m + ame + be ). Let d = gcd(m, m2 + ame2 + be4 ) = gcd(m, be4 ) =
2 2 4
gcd(m, b). Since m(m2 + ame2 + be4 ) is a square, there exists m1 Z such
that m = d m21 m = d. Let b = pa1 . . . pat be the prime factor
1 t
decomposition of b with p1 , . . . , pt distinct prime numbers and ai 1 i =
1, t. Then d|b d = p11\ . . . pt t with i {0, 1} i = 1, t. It is easy
to see that these give at most 2t+1 possibilities for d, hence also for (P ).
These possibilities include d {1, b}, therefore we need not consider the
cases P {O, T }.
Now Im ' C(Q)/ ker = C(Q)/Im |C(Q) : Im| 2t+1 .
Remark 6.1.5. Similarly to the lemma above we can prove |C(Q) : Im|
2s+1 , where s is the number of distinct prime factors of b = a2 4b.
We have proved |A : Im| < and |B : Im| < and by 6.1.3 we
have the Weak Mordell-Weil Theorem.
The proof of 6.1.3 actually gives the estimation:
|C(Q) : 2C(Q)| 2s+t+2 .
By what we have seen in the previous lecture, we have finished the proof for
the Mordell-Weil Theorem in the case C(Q) has points of order 2.
6.2. COMPUTING THE RANK OF ELLIPTIC CURVES 57
and |G(2)| = 2 . We have seen that G(2) = {O[0 : 1 : 0]} {P (x, y) G|y =
0}. Since the equation f (x) = 0 has at most 3 integer roots, we find that
|G(2)| 4. G(2) contains at least the elements O and T (0, 0) of G and its
order is a power of 2 least or equal to 4, hence |G(2)| {2, 4}. |G(2)| = 4
if and only if there are 3 points of order 2 i.e. the equation f (x) = 0 has 3
integer solutions i.e. the equation x2 + ax + b = 0 has 2 integer roots. The
equation x2 + ax + b = 0 has 2 integer roots if and only if a2 4b = c2 for
some integer c. Note that b2 (a2 4b) 6= 0 implies that c 6= 0 and that the
roots of f are all distinct.
58 CHAPTER 6. LECTURE VI
|Im| |Im0 |
2r = .
4
We wish to find Im. Let P (x, y) G. By 2.2.5 there exist m, n, e Z
such that x = em2 , y = en3 , e 6= 0 and gcd(m, e) = gcd(n, e) = 1. The case
e = 0 corresponds to P = O.
m 2
= m = b\
1 M = b1 .
c
(P ) =
e2
If M = 0, then P = T (P ) = b. (O) = 1.
Conversely, if there exist M, N, e, b1 , b2 Z such that:
b1 M 4 + aM 2 e2 + b2 e4 = N 2
b1 b2 = b ,
gcd(M, e) = gcd(M, N ) = gcd(e, N ) = gcd(b2 , M ) = gcd(b1 , e) = 1
(6.2.1)
2
then x = b1eM
2 and y = b1 M N
e3
give a point P G and (P ) = b 1 .
Solution: We have seen, for example in 3.2.2, that the torsion group of
C(Q) is isomorphic to Z2 Z2 and as a set is {O, T, (1, 0), (1, 0)}.
0|
We now compute the rank r of C(Q). We know 2r = |Im||Im 4 . We
use 6.2.1 to determine |Im| and |Im0 |.
To find |Im|, we must solve the systems 6.2.1 for all b1 b2 = b = 1
i.e. (b1 , b2 ) {(1, 1), (1, 1)}. The least we can say is |Im| 2. But
{1, 1}
c Im and we get |Im| = 2.
C(Q) : y 2 = x3 + 4x. We apply again 6.2.1 to find |Im0 |. We must
solve systems of type 6.2.1, but the equation is b1 M 4 + b2 e4 = N 2 and
b1 b2 = b = 4. For the cases (b1 , b2 ) {(1, 4), (2, 2), (4, 1)}, the
equation b1 M 4 + b2 e4 = N 2 only has the trivial solution (M, N, e) = (0, 0, 0)
which fails the restriction gcd(M, N ) = 1 in 6.2.1. The remaining cases
b1 = b2 = 2 and (b1 , b2 ) {(1, 4), (4, 1)} are bound to give solutions as they
0|
correspond to 2 or 1 being in Im0 , and 2r = |Im 2 |Im0 | 2. In
conclusion r = 0 and C(Q) is a torsion group isomorphic to Z2 Z2 .
We have proved that the only rational solutions of the equation y 2 =
x3 x are (0, 0), (1, 0).
Example 6.2.3. Our second example is a famous one, Eulers equation.
Find the rational solutions of the equation y 2 = x3 + 1.
In the next lecture we will provide an elementary solution for it. Com-
paring the lengths and depths of the two proofs will show the strength of
Nagell-Lutzs and Mordell-Weils Theorems.
Proof: We will do more than finding the rational solutions. We will
also characterize the abelian group C(Q) : y 2 = x3 + 1. The torsion points
of C(Q) are, as seen for example in 3.2.3, O, (0, 1), (1, 0) and (2, 3).
Because our proof for the Mordell-Weil Theorem and our algorithm for
finding the rank of an elliptic curve are restricted to the assumption T
C(Q), we will first make the substitution x x 1 to find the elliptic curve
that we also denote C(Q) : y 2 = x3 3x2 + 3x. We will prove that the rank
r of C(Q) is 0.
We first find |Im|. We have a = 3 and b = 3. The equations to
solve in the systems 6.2.1, are b1 M 4 3M 2 e2 + b2 e4 = N 2 for b1 b2 = 3.
We have seen that 1, 3 are in Im as the images of O and T respectively.
1
c and 3 c are not as the associated equations only give the trivial solution
(M, N, e) = (0, 0, 0) which contradicts gcd(N, M ) = 1. Therefore |Im| = 2.
For |Im0 | we have a = 6 and b = 3. Again 1, 3
c Im0 . For 3 to be
0 4 2 2 4 2
in Im , the equation 3M +6M e e = N must have a solution satisfying
the conditions in 6.2.1. Reducing mod 3 we find e4 = N 2 (mod3) which
implies 3|e and 3|N , thus contradicting gcd(N, e) = 1. Therefore 3 6 Im0
c 6 Im. We have |Im0 | = 2 and r = 0. It is not hard
and similarly 1
to prove that the rank of the elliptic curve associated to Eulers equation
is also 0 and conclude that the only rational solutions of the equation are
(0, 1), (1, 0) and (2, 3).
6.2. COMPUTING THE RANK OF ELLIPTIC CURVES 61
t = 1 x = 0, y = 1.
t = 3 x = 1, y = 1.
Therefore the only rational solutions to 1 + 2x3 = y 3 are (0, 1) and (1, 1).
(4xy, t) {(0, 1), (0, 1), (1, 0), (2, 3), (2, 3)}.
Modulo 4 we see that these possibilities restrict to (4xy, t) {(0, 1), (0, 1)}.
This means 4xy = 0. y 6= 0, otherwise we get a contradiction modulo 2 in
1 + 4x3 = 0, hence x = 0 and y = 1.
62 CHAPTER 6. LECTURE VI
Chapter 7
Lecture VII
63
64 CHAPTER 7. LECTURE VII
Proof of 7.1.1: The main idea of the proof is Fermats descent ar-
gument. Assume that there exists a solution with the required proper-
ties. Chose one for which b is minimal. We have gcd(b, c2 3bc + 3b2 ) =
gcd(b, c2 ) = 1 and gcd(c, c2 3bc + 3b2 ) = gcd(c, 3b2 ) = 1. Since bc(c2
3bc + 3b2 ) is a square and b, c, c2 3bc + 3b2 are pairwise coprime positive
integers, they must all be squares. Hence there exists k N such that
c2 3bc + 3b2 = k 2 . Let k+c m
b = n with gcd(m, n) = 1, m, n N . Then
m
k = n b c.
