Notes On Diophantine Geometry: Rational Points On Curves of Genus Zero
Notes On Diophantine Geometry: Rational Points On Curves of Genus Zero
Notes On Diophantine Geometry: Rational Points On Curves of Genus Zero
June 5, 2008
An algebraic set over a field K is (the solution set of) a system of equations:
f1 = 0
X: .. .
.
fm = 0
with f1 , . . . , fm ∈ K[x1 , . . . , xn ].
Let X be an irreducible algebraic set, i.e., (f1 , . . . , fm ) is a prime ideal.
Next, let R = K[x1 , . . . , xn ]/(f1 , . . . , fm ) and let K(X) denote the field of
fractions of R. Then the dimension of X is the transcendence degree of
K(X) over K:
Tr degK K(X) = dim X.
X is a curve if dim X = 1, that is, if some xi is not in K, and all other xj
are algebraic over K(xi ).
We say that X, Y , irreducible algebraic sets, are birationally isomorphic
if their function fields K(X), K(Y ) are isomorphic as field extensions of K.
1
where, if the singularity is an ordinary singularity with m branches, then
δp = 12 m(m − 1). (So, δp depends on the singularity. There is an algorithm
for computing it in general, but you won’t find it in these notes.)
∂f ∂f
f (x, y) = (x, y) = (x, y) = 0.
∂x ∂y
g = 0 ⇐⇒ d = 1, 2
g = 1 ⇐⇒ d = 3
g > 1 ⇐⇒ d ≥ 4.
2
the curve in exactly one other point (unless it is tangent at P ). The equation
of a line through (1, 0) is given by y = t(x − 1). Substituting this into X for
2 −1
y and solving for x gives: x = tt2 +1 . Then, y = t(x − 1) = t−2t
2 +1 . This is our
parametrization.
Corollary 2. A singular cubic always has lots of points over the ground
field.
Example 6. Another
S example is x2 + y 2 = 0. But, this curve is reducible:
X = {x = iy} {x = −iy}.
3
A conic is an absolutely irreducible curve of degree 2. Given by f (x, y) =
0 where deg f = 2.
4
As a polynomial,
ax2 + by 2 + cz 2 ≡ by 2 + cz 2 mod p
≡ c(z 2 − u2 y 2 ) mod p
So for every odd prime p | abc there are linear forms Lp and L0p in x, y, z with
Also, for p = 2,
By the Chinese Remainder Theorem, there exist linear forms L and L0 such
that
5
By the Pigeonhole Principle there exists distinct such triples (x0 , y0 , z0 )
and (x00 , y00 , z00 ) with
and consequently
By condition (i) we may suppose without a loss of generality that a < 0, b <
0, and c > 0. This implies that |abc| = abc. As a result
and
ax2 + by 2 + cz 2 = 0 or − abc
6
If ax2 + by 2 + cz 2 = 0, then we are done. We thus suppose that
ax2 + by 2 + cz 2 = −abc
a(xz +by)2 +b(yz −ax)2 +c(z 2 +ab)2 = (ax2 +by 2 +c(z 2 +ab))(z 2 +ab) = 0
a12 + by 2 + cz 2 ≡ 0 mod p
0 = ax2 + by 2 + cz 2 ≡ cz 2 mod p
7
with a, b, r ∈ Z and p - ab.
Z Z
lim := {(a1 , a2 , . . .) | ai ∈ , ai+1 ≡ ai mod pi }
← pn Z pi Z
Z Z
⊆ × × ···
pZ p2 Z
Theorem. The following are equivalent.
8
(ii) ∃ solutions to f (x1 , . . . , xn ) = 0 with xi ∈ Zp .
Proof. (i) =⇒ (ii)
(r) (r) (r) (r)
For each r let x1 , . . . , xn ∈ Z with f (x1 , . . . , xn ) ≡ 0 mod pr . Since
(r)
Zp is compact ∃ a convergent subsequence for the xi and the limit satisfies
f (x1 , . . . , xr ) = 0
suitable for generalization.
Theorem 4. A conic defined over Q has a Q-rational point if and only if
it has a Qp -rational point for all p, p prime or p = ∞.