2
c2 3bc + 3b2 = ( m 2 2 2 m 2 m
n b c) c 3bc + 3b = n2 b 2 n bc + c
2
2
3(b c)n2 = m2 b 2cmn b(m2 3n2 ) = c(2mn 3n2 ) cb = 2mn3n
m2 3n2
.
m
Lets prove that m2 3n2 > 0 n > 3. This is equivalent to
k+c
b > 3 k > b 3 c. If b 3 c 0 then the inequality is obviously
equivalent to k 2 >
true. Ifb 3 c > 0 then the inequality is (b 3 c)2 =
3b2 2 3bc + c2 c2 3bc + 3b2 > c2 2 3bc + 3b2 3 < 2 3 9 < 12
which holds.
b 2mn3n2 2 2 2 2 2
c = m2 3n2 and m > 3n imply 2mn 3n , m 3n N .
2
Assume 3|m. Then m = 3l cb = 2nln3l2 n2
and gcd(3l, n) = 1. Assume
there exists a prime number p such that p|2ln n2 and p|3l2 n2 . Then
p|n(2l n) and p|(3l2 n2 ) (2ln n2 ) = l(3l 2n). These give four
possibilities:
1. p|gcd(n, l) = 1;
3. p|gcd(2l n, l) = gcd(n, l) = 1;
b = 2nl n2 = b21
. Recall that b, c are squares. So we have c21 +n2 = 3l2 .
c = 3l2 n2 = c21
It is well known that:
Remark 7.1.3. If p is a prime number p 3(mod4) then p|a2 +b2 p|a, p|b
for all a, b Z.
Using this remark, we have 3|c1 and 3|n which contradicts gcd(3l, n) = 1.
Actually, a descent argument proves that the only integer solutions of c21 +
n2 = 3l2 are c1 = n = l = 0. The conclusion is 3 6 |m. As in the case 3|m, we
b 2 = b = 2mn 3n2
prove gcd(2mn 3n2 , m2 3n2 ) = 1. This implies 1 .
c21 = c = m2 3n2
Let pq = c1n+m such that p, q N , gcd(p, q) = 1 and the sign is taken
such that 3 6 | c1 + m. Lets see that this choices are possible. We have
(m c1 )(m + c1 ) = m2 c21 = m2 c = 3n2 > 0 m c1 > 0, so we can
chose positive p, q. Since 3 6 |m, we cannot simultaneously have 3|m + c1 and
3|m c1 .
Lets see that (q, p) is a solution to the problem with q < b. If we prove
this, then we contradict the minimality of b in the choice of (b, c) and we
solve the problem.
We have (n pq m)2 = (c1 )2 = c = m2 3n2 n( pq )2 2m pq =
m p2 +3q 2 b 2mn3n2
3n np2 2mpq + 3nq 2 = 0 n = 2pq . n2 = n2
= 2m
n 3 =
p2 +3q 2 p2 3pq+3q 2 b1 2
pq 3 = pq pq(p 3pq + 3q ) = (pq n ) pq(p 3pq + 3q 2 )
2 2 2
point at infinity of C(Q) and let T = (0, 0). Let C(Q) be the smooth rational
elliptic curve given by y 2 = x3 + ax2 + bx with a = 2a and b = a2 4b.
Let : C(Q) Q /(Q )2 be defined by
x, if P = (x, y) 6= (0, 0), O
(P ) = b, if P = T (0, 0) .
1, if P = O
N 2 = b1 M 4 + aM 2 e2 + b2 e4
The weak form of Birch and Swinnerton-Dyer conjecture you can read
about in the next subsection implies that every positive integer congruent
modulo 8 to 5,6 or 7 is a congruent number.
Theorem 7.3.3 (Tunnel). Let d be an odd, square-free positive integer.
Then d is a congruent number if and only if
for <e(s) > 1, where the product is taken over all the prime numbers, the
following function was introduced:
1
Y ap 1
L(C, s) = 1 s + 2s1 ,
p p
p6| 2f , p is prime
where ap = p+1Np . It can be proved that L(C, s) converges for <e(s) > 32 .
Hasse conjectured that L(C, s) has a holomorphic extension to C. In 1999,
Taylor, Breuil, Conrad and Diamond proved Hasses Conjecture.
A weak form of The Birch and Swinnerton-Dyer Conjecture states:
Using results of Wiles, Coates (1977) and Zagier, Gross (1983), Koly-
vagin proved in 1990 that for modular elliptic curves the following hold:
L(C, 1) 6= 0 rank(C(Q)) = 0
Lecture VIII
73
74 CHAPTER 8. LECTURE VIII
Probably the most famous examples of rings ofintegers are the rings of
integers associated to quadratic extensions Q Q( d) = K with d a square
free integer, d 6= 1. The integral elements of K are:
(
Z[ d], if d 2, 3(mod4)
A= .
Z[ 1+2 d ], if d 1(mod4)
For any two ideals I and J of A, not both 0, there exists the great-
est common divisor gcd(I, J) of I and J defined just like for integers by
min{1 ,1 } min{k ,k }
gcd(I, J) = p1 . . .pk if I = p1 1 . . .pk k and J = p1 1 . . .pk k
with p1 , . . . , pk distinct prime ideals of A and i , i N for all i = 1, k. Since
I I+J and J I+J, I+J|I and I+J|J, hence I+J|gcd(I, J). Conversely
gcd(I, J)|I and gcd(I, J)|J imply gcd(I, J)|I + J, so gcd(I, J) = I + J. In
particular, I and J are coprime if and only if I + J = A. We see that the
ideals of the ring of integers of a number field have similar properties to the
integers in Z.
By the Structure Theorem for Finitely Generated Abelian Groups, U (A) '
Zr W , where W is the torsion part of U (A). W is the set of roots of unity
of K and it can be shown that it is cyclic. We can also determine r. There
exist exactly n distinct field embeddings i : K , C, i = 1, n. Of these
field morphisms, some are real (the image is a subfield of R), and the rest
can be coupled in pairwise complex conjugate homomorphisms. Let s be
the number of real embeddings, and 2t the number of remaining complex
homomorphisms. Then
s + 2t = n
.
s+t1=r
Lets evaluate
the strength of this result by applying it to Pells equation.
Let K = Q( d) with d a square free integer, d 6= 1. Then [K : Q] = 2 and
the two field embeddings of K in C are completely characterized by
1( d) = d
.
2 ( d) = d.
{1, i}, if d = 1
W = 1 3
{1, 2 , if d = 3 .
1,
elsewhere
(A)(B)(C) = ((1 + )2 , (\ 2 2
2 + ) , (3 + ) ) = 1.
(f 0 (1 )2 , (2\
+ )2 , (3 + )2 ) = 1.
Similarly we treat the cases A = (2 , 0) and A = (3 , 0). These prove that
is a group homomorphism.
C(Q)
Lemma 8.2.4. ker = 2C(Q) and as a simple consequence 2C(Q) ' Im
C(Q)
and 2C(Q) = |Im|.
U Q(i )
It is obvious that ker ker i , where i : U2
(Q( 2
i) )
denotes the
U Q3 Q(i )
canonical projection via the isomorphism U2
' i=1 (Q(i ) )2 . The reverse
inclusion now follows from i = ji j .
So it suffices to prove that ker 1 = ker 1
1 1 2C(Q), where i :
U (Q[X]/f Q[X]) Q(i )
U (Q[X]/f Q[X])2
(Q(i ) )2 is the isomorphism obtained canonically from i .
1
Denote = 1 1 . Let P ker . If P = O, then P = 2O 2C(Q). Let
P = (, ) with , Q. By the assumption that f is irreducible over Q,
j 6 Q j = 1, 3, hence C(Q) does not have points of order 2. We want to
find Q C(Q) such that 2Q = P .
We have 1 = (P ) = 1 1 \
1 1 (P ) = 1 1 (( 1 , 2 , 3 )) =
1 \
1 ( 1 ) = \ X ^ X = (1 X 2 +^ 2 X + 3 )2 for some 1 , 2 and
3 in Q.
( X 2 +
1
^ X + )( ^
2 3 X ) = (2 X 3 2 X^
1 2 1 + X )=
2 1 3 2 3
(P ) = ( 1 , \
2 , 3 ).
a\ a\ b1
1 = = a\ b1 = \ u 2 = d u,
b e2
for some u U (A), 0 6= K = Q(1 ) and for some belonging to the
finite set in 8.2.7.
By 8.1.5, U (A) is a finitely generated abelian group, hence there exists
r N such that U (A) is generated by u1 , . . . , ur . It is easy to see that every
element of U (A)/U (A)2 is of the form u11 \ . . . urr with i {0, 1} for
2 r
i = 1, r. So |U (A)/U (A) | 2 . Manifestly, there is only a finite number of
possibilities for ab\ 1 . We treat the second and third coordinates similarly
and keeping in mind that
8.3 C17
We return to 7.2.12 and prove that even though the rank of C17 (Q) : y 2 =
x3 + 17x is 0, one of the equations that we will stumble upon, namely
x2 + 4y 4 = 17e4 , has nontrivial solutions modulo any positive integer m > 1,
but has no nontrivial solution in integers. To simplify the notation, we
will use C(Q) instead of C17 (Q). The procedure is standard. We consider
the associated curve C(Q) : y 2 = x3 68x and the group homomorphisms
: C(Q) Q, 0 : C(Q) Q, defined by
x, if P (x, y) 6= O[0 : 1 : 0], T (0, 0)
(P ) =
17, if P = T ,
1, if P = O
x,
b if P (x, y) 6= O[0 : 1 : 0], T (0, 0)
0
(P ) = 17,
d if P = T ,
1, if P = O
0
where Q = (QQ )2 . It was proved that 2r = |Im||Im
4
|
, where r is the rank of
C(Q). We have proved that for curves of type Cp (Q) with p a prime integer,
|Im| = 2, so all that is left to prove is that |Im0 | = 2.
8.3. C17 81
Since n 3, 2n2 n+1, so 2n+1 |(xn +2n1 t)2 17 if and only if 2|yn +xn t.
Such t exists because 2n |x2n 17 xn is odd. Take xn+1 = xn + 2n1 t.
Alternatively we could have used the following lemma for f (x) = x2 17,
n = 3 and k = 1:
Lemma 8.3.4. Let f Z[X] with f 0 its derivate and let p be a prime
number. Let x Z, n, k Z such that 0 2k < n, pn |f (x) and vp (f 0 (x)) =
k i.e. pk |f 0 (x) and pk+1 6 |f 0 (x).
Then there exists y Z of the form x + pnk z such that pn+1 |f (y) and
vp (f 0 (y)) = k.
Lets see what we have worked so hard for. We have proved that the
equation x2 + 4y 4 17z 4 = 0 has no nontrivial solution in Z even though it
has nontrivial solutions in Zm for any m > 1. This means that generally we
cannot hope to have an algorithm, based on reductions by various integers,
that would help us prove that an equation of the form N 2 = b1 M 4 +aM 2 e2 +
b2 e4 has no nontrivial integer solutions. The reason why such an algorithm
was expected until the discovery of a counterexample like the one above was
that such an algorithm exists for quadratic forms:
Theorem 8.3.5 (Minkowski-Hasse). The equation with nonzero integer
coefficients a1 x21 + . . . + an x2n = 0 has nontrivial solutions ((x1 , . . . , xn ) 6=
(0, . . . , 0)) in Z if and only if it has nontrivial solutions in R and in Zm for
any m > 1.
We should now return to C17 (Q). We have proved that the equations
N 2 = M 4 + 68e4 and N 2 = 4M 4 + 17e4 have no nontrivial integer solu-
c 6 Im0 .
tions. This means that 1
such that a and b have the same parity and (a2 17b2 )2 =
for some a, b, n Z
64m . Since M + 17 e2 > 0, we have
2 2
!2
a + 17 b
M 2 + 17 e2 = (4 + 17)n (3 + 17) .
2
Let (4 + 17)r = Ar + 17 Br with Ar , Br Z uniquely defined by the
recurrence:
Ar+1 = 4Ar + 17Br
, r Z
Br+1 = Ar + 4Br
8.3. C17 85
with initial terms A0 = 1 and B0 = 0. Note the recurrence goes both ways.
Ar+1 and Br+1 are obviously uniquely defined by Ar and Br . Conversely,
Ar and Br are uniquely defined by Ar+1 and Br+1 as a consequence of
4 17
det = 1.
1 4
The equation can be rewritten as:
4 (M 2 + 17 e2 ) = (An + 17 Bn ) (3 + 17) ( + 17 )
for = a2 + 17b2 and = 2ab. So 4(M 2 + 17e2 ) = ((3An + 17Bn ) +
(An 3Bn ) 17) ( + 17 ). We
have the same everywhere it appears
in the preceding equality. Since 17 is not a rational number, we find the
system:
4M 2 = (3An + 17Bn ) + 17(An 3Bn )
2
so a+b
2 3(mod 4) which is impossible. Therefore a and b are both even.
Then there exist u, v Z such that a = 2u, b = 2v and (u2 17v 2 )2 = 4m2 .
This means that u and v have the same parity. Just like before, we get the
system:
We have used that u + v is even because u and v have the same parity, hence
4|(u + v)2 . But 8|M 2 + e2 implies that M and e are even which is false.
We have finished proving that 2 6 Im0 .
We know that Im0 is an abelian group containing 1 and 17. d Also
0 0
1, 2 6 Im . Using these and the group structure of Im , it is easy to
c c
d 6 Im0 . Hence |Im0 | = 2 and the rank r of C17 (Q) is 0.
b 34
prove that 17,
Lemma 8.3.6. Z[ 1+2 17 ] is a principal ideal domain.
1. () = 0 = 0;
Since obviously (x) 0 for all x A, we have proved that (x) N for
all x A. can be extended on Q( 17) by (a + b 17) = |a2 17b2 |.
is multiplicative
on Q( 17) and on A.
2 17b2 | = 0 a2 = 17b2 . It is easy to prove
+ b 17) = 0 2 |a
(a
2
using 17 6 Q that a = 17b a = b = 0. Therefore (x) = 0 x = 0.
Let x, y A such that y 6= 0 and y 6
|x. It is clear that 0 < (xu + yv) <
(y) 0 < ( xy u + v) < 1. xy Q( 17) a, b, c Z, c N such
that gcd(a, b, c) = 1 and xy = a+bc 17 . Because gcd(a, b, c) = 1, there exist
d, e, f Z such that ad + be + cf = 1. Let
u = e + d 17 A.
Then xy u = (a+b 17)(e+d
c
17)
= ae+17db+(ad+be)
c
17
. Let q, r Z be defined
by ae + 17bd = cq + r with 2 r 2c , and let
c
v = q + f 17 A.
x ae+17dbcq+(ad+be+cf ) 17 r+ 17
Then y u+v = c = .
2 c 2
For c 5, we have 0 < ( xy u + v) = r c17 rc
+ 17 14 + 17
25 < 1.
2 c2
If c = 1, then we contradict y 6 |x.
If c = 2, then a and b must have different parities otherwise xy A which
contradicts y 6 |x. Let b 17 and v = q such that a2 17b2 = 2q + 1.
u = a
Then xy u + v = (a+b 17)(ab2
17)2q
= 12 0 < ( xy u + v) = 14 < 1.