We will see that checking p-adic solubility is easy. Here are some (chrono-
logical) remarks.
|·|:K →R
such that
1. ∀x ∈ K, |x| ≥ 0 and |x| = 0 ⇔ x = 0.
9
| · |p is the p-adic absolute value defined last lecture.
If K/Q is finite extension (i.e. K is a number field), then for each ab-
solute value | · | on Q there is only a finite number of equivalence classes of
absolute values on K that gives | · | when restricted to Q. We denote by
MK the set of equivalence classes of absolute values on K. If v ∈ MK , then,
there exists p=prime,0 or ∞ such that |x|v = |x|p ∀x ∈ Q. We say that v|p
(v divides p).
1 if p 6= pi , ∞
|x|p = p−αi if p = pi
pα1 1 · · · pαr r if p=∞
10
Y
|x|nv v = 1
v
|x|∞ = cdeg(a)−deg(b)
1 1
For instance, | |∞ = .
t c
Using valuation to define an absolute value, we have a way of writing
the product formula eliminating the constant:
a
For x = pr ∈ F(t), p ∈ F[t] monic irreducible and p - ab, we set
b
vp (x) = r and |x|p = c−vp (x) deg p and then
Y Y X
|x|p = 1 ⇔ c−vp (x) deg p = 1 ⇔ vp (x) deg p = 0
Notice that the last part is just expressing the fact that for any rational
function we have # zeros=#poles.
11
Definition. If X is a collection of varieties defined over a global field K,
then we say that X satisfy the Hasse principle if ∀X ∈ X
X(K) 6= ∅ ⇐⇒ X(Kv ) 6= ∅ ∀v ∈ MK
It is known that 3x3 + 4y 3 + 5 = 0 has points in Qp for all p but does not
have a rational point and there is an example with three variables as well.
The case of cubics with 4, 5, 6 and 7 variables is open.
Remark. The proof given below works for any global field.
12
space Vn,d of dimension N = n+d
n . Under this identification, we define
the map φk : Vn,k × Vn,d−k → Vn,d , where 1 ≤ k ≤ d − 1, by (g, h) 7→ gh,
where g and h are polynomials in the same n variables with deg g = k,
deg g = n − k, and gh is the product of g and h as polynomials. By the
formula of multiplication of polynomials in terms of their coefficients, the
image of φk is an algebraic set in the N -dimensional affine space Vn,d , so is
the union U of the image of φk over 1 ≤ k ≤ d − 1. Therefore U is the set of
common zeros (c1 , . . . , cN ) of some F1 , . . . , Fr ∈ Z[y1 , . . . , yN ]. This means
that f factors nontrivially if and only if F1 = . . . = Fr = 0 at f ∈ Vn,d , that
is, the coefficients of f satisfies the polynomial equations F1 = . . . = Fr = 0
in N variables. In fact, the polynomial equations can be written explicitly
in some case as the following exercise shows.
Exercise. Let f (x, y) = a1 x2 + a2 xy + a3 y 2 + a4 x + a5 y + a6 be a polyno-
mial of degree 2 over a field whose characteristic is not 2. Then f factors
nontrivially if and only if
a1 a22 a24
13
them to be reducible, it needs at least two conditions for the coefficients.
If X is smooth, primes p with X mod p smooth is called the primes of good
reduction (associated with X).
Now let us get some geometric intuition about good or bad reductions. As
mentioned in the beginning of semester, in order to see the geometry, we
have to work on algebraically closed fields. Thus let F be an algebraically
closed field and R = F [t], K = F (t). Since F is algebraically closed, the
prime ideals in R are of the form (t − c)R with c ∈ F . Thus the set of
prime ideals in R can be identified with F . In this case, F [t] plays the
role of Z, and F plays the role of the set of primes p ∈ N. Now consider
a polynomial f ∈ R[x1 , . . . , xn ], i.e. the coefficients of f are in F [t]. We
write f as ft to indicate the dependence. For any c ∈ F , the element
f ∈ (R/(t − c)R) [x1 , . . . , xn ] is simply fc ; this corresponds the algebraic
set Vc defined by fc = 0 in the n-dimensional affine space. The primes of
good reduction corresponds the points c ∈ F such that Vc is smooth. In this
context, one can show the set of all c ∈ F such that Vc is not smooth is closed
in the Zariski topology of F , hence is finite. Step 1 is the corresponding
statement when F [t] is replaced by Z.