If c = 3, then let u = a b 17 and v = q such that a2 17b2 = 3q + r
and r {0, 1, 2}. If r = 0, then 3|a2 17b2 3|a2 + b2 3|gcd(a, b) which
2 2 3q
contradicts gcd(a, b, 3) = 1. Hence r 6= 0. Then xy u + v = a 17b 3 =
r x
3 0 < ( y u + v) < 1.
Assume c = 4. If a and b have different parities, let u = a b 17.
a2 17b2 is an odd number, so there exist q, r Z such that a2 17b2 = 4q+r
and r {1, 3}. Let v = q. Then xy u + v = 4r 0 < ( xy u + v) < 1.
If a and b have the same parity, then they must both be odd because
gcd(a, b, 4) = 1. We have the two cases two consider:
1. If a 3(mod4), then let u = 1 and a = 4k +3.
Let v = k l 17 with
l defined by b = 4l 1. Then xy u+v = 34 17 0 < xy u+v = 12 < 1.
Lemma 8.3.8. The group of units of A = Z[ 1+2 17 ] is
U (A) = {(4 + 17)n |n Z}.
88 CHAPTER 8. LECTURE VIII
Proof: It is very easy to see that (4 + 17)n is a unit for every integer
n.
Let now u = a+ 2 17b be a unit in A. a, b Z and have the same parity.
Since u is invertible, there exists v A such that uv = 1. Since is
2 2
multiplicative, we find (u) (v) = 1 (u) = (v) = 1 a 17b 4 = 1.
2
a 17b2
a+b 17 ab 17
Assume first 1 < u < 4 + 17. 1 = 4 = 2 2
ab 17
2 = a+b217 = u1 < 1 1 < b 17a < 1. Since 1 < a+b 17
<
2 2
4 + 17, we get 0 < b 17 < 5 + 17 0 < b < 1 + 5 1717 < 3. We have the
two cases:
Lecture IX
C(Q) : y 2 = x3 82x.
1
Working time 150 minutes. This test paper only counts as extra for the final exam.
89
90 CHAPTER 9. LECTURE IX
9.1.1 Solutions
Solution to 9.1.1: Let C(Q) : y 2 = x3 + 328x. Let r denote the rank of
the elliptic curve C(Q). We have proved that
|Im| |Im0 |
2r = ,
4
where : C(Q) Q = (QQ )2 and 0 : C(Q) Q are group homomor-
phisms. We have seen that x Im if and only if there exists b1 |82 such
that b1 = x and the equation
82 4
N 2 = b1 M 4 e
b1
has nontrivial solutions i.e. (N, M, e) 6= (0, 0, 0). Similarly, for 0 we have
the equations
328 4
N 2 = b1 M 4 + e .
b1
We first compute Im. The divisors of 82 are {1, 2, 41, 82}.
Therefore Im {1, c 2,
c 41,
d 82}.
d We prove that the previous inclusion
is in fact equality. To do so, we give nontrivial solutions to each of the equa-
tions for b1 {1, 2, 41, 82}. To make computations easier, notice that
if (N, M, e) is a nontrivial solution to the equation N 2 = b1 M 4 82 4
b1 e , then
(N, e, M ) is a nontrivial solution to N 2 = 82 4 82 4 4 82
b1 M 82 e = b1 e b1 M .
4
b1
82
This means that b1 Im
d
b1 Im. We have the solutions:
b1 N M e
1 1 0 1
2 11 1 3 .
41 3 2 1
82 1 1 3
Hence |Im| = 8.
We now compute Im0 . We have b1 < 0 328 b1 < 0. In this case,
2 4 328 4
N = b1 M + b1 e implies N = M = e = 0. Therefore the equations
N 2 = b1 M 4 + 328 4
b1 e give only trivial solutions for b1 < 0.
The positive divisors of 328 are {1, 2, 4, 8, 41, 82, 164, 328}, so Im0
82}.
{1, 2, 41, We prove that this last inclusion is in fact an equality. Just like
for , we need consider only half the cases, because of the pairing b1 , 328 b1 .
b1 b1 N M e
1 1 1 0 1 .
2 8 7 1 1
9.1. TEST PAPER 91
rankC17 (Q) = 0,
An unexpectedly hard
problem
93
94 CHAPTER 10. AN UNEXPECTEDLY HARD PROBLEM
C 0 : y 2 z = x3 + 5x2 z 32xz 2 .
n xn yn
1 1 2
2 2 0
3 14 58
4 29 22
27
49 259
5 25 125
1 x yx
[x : y : z] [ : : z].
4 8
Since (Pn ) C 0 (Q), there exist a0n , b0n , t0n Z such that:
(Pn ) = [a0n t0n : b0n : (t0n )3 ], t0n > 0 and gcd(a0n , t0n ) = gcd(b0n , t0n ) = 1.
0 0
This corresponds to the affine situation (Pn ) = ( (ta0 n)2 , (tb0n)3 ) which follows
n n
from 2.2.5 if we prove that (Pn ) 6= O for all n 1. Assume (Pn ) = O
for some n 1. Then (n P + T ) = O n (8, 24) + (0, 0) = O 2n
(8, 24) + 2 (0, 0) = O in C 0 (Q). In C 0 (Q), we have 2T = 0, so 2n (P ) = O.
If that was so, then (P ) would be a torsion point of C 0 (Q). Since T is also
95
0 0 0 0 0 0
(
(an , bn , tn ) = n ,0 bn 0 a0n t0 n , 2tn ), if an is odd .
(a
an bn an tn 0
4 , 8 , tn , if a0n is even
The only thing left to prove is the recurrence tn+5 tn = tn+4 tn+1 +
tn+3 tn+2 . Lets see that this is the same as proving 10.0.4. The recurrence
is the same, but the initial terms slightly differ. If we look at Pn for n = 1, n,
we get t1 = t2 = 1, t3 = 2, t4 = 3 and t5 = 5. The first five terms in 10.0.4
are t1 = t2 = t3 = t4 = t5 = 1. However t6 = 2, t7 = 3 and t8 = 3. So if
we manage to solve 10.0.5 we also obtain the other sequence, but shifted 3
96 CHAPTER 10. AN UNEXPECTEDLY HARD PROBLEM
places to the right which changes nothing on the recurrence or on the terms
of the sequence being integers.
Consider the projective rational transformation given by
[x : y : z] [2z y 3x : 2z + y 2x : 2z x].
The corresponding affine map is (x, y) = 2y3x2x , 2+y2x
2x .
Let C1 : (xy + z 2 )(5z x y) = 6z 3 . We will prove that (C(Q)) =
C1 (Q). For this, let [x : y : z] C(Q), u = 2z y 3x, v = 2z + y 2x
and t = 2z x. We must prove that (uv + t2 )(5t u v) = 6t3 . uv =
4z 2 10xz (y + 3x)(y 2x). u + v = 4z 5x. (uv + t2 )(5t u v) =
(4z 2 10xz y 2 xy + 6x2 + 4z 2 4xz + x2 )(10z 5x 4z + 5x) =
(y 2 xy+7x2 14xz+8z 2 )6z = 6(y(y+x)z+7x2 z14xz 2 +8z 3 ). Now we
use [x : y : z] C(Q). y(y + x)z = x(x z)(x + 2z) (uv + t2 )(5t u v) =
6(x3 x2 z + 2xz 2 + 7x2 z 14xz 2 + 8z 3 ) = 6(x3 + 6x2 z 12xz 2 + 8z 3 ) =
6(2z x)3 = 6t3 .