In general, if we do not assume that F is algebraically closed, the primes of
F [t] corresponds the monic irreducible polynomials, which can be identified
with the Galois orbits of elements in the algebraic closure of F .
Step 2: If f is absolutely irreducible in (Z/pZ)[x1 , . . . , xn ] and p is large
with respect to deg f and n, then the variety defined by f = 0 has a smooth
point in (Z/pZ)n .
To show this, we use the Lang-Weil estimate (which we do not prove):
Theorem 10. Given n ≥ 1, d ≥ 1, there are constants C(n, d), C1 (n, d)
such that for any primes q and any absolute irreducible f ∈ Fq [x1 , . . . , xn ]
with degree d,
3
|#{p ∈ Fnq |f (p) = 0} − q n−1 | ≤ C(n, d)q n− 2 + C1 (n, d)
Hyperplane sections reduce to n = 2 which is due to Hasse-Weil. (Func-
tion field analogues of the Riemann Hypothesis)
Fact: C(2, d) = (d − 1)(d − 2)
To do Step 2, we want a solution c ∈ (Z/pZ)n to f = 0 (mod p) with
∂f
∂xi (c) 6= 0 (mod p) for some i.
Lang-Weil gives that the number of solution to f = 0 (mod p) is about pn−1 .
∂f ∂f
Solutions to f = 0 (mod p), ∂x 1
= 0 (mod p) must satisfy Resxn (f, ∂x1
) = 0,
a equation in x1 , . . . , xn−1 ). The latter one has about pn−2 solutions by
Lang-Weil; and given one of its solution x1 , . . . , xn−1 , there are at most d
14
values for xn with f (x1 , . . . , xn ) = 0. Hence we conclude the number of so-
∂f 5
lutions to f = ∂x1
= 0 (mod p) is at most d(pn−2 +C(n−1, d)pn− 2 +C1 (n−
3
1, d). For p large enough, this is less than pn−1 − C(n, d)q n− 2 − C1 (n, d),
which is a lower bound of the number of solution to f = 0 (mod p). There-
fore, for p large enough, there exists a solution to f = 0 (mod p) which is
∂f
not a solutions to ∂x 1
= 0 (mod p). Step 2 is done.
Under the given assumption, we have established the following conse-
quences:
1. f ∈ (Z/pZ)[x1 , . . . , xn ] is absolutely irreducible for all sufficiently large
primes p.
2. For all sufficiently large primes p, the variety defined by f = 0 has a
smooth point in (Z/pZ)n .
The final step is to turn the smooth point in (Z/pZ)n of the variety de-
fined by f = 0 to a p-adic point.
Remark. Note that the assumption f (a) = 0 (mod p) is equivalent to |f (a)|p <
1; the assumption f 0 (a) 6= 0 (mod p) is equivalent to |f 0 (a)|p = 1; the conclu-
sion asserts that there exists an α ∈ Zp such that |α − a|p < 1 and f (α) = 0.
Therefore Hensel’s Lemma is an analogue to the Newton’s method of finding
zeros of differentiable functions on the real line.
f 00 (an−1 )
f (an−1 + zpn−1 ) = f (an−1 ) + f 0 (an−1 )zpn−1 + (zpn−1 )2 + · · · .
2!
We claim f (an−1 + zpn−1 ) ≡ f (an−1 ) + zpn−1 f 0 (an−1 ) (mod pn ) by showing
15
f (k) (c)
k! ∈ Z for any c ∈ Z and 1 ≤ k ≤ d. Note that
d
f (k) (c) X i(i − 1) · · · (i − k + 1) i−k
= ci c ,
k! k!
i=k
16
Question: How to write negative rational integer as p-adic numbers?
Answer: For example, we have −1 ≡ pn − 1 = (p − 1)(1 + p + p2 + . . . +
pn−1 ) mod pn . Thus as a p-adic number, −1 = (p − 1)(1 + p + p2 + . . .).