1
The inverse of is given by [x : y : z] [2x
+ 2y 4z : 2x 4y + 2z :
1 x+y2 x2y+1
x + y 5z] and the affine map is (x, y) = 2 x+y5 , 2 x+y5 .
induces an isomorphism from C(Q) to C1 (Q). The neutral element of
C1 (Q) is (O) = [1 : 1 : 0] and not O like for C and C 0 .
tn tn+3
Lemma 10.0.6. Let vn = tn+1 tn+2 . Then (Pn+2 ) = (vn , vn+1 ) for every
n 1.
Assume we have proved it. Then notice that (u, v) C1 (Q) (v, u)
C1 (Q) because the affine equation of C1 , (xy+1)(5xy) = 6, is symmetric
in x and y. Actually we can prove that if (u, v) C1 (Q), then (u, v), the
inverse of (u, v) in C1 (Q) is (v, u). If uv 6= 0, then (uv + 1)(5 u v) = 6
v 2 + ( u1 + u 5)v + u+1
u = 0. By Vietes relations, the last equation has the
solution v for fixed u if and only if it has also the solution u+1 uv . The same
argument holds for proving that if u 6= 0, then the affine line x = u cuts
C1 (Q) in at most two points. So, for uv 6=0, we have v, v+1 uv C1 (Q)
(v, u) C1 (Q) (u, v) C1 (Q) u, u+1 uv C1 (Q).
Since tn > 0 for all n 1, it is easy to see that vn > 0 for all n 1.
Assuming the lemma, we have (Pn+2 ) = (vn , vn+1 ) C1 (Q) (vn+1 , vn )
vn+1 +1
C1 (Q). Also (vn+1 , vn+2 ) = (Pn+3 ) C1 (Q) and vn+1 , vn vn+1 C1 (Q).
Sincethe affine line x = vn+1 cuts C1 (Q) in at most the two points (vn+1 , vn )
and vn+1 , vvn+1 +1
n vn+1
, we have vn+2 = vn or vn+2 = vvn+1 +1
n vn+1
.
vn+1 +1 1 tn+2 tn+3 tn tn+3
If vn+2 = vn vn+1 , then 1 + vn+1 = vn vn+2 1 + tn+1 tn+4 = tn+1 tn+2
tn+2 tn+5
tn+1 tn+4 + tn+2 tn+3 = tn tn+5 .
tn+3 tn+4
Assume for a contradiction that there exists an n such that vn = vn+2 .
Then (vn , vn+1 ) = (vn+1 , vn+2 ) (Pn+2 ) = (Pn+3 ) (Pn+2 +
Pn+3 ) = O Pn+2 + Pn+3 = O (n + 2)P + T + (n + 3) + T = O
97
xn+1 < 0 4 6yn + 2x2n + 2xn < 0 x2n + 2xn 2 < 3yn yn (yn +
xn ) + x2n = x3n + 2x2n 2xn > 3xn yn which holds if xn 6= yn . If not, then
xn+1 = 0 which implies yn+1 = 0, so Pn+1 = T . This is impossible since P
and Pm are not torsion points of C(Q) for any m.
b b
Proof: Denote an+2 = a, bn+2 = b and tn+2 = t. Pn+2 C(Q)
(
t3 t3
+ ta2 ) = ta2 ta2 1 ta2 + 2 b(b + at) = a(a t2 )(a + 2t2 ). We have
Pn+4 = Pn+2 + 2 (2, 2).
Notice that if Q(x, y) C(Q), then Q = Q O is the second point
of intersection of C and {X = x}. It is not hard to see that this point is
Q(x, y x).
The tangent to C at P (2, 2) is 2x y 2 = 0. It cuts C again at (1, 0),
so 2P = O (1, 0) = (1, 1).
b
y+1 yn+2 +1 +1 b+t3
t3
The equation of the line Pn+2 (1, 1) is x1 = xn+2 1 = a
1 = att3
=
t2
. The x-coordinates of the intersection of Pn+2 (1, 1) and C(Q) are given
by ((x 1) 1)((x 1) + (x 1)) = x(x 1)(x + 2). Since we already
know that this equation has the solutions 1 and xn+2 , we find xn+4 =
2
b+at
(+1)2 t2 (b+at)2 2 2 t2 a(at2 )(a+2t2 )+abt+a2 t2
att3
xn+2 = a = a(at2 )2
= b +2abt+a
a(at2 )2
= a(at2 )2
=
a2 +at2 2t4 +bt+at2 4t2 (at2 )+t(b+t3 )
(at2 )2
=1+ (at2 )2
.
2 2
Similarly, Pn = Pn+2 2(2, 2) = Pn+2 + (1, 0) and xn = 1 + 3t (at (at )tb
2 )2 .
Since xn = at2n , xn+4 = at2n+4 and gcd(an , tn ) = gcd(an+4 , tn+4 ) = 1, the
n n+4
certain thing to say is tn+4 |a t2 and tn |a t2 .
Let p be a prime number such that p|a t2 . Then p|b(b + at) = a(a
t2 )(a + 2t2 ). If p was to divide both b and b + at, then p|at which together
with p|a t2 yields p|gcd(a, t) which contradicts gcd(a, t) = 1. By Lemma
10.0.7, xn+2 < 0 ta2 < 0. In particular, this implies a 6= t2 . If p|b + at,
then since p|a t2 , also p|b + t3 . Let vp (a t2 ) = k i.e. pk |a t2 and
pk+1 6 |a t2 . Then it is easy to prove that pk |b or pk |b + t3 .
98 CHAPTER 10. AN UNEXPECTEDLY HARD PROBLEM
for all n 1. Following the ideas in Lemma 10.0.8 it can be proved that:
2 2 )6t(b2t3 )
(
xn+3 = 2 + 12t (a2t (a2t 2 )2 2t2 a
18t2 (a2t2 )+6t(b+4t3 ) tn+3 tn+1 = gcd(2t2 a,6)
xn+1 = 2 +
(a2t2 )2
( 12t2 (a2t2 )6t(b2t3 ) .
xn+3 = 2 + (a2t 2 )2
2 3atb+2t 3
tn+3 tn = gcd(2t2 a,6)
2 (at2 )tb
xn = 1 + 3t (at
2 )2
So far we have only been interested in the structure of the rational points
C(Q) of an elliptic curve given by C : y 2 = x3 + ax2 + bx + c. But for ages
number theorists have been interested in Diophantine Equations. To honor
their work, we begin the study of integer points on elliptic curves. The
Diophantine Equation to consider is y 2 = x3 + ax2 + bx + c with a, b, c Z.
In the spirit of the course, we denote the set of integer solutions of this
equation by C(Z).
One of the strongest results in connection to this problem is:
Theorem 11.0.10 (Siegel). If C is nonsingular, then the equation y 2 =
x3 + ax2 + bx + c has a finite number of integer solutions i.e. C(Z) is finite.
Notice that Siegels Theorem does not hold if we drop the assumption
on the non-singularity of C. For example, the equation y 2 = x3 has an
infinite number of integer solutions. They can actually pe parameterized as
(x, y) {(t2 , t3 )| t Z}.
For the time being we leave aside the proof of Siegels Theorem and shall
content ourselves to prove another result concerning equations of a more
particular form. This is Thues Theorem proved by Thue around 1909. We
dedicate to it the next section.
101
102 CHAPTER 11. INTEGER POINTS ON ELLIPTIC CURVES
Liouville(1850) (d) = d
T hue(1909) (d) = d2
+1
Siegel(1921) (d) = 2 d
Gelf on, Dyson(1947) (d) = 2d
Roth(1955) (d) = 2
Roth also proved that the bound (d) = 2 cannot be improved. He was
awarded the Fields Prize for his contribution to this problem.
To solve our problem, we will prove the following particularization of the
previous approximation theorem:
Theorem 11.1.7 (Thue). If b N such that = 3 b 6 Q and C is a
positive real
number,
then there is a finite number of pairs (p, q) Z N
p
such that q < qC3 .
Proof: Let A MM,N (Z) be the matrix whose entries are (aij )i=1,N ,j=1,M .
Let t MN,1 (Z) be the vertical vector whose entries are (Ti )i=1,N . Denote
k t k= max{|Ti | | i = 1, N } and k A k= max{|aij | | i = 1, M , j = 1, N }.
M
We are looking for t such that At = 0 and k t k< 2(4N k A k) N M .
For arbitrary H > 1, let
TH = {t ZN | k t k H}.
A simple count yields |TH | = (2[H] + 1)N and |UH | = (2[N H k A k] + 1)M ,
where [H] denotes the integer part of H, i.e. the greatest integer least or
equal to H.
Since N > M , there exists H > 0 such that (2[H] + 1)N > (2[N H k A k
] + 1)M . For such H we have |TH | > |UH |. Since A sends TH into UH , there
exist t1 6= t2 in TH such that At1 = At2 . Then for t = t1 t2 we have t 6= 0,
At = 0 and k t k 2H.
To finish the proof of the lemma we must prove that we can choose such
M
H with the additional restriction H < (4N k A k) N M . So we need H > 0
M
with (2[H] + 1)N > (2[N H k A k] + 1)M and H < (4N k A k) N M .
We have (2[H] + 1)N > (2H 1)N H N and (2[N H k A k] + 1)M
(2N H k A k +1)M (3N H k A k)M . Thus for having (2[H] + 1)N >
(2[N H k A k] + 1)M , it suffices H N (3HN k A k)M H (3N k A k
M M
) N M . We take H = (3N k A k) N M to finish the proof of the lemma.
Theorem
11.1.9 (Auxiliary Polynomial). Let b N such that =
3
b 6 Q. Let m, n N such that m + 1 > 2n 2n
3 m 3 i.e. 3 m = 3 .
Then thereP exist P, iQ Z[X],Pnot both 0, both of degree at most m + n,
P (X) = m+n i=0 u i X , Q(X) = m+n i
i=0 vi X , such that F
(k) (, ) = 0 for all
k = 0, n 1 and
not 1 k F
where F (X, Y ) = P (X) + Y Q(X) and F (k) (x, y) = k! X k
(x, y).
m+nk
X
j+k
j j j+k
= uk+j x + vk+j x y .
k k
j=0
n n!
, also denoted by Cnk .
kdenotes the binomial coefficient equal to (nk)!k!
When we substitute x = y = in the previous polynomial equality, we get
m+nk
X
(k) j+k j j+k j+1
F (, ) = uk+j + vk+j =
k k
j=0
m+nk+1
X
i+k i+k1
= uk+i + vk+i1 i .
k k
i=0
106 CHAPTER 11. INTEGER POINTS ON ELLIPTIC CURVES
can write
2 m+nk+1
X i + k
X i+k1
F (k) (, ) = ui+l + vi+l1 bj l .
k k
l=0 0=i=3j+l
[ m+nkl+1
3 ]
X 3j + l + k 3j + l + k 1
uk+3j+l + vk+3j+l1 = 0
k k
j=0
n (2m + 2n + 3)
< 12(m + n)
2m n + 2
11.1. THUES THEOREM 107
and
(m + n) n
< 9(m + n)
2m n + 2
to be done.
(m+n)n 9m+9
2mn+2 < 9(m + n) n < 18m 9n + 18 n < 4 , which follows
from n < 3(m+1)
2 < 9(m+1)
5 .
4n
2m+2n+3 +2n+3 n(2m+2n+3)
2mn+2 <
3
4n
n
= 10n+9
n . For 2mn+2 < 12(m + n) it suffices to
3
prove 10n + 9 < 12(m + n) 9 < 12m + 2n. By hypothesis, m 3 and the
conclusion follows.
m+n
X m+n
X
F (x, y) = F (k) (, ) (x )k + (y ) Q(k) () (x )k .
k=n k=0
By induction,
m+n m+n
(t)
X
(k) kt k X
(t) kt k
F (x, y) = F (, )(x) +(y) Q ()(x)
t t
k=n k=t
Pm+n
Recall that F (x, y) = i=0 (ui xi + vi xi y). From this it follows
m+n
X
(k) i ik i ik+1
F (, ) = ui + vi
k k
i=0
Proof: The main ingredient of this proof is what is called the Wronskian
of P and Q, defined as
P (x) Q(x)
W (x) = 0 = P (x)Q0 (x) P 0 (x)Q(x).
P (x) Q0 (x)
ln |as |
T 1 .
ln q1
Pm+n Pm+n P
m+n
W (x) = P Q0 P 0 Q = i
i=0 ui x j=0 j vj x
j1
j=0 j uj x
j1
Pm+n i
i=0 vi x . The leading term of this polynomial expression is, by nota-
tion, the coefficient of the degree s part. This means
X
as = j (ui vj uj vi ).
i+j=s+1
5n
ln |as | (m+n) ln 4+ln 4+18(m+n) ln(16b) ln 4+ln 4+45n ln(16b)
3
5
ln 4 + 45 ln(16b) + 1 n.
3
| {z }
c2
c2 n
c2 indeed only depends on b. Also, T 1 ln q1 .
16 + 46n 47n 54 70 n
x2 x + 1 = y 3
are (x, y) {(0, 1), (1, 1), (19, 7), (18, 7)}.
It is easy to prove that every element of the set given above is a solution
to the equation. However, proving that these are the only solutions is a
very difficult problem and for now we will content ourselves to solving a
particular case of the problem.
Solution: First of all, notice that we can assume p > 3. This is because
the equations k 3 = 3 = 22 2 + 1 and k 3 = 7 = 32 3 + 1 have no solutions
in integers.
By multiplying by 4, the equation can be rewritten as (2p1)2 +3 = 4k 3 .
Assume p 1(mod 3). Then 3| 2p 1 3| 4k 3 3| k. Reducing modulo
9 in the equation we obtain 3 0(mod 9) which is impossible.
Therefore p 1(mod 9). Then p2 p + 1 = k 3 1 k(mod 3). Let
k = 3s + 1. Then k 3 = 9 (3s3 + 3s2 + s) + 1 9| k 3 1 = p2 p 9| p 1.
We have
p1 k3 1 k 1 k2 + k + 1
p = = .
9 9 3 3
Since p > 1 and k 2 +k +1 > 0, it is easy to prove that k > 1. Since p| p p1
9 ,
k1 k2 +k+1
it follows that p| 3 or p| 3 .
112 CHAPTER 11. INTEGER POINTS ON ELLIPTIC CURVES
k1
k1 k1 p1 2 +k+1 k2 +k+1
If p| 3 , then p 3 . Also 9 = 3
k 3 3 p1
p
|{z}
N
3k 2 +3k+3. By combining the two inequalities, we have k1 2
3 3k +3k+4
2 k2 +k+1
0 9k + 8k + 13 which is impossible in integers. Therefore p| 3 .
k2 +k+1 p1 k1 k2 +k+1 p1 k1
Set 3 = pt with t N . Then p 9 = 3 3 9 = 3 t.
k2 +k+1 k1 2 k1 k2 +k+1
1(mod k1
We have 3 = 3 3 + 3 3 + 1, hence 3 3 ).
k1 p1 k1 k2 +k+1 k1
3 | 9 p 1(mod 3 ). 1 3 = pt(mod 3 ) t
k1
1(mod 3 ).
k2 +k+1
Assume t > 1. Since t 1(mod k1 3 ), t 3
k+2
3 = p t p k+23 .
k1 p1 2
k +k+1 (3k2)(k+2) 3k2 +4k4
Since t 3 = 9 , p 3k 2. Then 3 3 = 3
0 2k 2 + 3k 5 = (k 1)(2k + 5) 25 k 1 which contradicts k > 1.