Alternatively, using a bomb to kill an ant, we may apply Hensel’s Lemma
to the polynomial x + 1 in order to write −1 a p-adic number.
∂f
To turn a solution (a1 , . . . , an ) ∈ (Z/pZ)n of f = 0 and ∂xi 6= 0 for some
i into a p-adic one, simply apply Hensel’s Lemma to
#Bn,d,H
lim >0 if n ≥ 2, d ≥ 2 and (n, d) 6= (2, 2).
H→∞ #An,d,H
17
degree d in n variables. So there are at least 2 numbers that need to be
divisible by p in order for a polynomial in Z[x1 , . . . , xn ] to be reducible in
(Z/pZ)[x1 , . . . , xn ]. Thus the probability of a polynomial in Z[x1 , . . . , xn ]
being bad modulo p is no bigger than p12 . Equivalently, the probability of
a polynomial in Z[x1 , . . . , xn ] being good modulo p is no less than 1 − p12 .
Hence the probability of a polynomial in Z[xQ 1 , . . . , xn ] being good, i.e., good
modulo any primes p ∈ N, is no bigger than p (1 − p12 ) = π62 > 0. Of course
this is a heuristic argument.
In Step 2, given fixed n and d, all but finitely many primes p ∈ N are
good for all polynomials in Z[x1 , . . . , xn ] with total degree d. For each q of
finitely many good primes, the probability of a polynomial in Z[x1 , . . . , xn ]
having a smooth zero is at least
∂f 1 1
Prob f (0, . . . , 0) = 0 mod q, (0, . . . , 0) 6= 0 mod q = (1 − ) > 0.
∂x1 q q
Conjecture 13. Let An,d,H be as in Theorem 11, and Bn,d,H0 be the set of
n
polynomials in An,d,H which admits a zero in Q . Then
0
#Bn,d,H Cn,d > 0 if d ≤ n
lim = ? if d = n + 1
H→∞ #An,d,H
0 if d ≥ n + 2.
18
#{f : |coeff(f )| ≤ H, f (a) = 0} = c(a)H N + O(H N −1 )
Using this lemma we see that:
X X
#{f : |coeff(f )| ≤ H, f (a) = 0} = c(a)H N + O(H N −1 )
a∈A a∈A
for A finite. If the error terms didn’t matter this would give us:
X
#{f : |coeff(f )| ≤ H, ∃a ∈ A, f (a) = 0} ≤ c(a)H N
a∈A
and
#{f : |coeff(f )| ≤ H} ≈ cH N +1
so:
N
P P
#f with solutions a c(a)H a c(a)
≤ =
#total cH N +1 cH
which goes to 0 as H → ∞.
Let K be a field. We will define n-dimensional projective space over K:
19
f (a) = 0 ⇔ f (b) = 0
Thus the set of points in Pn (K) where f (homogeneous) vanishes is well
defined. Thus we can make the following definition:
f (x, y) = 0
so
x y
f = zdf ( , )
z z
and f = 0 defines a curve in P2 . The solutions to f (x, y, 0) = 0 are
points at infinity.
y 2 = x3 + ax + b
y 2 z = x3 + axz 2 + bz 3
so the points at infinity correspond to z = 0 which implies x3 = 0 or
x = 0. So (0 : y : 0) ∼ (0 : 1 : 0) is the point at infinity.
20
Theorem 16. Let P, Q ∈ P2 (K) P 6= Q then there exists a unique line L
that contains P and Q. Given L1 and L2 lines in P2 and L1 6= L2 then there
exists a unique point P ∈ L1 ∩ L2 .
1 Elliptic curves
Definition: If K is a field then an elliptic curve E over K is a smooth
irreducible projective curve of genus 1 with a point P ∈ E(K)
y 2 = x3 + ax + b
21
Group law on E:
3x21 + a
α := .
2y1
x3 := α2 − x1 − x2
y3 := −(α(x3 − x1 ) + y1 )
(α(x − x1 ) + y1 )2 = x3 + ax + b
which becomes
x3 − α2 x2 + · · · = 0
since x1 , x2 , x3 are the roots we get x1 + x2 + x3 = α2 and so forth.
One can prove associativity of the group law either by brute force com-
putation or by using methods of classical geometry.
Two elliptic curves are isomorphic over an algebraically closed field if they
have the same j-invariant. In other words, E is ismorphic to E 0 iff j(E) =
j(E 0 ), where
1728a3
j(E)= 3 ,
4a + 27b2
22
so for example, changing x to λ−2 x and y to λ−3 y gives us y 2 = x3 + λ4 ax +
λ6 b which is isomorphic to our original curve.
Aside Question: Are there other varieties with group laws, or non-abelian
2
ones? Yes, known as group varieties. For example GLn ⊆ An , the set of
n × n matrices with nonzero determinant.
In general, the analogue of the Mordell-Weil does not hold. For example,
if X denotes the affine conic xy = 1 as above, X(Q) = Q× is not finitely
generated. But see the remark below.
Exercise: If y 2 = x3 + x2 (a nodal cubic) then C0 (K) ∼
= K × , and if y 2 = x3
∼
(a cuspidal cubic) then C0 (K) = (K, +), where C0 := C − {(0, 0)}. In other
words, the same formulas define a group law for the given curves (with the
groups mentioned) after we remove the singular point.
23
variety. For those, the integer points form a finitely generated group. This
generalizes both the Mordell-Weil Theorem and the Dirichlet Unit Theorem.
m2 h(R) + c0 ≤ 2h(Q) + c
2 c − c0
h(R) ≤ h(Q) +
m2 m2
so letting c00 = (c − c0 )/m2 we have
2
h(Q) + c00 ≥ h(R)
m2
24
so, choosing λ, 2/m2 < λ < 1, either h(R) < λh(Q) or
2
h(Q) λ − 2 ≤ c00 .
m
Therefore we have h(R) < λh(Q) for some λ < 1, or h(Q) ≤ c000 where
c000 = c00 /(λ − 2/m2 ). Note that the restriction h(Q) ≤ c000 gives us a finite
set.
Start with an arbitrary Q and put Q0 = R and repeat until we have
h(Q(k) ) ≤ c000 , that is,
Q = mQ0 + Pi
Q0 = mQ00 + Pi0
..
.
Q(k)
where h(Q(k) ) ≤ c000 which will happen eventually since otherwise h(Q(i) ) ≤
λi h(Q) which goes to 0 as i grows.
Since Q = mQ0 + Pi = m(mQ00 + Pi0 ) + Pi = m2 Q00 + mPi0 + Pi , . . . we
have
Xr
Q = mk Q(k) + ai Pi
i=1
therefore {P1 , . . . , Pr } ∪ {P ; h(P ) ≤ c000 }, a finite set, generate A(K).
25
Since 0 + 0 = 0, we have Fi (0, . . . , 0, 0, . . . , 0) = 0 for all 1 ≤ i ≤ n; since t +
0 = 0+t = t, we have Fi (t1 , . . . , tn , 0, . . . , 0) = ti and Fi (0, . . . , 0, t01 , . . . , t0n ) =
t0i , for all 1 ≤ i ≤ n. It follows that Fi (t1 , . . . , tn , t01 , . . . , t0n ) = ti + t0i +
higher order terms (as a power series) and hence, m(t1 , . . . , tn ) = (t1 , . . . , tn )+
· · ·+(t1 , . . . , tn ) = (mt1 +higher order terms, . . . , mtn +higher order terms).
By assumption, charK - m, so m(t1 , . . . , tn ) 6= 0 for (t1 , . . . , tn ) close but
unequal to 0, which implies 0 is an isolated point in A[m].
Combine the two results, we see each irreducible component X must be
a single point, so A[m] is a finite set of points.
26
Secondly, for any σ, τ ∈ Gal(L/K), we have λ(στ ) = στ Q − Q = στ Q −
σQ+σQ−Q = σ(τ Q−Q)+(σQ−Q) = σ(λ(τ ))+λ(σ). As λ(τ ) is an element
in A[m] ⊆ A(K), it is fixed by σ ∈ Gal(L/K), and so λ(στ ) = λ(τ ) + λ(σ).