2
Therefore t = 1 which implies k +k+1 3 = p and p 1 = 3(k 1) p =
3k2. Substituting yields k +k+1 = 9k6 k 2 8k+7 = 0 k {1, 7}.
2
The thematic of this lecture is obvious from the title. We have seen some
methods for computing the rank of elliptic curves in a few particular cases.
We will illustrate some examples to prove that computing explicit generators
for the abelian group of an elliptic curve is a much harder problem.
We recommend (re)reading the Algebraic Number Theory Prerequisites
section in the beginning of Lecture VIII.
Example 12.0.3. Find a set of generators for the abelian group C(Q) as-
sociated to the elliptic curve y 2 = x3 13.
2. rank(C(Q)) = 1.
113
114 CHAPTER 12. GENERATORS FOR ELLIPTIC CURVES
units, N (x) = N (y) = 2. Let x = a + b i 13 with a, b Z. Then
N (x) = 2 a2 + 13b2 = 2. But the last equation obviously has no integer
solutions, so we get a contradiction,
therefore
2 is irreducible. We have
2|14 = (1 + i 13)(1 i 13) and 2 6 |1 i 13 in A, hence 2 is not prime.
That A is not factorial is also a consequence of the ideal class group C of A
(see 8.1.2) not being trivial. We will prove that |C| = 2 and that a complete
set of representatives is A (the class of principal ideals of A) and the class
of the maximal ideal P = 2A + (1 + i 13)A.
Assume we have proved these assertionson C. Then the equation y 2 +
3 3 3 3
13 = x is equivalent to (y + i 13)(y i 13) = x (xA) = x A =
(y + i 13)A (y i 13)A. By 8.1.2, every nonzero ideal of A decomposes
uniquely as a product of prime ideals, and, as a consequence, we have proved
the existence of an ideal theoretic greatest common divisorfor two nonzero
ideals. With this in mind, let I = gcd((y + i 13)A, (y i 13)A). Then
I|(y i 13)A (y i 13)A I 2i 13 I I|2i 13A.
2 = (4, 2 + 2i 13, 12 + 2i 13) = (4, 2 +
P = (2, 1 + i 13) P
2i 13, 2i 13) = (4, 2, 2i 13) = 2A. It is not hard to prove that i 13 is
irreducible and prime, hence Q = i 13A is a maximal ideal of A. Therefore
I|P 2 Q.If Q|I, then I Q and since y i 13 it follows
I, 2y Q.
Since i 13 6 | 2 in A, i 13| 2y i 13| y y = i 13 (a + b i 13) for
some a, b Z. We obtain y = 13b. But (13b)2 13 = x3 13| x
132 | (13b)
2 13 which is impossible. So Q 6 | I and I| P 2 . If P | I, then
P 2 | (y+i 13)A(yi 13)A P 2 | (xA)3 2A| (xA)3 (xA)3 2A x
is even. We have y 2 + 13 = x3 and modulo 8 we get y 2 3(mod 8) which is
impossible. Therefore P 6 | I and finally I = A.
The unique decomposition
of the nonzero ideals of A as products of
prime ideals, (y + i 13)A (y i 13)A = (xA) 3 and gcd((y + i 13)A, (y
i 13)A) = A prove that (y + i 13)A = J 3 for some ideal J of A. By going
to classes in C, we get A = J3 = J sincewe assume |C| = 2, thus J is a
principal ideal of 3
A. It follows that y + i 13 is associated to 2a cube2 z of
A. u = a + b i 13 is a unit of A if and only if N (u) = 1 a + 13b = 1.
It is easy to see that the
only solutions are u = 1. Since 1 is also a cube
3
in A, we get that y + i 13 = z for some z A. If z = a + b i 13 for some
a, b Z, then
y = a3 3ab2 13
.
1 = 3a2 b b3 13
The last equation can be rewritten as b(3a2 13b2 ) = 1 b = 1. b = 1
leads to 3a2 13 = 1 which is impossible. Therefore b = 1 and 3a2 13 =
1 a = 2. These lead to the solution (x, y) = (17, 70).
We are left to prove the assertions made on C. To prove them we will
need some more results from Algebraic Number Theory.
115
where K = det(T rK (ei ej ))i,j=1,n , (ei )i is a basis for K over Q such that
A = ni=1 Zei , s + 2t = n and s is the number of real embeddings of K i.e.
field homomorphism K , R. T rK (x) is by definition [K : Q(x)] T r(x),
where T r(x) is the sum of conjugates of x i.e. the sum of all the distinct
complex roots of the irreducible polynomial of x over t Q. p
K is called the discriminant of K over Q. 4 nn!n |K | is called the
Minkowski constant of K.
In our case, we have K = Q(i 13), A = Z[i 13], u = i 13, n = 2,
s = 0, t = 1, e1 = 1, e2 = i 13. With these,
T rK (1) T r K (i 13) 2 0
K = det
= det
= 52.
T rK (i 13) T rK (13) 0 26
and i=1 ei fi = n.
then the proofs of 8.2.4 and 8.2.5 show that is a well defined group homo-
morphism whose kernel is 2C(Q). By the fundamental theorem of isomor-
phism it is enough to determine Im. From 8.2.7 we know that there exist
a finite number of algebraic integers A for whichthere exist u U (A)
and K such that = u 2 as P = , w varies in C(Q) \ {O}
with , Z, > 0 and gcd(, ) = 1. Then (P ) = ud . Define
I(P ) = ( )A + 2 g A,
1 (u) = a + b + c2
2 (u) = a + b + c 2 2 .
3 (u) = a + b 2 + c2
Proof:
By Minkowskis Theorem 12.0.5, for every nonzero ideal M of
A = Z[ 3 13], there exists an ideal I of A such that M I(mod P rA) and
t
N (I) 4 nn!n |K |, where n = [K : Q] = [Q() : Q] = 3, t = 1 is half
p
T rK (1) T rK () T rK (2 )
3 0 0
K = det T rK () T rK (2 ) T rK (3 ) = det 0 0 39 = 27 169.
T rK (2 ) T rK (3 ) T rK (4 ) 0 39 0
Therefore N (I) 4 33!3 27 169 < 20. This means that every class of C
has a representative whose norm is at most 19.
We are now looking for ideals of A of norm least or equal to 19. By
using Theorem 12.0.6, we can find the prime ideals of A of norm at most 19.
We have A = Z[] who is the ring of integers of an algebraic extension of Q
3
of degree 3, K = Q( 13), and f = IrrQ () = X 3 13 so Theorem 12.0.6
3
applies merrily. Remember the notation = 13.
There is only one ideal of norm 1, and this is A.
Since for every ideal I of A, we have N (I) I N (I) A I
I| N (I) A, the ideals of norm 2 can be found among the divisors of 2A.
To decompose 2A as a product of prime ideals, we apply Theorem 12.0.6.
Modulo 2 we have the decomposition in irreducible factors f = X 3 13 =
2
(X 1)(X + X + 1). Therefore 2A = P1 P2 with P1 = 2A + ( 1)A,
N (P1 ) = 2 and P2 = 2A + (2 + + 1)A, N (P2 ) = 4. P1 and P2 are both
maximal ideals of A. Our main goal is to prove that A, P1 and P12 is a
complete set of representatives of C. So far we have P2 P11 (mod P rA).
Modulo 3, f = (X 1)3 , so 3A = P33 with P3 = 3A + ( 1)A and
N (P3 ) = 3. We have N (( 1)A) = |NK ( 1)| = | 1 + 13| = 12. Also
P3 | ( 1)A and P1 | ( 1)A. We conclude ( 1)A = P1 P3 Q for some
prime ideal Q of norm 2. Since P1 is the only such ideal, P12 P3 = ( 1)A
P3 P12 (mod P rA).