Besides, suppose λ(σ) = 0, then σ(Q) = Q =⇒ L = K(Q) is fixed by σ =⇒
σ = 1. This shows λ is a group monomorphism.
Our goal today is prove the Weak Mordell-Weil Theorem. Recall our
map δ:
δ : A(K) −→ Hom(Gal(Ksep /K), A[m]) = H
We can also identify H with the (a priori) set
(in fact, this set has a group structure), and we will often go back and forth
between these two interpretations of H. We will also assume A[m] ⊆ A(K).
We defined δ by the following: given P ∈ A(K), let Q be a point on our
abelian variety such that mQ = P . Then we set λ(σ) = σQ − Q and
L = K(Q), where K(Q) is the extension of K gotten by adjoining the
coordinates of Q. We first want to prove some claims from last time.
(λ1 +λ2 )(σ) = σQ1 +σQ2 −Q1 −Q2 = σ(Q1 +Q2 )−(Q1 +Q2 ) = σQ−Q = λ(σ),
so λ = λ1 + λ2 as desired.
Claim. ker δ = mA(K).
27
Proof. We first show that mA(K) ⊆ ker δ: say P ∈ mA(K), then there
exists a Q ∈ A(K) satisfying mQ = P , so let us choose this Q. Then L = K
so λ is the zero map, so P ∈ ker δ.
Now assume P ∈ ker δ, so λ(σ) = O for all σ. Then σQ = Q for all σ,
hence Q ∈ A(K) and, since mQ = P , we see that P ∈ mA(K).
Using this last claim, we see that δ : A(K)/mA(K) ,→ H.
Example. Let A = E : y 2 = x3 + ax + b be an elliptic curve and m = 2.
We are assuming that E[2] ⊆ E(K), so the points of order 2 (i.e. those
points with y-coordinate 0) are in E(K). Thus our cubic factors over K:
y 2 = (x − e1 )(x − e2 )(x − e3 ), ei ∈ K.
nn L PPPPP
nnnnn PPP
nnn PPP
n PPP
√ nnn √ √
F1 = K( d1 ) K( d1 d2 ) F2 = K( d2 )
PPP
PPP nnn
PPP nnnnn
PPP nn
P nnn
K
From this,√the possible λ’s are easily √ determinable.
√ Noting that we can
adjust K( d) by a square (i.e. K( d) = K( r2 d)), we readily see that
H ∼= K × /(K × )2 ⊕ K × /(K × )2 , where (K × )2 is the subgroup of squares in
×
K . Choosing the basis {(e1 , 0), (e2 , 0)} for E[2], for (x, y) 6= (e1 , 0), (e2 , 0),
we see that
δ : (x, y) 7→ (x − e1 , x − e2 ) ∈ E[2].
For (x, y) = (e1 , 0), (e2 , 0), one can readily work out δ by considering what
to do in order to divide by 2.
Remark. In general, if p6 |m, A[m] = (Z/mZ)2 dim A , so H ∼
2 dim A
= (K × /(K × )m ) .
So far we had not assumed any conditions on our field K. We will now
restrict our arguments to a global field K. Let MK be the set of places
(i.e. equivalence classes of absolute values) of K, Kv be the completion of
K with respect to a place v ∈ MK , Ov = {x ∈ Kv : |x|v ≤ 1} ⊆ Kv , and
Mv = {x ∈ Kv : |x|v < 1} the unique maximal ideal in Ov . By clearing
28
denominators, we can view A as being defined over Ov and reduce modulo
Mv to get a variety Av defined over the finite field Ov /Mv . If Av is an
abelian variety, we say A has good reduction at v; otherwise we say A has
bad reduction at v.
We next define a set S ⊂ MK such that S contains the archimedean
places, all places of bad reduction for A, and all places v | m (i.e. |m|v < 1).
We claim without proof that S is finite. This can be shown by extending
techiques from a previous lecture. We will need the following definitions to
prove the Weak Mordell-Weil Theorem:
Definition 21. Let K be a field, L/K a finite extension, and |·| an absolute
value on L. We say that | · | is unramified in L if {|x| : x ∈ K} = {|x| :
x ∈ L}.
Definition 22. A place v ∈ MK is unramified in L if for all | · | of L that
are in the class of v when restricted to K are unramified in L.
√
Example. Let K = Q, L = Q( D) with D a square-free integer. √ Observe
that for odd p, the p-adic absolute value in Q is ramified in Q( D) if and
only if p | D. (We say |·|/Q is ramified in L if there exists an absolute value
in L that restricts to | · | in Q which is not unramified.)
√ If p | D, D = pc
where (c, p) = 1 (since D is square-free). So | D|2p = |D|p = |pc|p = 1/p.
√ √
Thus | D|p = 1/ p 6∈ {|x|p : x ∈ Q} = pZ ∪ {0}. On the other hand if p
√ √
does not divide D, then we have | D|p = 1. Thus√{|x|p : x ∈ Q( D)} =
{|x|p : x ∈ Q}. We also note that 2 ramifies in Q( D) if and only if D is
even or D ≡ 3 (mod 4).
We will make use of the following important theorem:
Theorem 23. If P ∈ A(K), mQ = P . Then K(Q)/K is unramified outside
of S.
Note that S depends on the abelian variety A and the field K. So as you
vary P ∈ A(K), the theorem tells us that the extensions K(Q)/K are all
unramifiied outside the same set S. We will also need the following theorem:
Theorem 24 (Hermite). If K is a global field, S ⊂ MK finite, d ≥ 2 an inte-
ger (with p6 |d if p = charK > 0), then the set {L/K : L unramified outside of S, [L :
K] = d} is finite.
Theorem 25. Theorem 23 and 24 imply the Weak Mordell-Weil Theorem.
29
Proof. By Theorem 23, if P ∈ A(K), K(Q) is unramified outside of S for
any mQ = P . We also know that [K(Q) : K] ≤ #A[m]. Then by Theorem
24, there are only finitely many choices for such K(Q) and only finitely
many λ : Gal(K(Q)/K) ,→ A[m]. Thus the image of δ is finite. Thus
A(K)/mA(K) is finite, giving us the Weak Mordell-Weil Theorem.
Thus we are only left with proving Theorems 1 and 2. We will only do
the case of an elliptic curve over Q and m = 2, leaving the general case to
one of the standard texts.
Proof of Theorem 1. We have
E : y 2 = x3 + ax + b = (x − e1 )(x − e2 )(x − e3 )
v 2 = (u − e1 z 2 )(u − e2 z 2 )(u − e3 z 2 ).
30
Thus Q(Q) is unramified outside of S.
Proof of Theorem 2. We want to count the number of quadratic extensions
Q unramified outside of S = {2, p1 , . . . , pr }. These extensions are those
of √
Q( D) such that D is square-free and for all p ∈ / S, p6 |D. Thus for i = 0, 1,
Y
D=± pi i .
pi ∈S
There are finitely many such D, so there are finitely many extensions Q(D).
The general case of these theorems requires finiteness of class numbers
and Dirichlet’s Unit Theorem, but we will leave it at that.
We now assume that K is a global field, and MK the set of places of K.
Then for v ∈ MK , we let Kv represent the corresponding completion. We
can now look at the induced maps
δ : A(K)/mA(K) / HK
δv : A(Kv )/mA(Kv ) / HK
v
Proof. While we do not go through the proof of this theorem here, it follows
from much the same argument as the proof of the finiteness of A(K)/mA(K).
Define
Sham = Sm /δ(A(K)).
We want to figure out how to compute A(K). In order to do this, we need to
compute A(K)/mA(K). It turns out that Sm is computable, but it’s hard
31
to determine the image of A(K)/mA(K) in Sm , as it is not clear how to
decide if an element of Sm actually comes from A(K)/mA(K). It turns out
that there is a way around this.
Let l be a prime, l 6= p if charK = p > 0. Then the following commuta-
tive diagram holds:
0 / A(K)/lA(K) / Sl / Shal /0
O O O
Descent “algorithm”
We may now use the above to compute the image of A(K)/lA(K) in Sl in
the following manner. Consider two distinct parallel processes.
At some point, it is our hope that the subgroups obtained by following these
two processes will be the same. That is, there will come a point where we
can no longer build up our subgroup as in part 1, or restrict the subgroup
further, as we are doing in part 2. At this point, we will have found the
exact image of A(K)/lA(K) in Sl as desired.