As a consequence of 8.1.2, the only ideals of norm 4 are P12 and P2 . For
the same purpose, at first, we will only be looking for the prime ideal of
norm at most 19.
The ideals of norm 5 are divisors of 5A. Modulo 5, f = X 3 13 =
2 2
(X 2)(X + 2X + 4) and X + 2X + 4 is irreducible modulo 5 because it
has no roots modulo 5, or because its discriminant, = 5, is not a square
modulo 5. By 12.0.6, 5A = P4 P5 with P4 = 5A + ( 2)A, N (P4 ) = 5 and
P5 = 5A + (2 + 2 + 4)A, N (P5 ) = 25. P4 and P5 are both maximal. Since
N (P5 ) = 25 > 19, P5 is not of great interest for us. We have N ( 2) =
8 + 13 = 5, so ( 2)| 5 in A, hence P4 = ( 2)A and P4 A(mod P rA).
Since N (( + 3)A) = |N ( + 3)| = |27 + 13| = 40 = 23 5, ( + 3)A = P4 Q,
where Q is an ideal of norm 8. But the only ideals of norm 8 are P13 and
P1 P2 . If Q = P1 P2 = 2A, then 2A| ( + 3)A 2| + 3 which is impossible.
Therefore ( +3)A = P4 P13 A P4 P13 (mod P rA) A P13 (mod P rA).
120 CHAPTER 12. GENERATORS FOR ELLIPTIC CURVES
are now selected as generators for the principal ideals in the set {A, P12 , P14 }.
Only A is principal in this set, so there is only one to choose and we choose
it to be 1. It follows that for every choice of P as above, = u 2
for some u U (A) and K . Then (P ) = \ = u[ 2 = u
{1, 0 }. Assume (P ) = 1. c Then = 2 for some K .
Then N ( ) = N ( ) = N ( )2 . But N ( ) = 3 13 3 = 2
2
Lecture XIII
We have reserved this lecture for the proof of the problem 11.1.2, related
to Thues Theorem 11.1.1. We will present it in an enriched form that will
provide us a rare example of an effective method of determining the solutions
of a diophantine equation.
N (a + b + c2 ) = a3 + b3 d + c3 d3 3abcd
123
124 CHAPTER 13. LECTURE XIII
(x + y)n = a + b U (A)
(x + y2 )n = a + b U (A)
Proof: Because the proofs of the two statements are similar, we will only
prove the first. We assume that d 3 and solve the case d = 2 separately
in the last part of the proof.
Assume there exist x, y Z and n Z with |n| 2 and (x + y)n =
a + b U (A). Then obviously x + y U (A). We can eventually replace
x, y by x, y to assume x3 + dy 3 = N (x + y) = 1. From Lemma 13.0.13,
|a + b| 1 and |x + y| 1. Their absolute value cannot be 1 because of
the assumption x, y Z . (x + y)n = a + b, |x + y| < 1 and |a + b| < 1
imply n > 0, so n 2.
We have 1 = N (x + y) = (x + y) |2 (x + y)|2 , so x + y > 0. Since
a+b = (x+y)n , we get a+b > 0. If |x| 1, then dy 3 = 1x3 |dy 3 | 2.
From the assumption d 3, we obtain y = 0. Since x = x + y U (A),
x = 1 and we contradict |x + y| < 1. Therefore |x| 2.
125
Since we have proved that |x| 2, there exists a prime number p dividing
x.
According to the mod 3 residue of n, we have three cases.
If n 2(mod 3), then
n n2 2 n n5 5 n n n2
x y + x y d + ... + y d 3 = 0.
2 5 n
n2
Reducing modulo x we see that x|y n d 3 . But x3 + dy 3 = 1 implies that x
and dy are coprime, so we obtain |x| = 1 which contradicts |x| 2.
If n 1(mod 3), then we can write the condition Cn = 0 as
n n2 n n5 5 n 2 n2 n4
x y+ x y d + ... + x y d 3 = 0.
n2 n5 2
It is enough to prove p| n2 2 3k 3k 3k
3k (3k+1)(3k+2) x . For this, since p | x , we
prove p3k 6 | (3k + 1)(3k + 2). Assume p3k | (3k + 1)(3k + 2). Then p3k | 3k + 1
or p3k | 3k + 2 because gcd(3k + 1, 3k + 2) = 1. But an easy induction proves
p3k > 3k + 2 > 3k + 1 for every prime p 2 and k 1, so we get our long
sought contradiction and we are done with this case.
126 CHAPTER 13. LECTURE XIII
and that
U (A) = {n0 | n Z}.
Since x + y U (A), there exists n Z such that x + y = n0 . Since
x + y and 0 are positive, the sign is actually + . Since x + y and
0 are strictly smaller than 1, n > 0. From 13.0.15 and 13.0.16, n is odd
and not a multiple of 3.
127
Let 0 = P + Q + R2 . We have 1 = N (0 ) = 1 (0 )2 (0 )3 (0 ) =
0 |2 (0 )|2 > 0 N (0 ) = 1
2 + = 2 2 (0 ) + 3 (0 ) = 2 (P + Q + R 2 2 ) + (P + Q 2 +
R2 ) = P + 2Q R2 A. It is clear that
n1
X
= (1)k (2 (0 ))k (2 (0 ))n1k .
k=0
n1
X n1
X
(1)k ( 2 (0 ))k ( 2 (0 ))n1k = (1)k (2 (0 ))k (2 (0 ))n1k =
k=0 k=0
n1
X n1
X
= (1)n1k (2 (0 ))k (2 (0 ))n1k = (1)k (2 (0 ))k (2 (0 ))n1k ,
k=0 k=0
2 + = P + 2Q R2 U (A),
We have 1 < 1 2 2 2
0 = = (P + Q + R )(P + Q + R ) =
2
P2 2 2 2 2 2
+ Q + dR P Q P R dRQ = (P dRQ) + (dR P Q) + 2
From the equation 13.0.4 we have that there exist a, b Z such that
Q = 2ab and one of the following occurs:
If P = a2 and R = 2b2 , then from 13.0.3, a6 + 8da3 b3 + 8d2 b6 +
12da3 b3 = 1 which yields a contradiction modulo 4. Similarly we prove that
the case P = 2a2 and R = b2 is contradictory.
If P = a2 and R = 2b2 , then a6 + 8da3 b3 8d2 b6 + 12da3 b3 = 1.
Set x = db3 . Then 1 = a6 + 20a3 x 8x2 (3a2 )3 (4x 5a3 )2 = 2.
It is known that the integer solutions of Fermats equation, 2 + x2 = y 3 ,
are (x, y) (5, 3). It follows that a2 = 1 and 4x 5a3 = 5 4x
{0, 10} x = 0 db3 = 0 b = 0 Q = R = 0 and P = 1 which
contradicts 6= 1.
If P = 2a2 and Q = b2 , then 1 = 8a6 + 20da3 b3 + d2 b6 . For x = db3 ,
we have x2 + 20a3 x 8a6 (x + 10a3 )2 108a6 = 1. Let t = x + 10a4 and
s = 3a2 . Then t2 1 = 4s3 t1 t+1 3 t1 t+1
2 2 = s . Since gcd( 2 , 2 ) = 1 we
t1 t+1
have 2 = A and 2 = B for some A, B Z. Then A + 1 = B 3 which
3 3 3
Similar methods can be used for a proof of Lemma 13.0.16. With the
same notations as in the later proof, the equation 13.0.4 is replaced by
P 2 R + P Q2 + dQR2 = 0.
It must be proved that = 1 is the only solution of this equation such that
P 3 + dQ3 + d2 R3 3dP QR = 1 i.e. N () = 1. The proof uses that the only
integer solutions of the equation x3 9y 3 = 1 are (x, y) {(1, 0), (2, 1)}.