Theorem 27. The descent “algorithm” will work if | Shalm | is bounded as
m → ∞.
Consider
Shal∞ := lim Shaln .
←−
n
L
We define the Tate-Shafarevich group by Sha = l Shal∞ .
32
Conjecture 28 (Tate-Shafarevich). Sha is finite.
If this conjecture indeed holds true in all cases, then our algorithm is
always valid. For now, the conjecture is known to be true only for some
elliptic curves over Q of small rank.
Example. Consider the abelian variety A over K = Q given by y 2 = (x −
e1 )(x − e2 )(x − e3 ) where ei ∈ Z. Let m = 2. Then the map δ is given by:
Let µ be the restriction of the map δ to the first coordinate. That is,
µ : A(Q) → Q× /(Q× )2
(x, y) 7→ x − e1
Sha2 = 0 ⇐⇒ the Hasse Principle holds for this particular set of equations.
For every curve C of genus 1 over a field K, one can associate an elliptic
curve E/K called its Jacobian. If K is a global field and C(Kv ) 6= ∅ for all
v ∈ MK , then C can be viewed as an element of Sha(E/K). Sha(E/K) = 0
if and only if the Hasse Principle holds for curves of genus 1 over K with
Jacobian E. If the conjecture of Tate and Shafarevich above holds then
failure of the Hasse Principle is measured by a finite quantity.
33
Suppose C/K has genus 1. If C(Kv ) = ∅ for some v, then C(K) = ∅.
On the other hand, if C(Kv ) 6= ∅ for all v ∈ MK , then [C] ∈ Sha(E/K). It
follows that we have
C(K) 6= ∅ ⇐⇒ [C] = 0.
If Sha(E/K) is finite, then we can verify whether [C] = 0 by a finite compu-
tation. To do this, we use a bilinear pairing (known as the Cassels pairing)
34
Remark: Sometimes the function
Y
H(a) = max {|ai |v }nv
0≤i≤n
v
35
Lemma: If a ∼ b, then h(a) = h(b).
Example: K = Q.
To see this, note that there must be some ai not divisible by p. For this ai ,
|ai |p = 1. For the rest, we note that each other aj is of the form aj = ps m
with p 6 |m and s ≥ 0, so |aj | = p−s ≤ 1. Hence we see that
36
Theorem: If K is a global field, given integers n, c ≥ 1, the set
{a ∈ Pn (K) : h(a) ≤ c}
is finite.
Before we give the proof of this theorem, the following two propositions
are left as exercises:
{α ∈ K : h(α) ≤ c}
37
It is then enough to bound the heights of the coefficients of this polynomial,
i.e. the symmetric functions in {σ(α) : σ ∈ Gal(K/Q)}. Defining
X
β= σi1 (α) · · · σik (α),
i1 ≤···≤ik
we note that X
h(β) = log+ |β|p .
p≤∞
|β|p = |β|v
X
= σi1 (α) · · · σik (α)
i1 ≤···≤ik
v
X
≤ |σi1 (α) · · · σik (α)|v
i1 ≤···≤ik
X
= |σi1 (α)|v · · · |σik (α)|v .
i1 ≤···≤ik
Since the function |σij (·)|v is yet another absolute value restricting to | · |p
on Q, label it | · |vij . Continuing our estimates,
X X
|σi1 (α)|v · · · |σik (α)|v = |α|vi1 · · · |α|vik
i1 ≤···≤ik i1 ≤···≤ik
n
≤ max{1, |α|v1 , . . . , |α|vn }k .
k
38
There is a stronger version of the theorem due to Northcott:
is finite.
Proof: By definition,
X
h(Φ(a)) = nv log max |φi (a)|v .
0≤i≤r
v
If we write φi (a) as
X
φi (a) = cid0 ,...,dn ad00 · · · adnn ,
d1 +···+dn =d
then we have
X
log |φi (a)|v ≤ log |cid0 ,...,dn |v |a0 |dv0 · · · |an |dvn
d1 +···dn =d
n+d
≤ log kv + d log max {|ai |v } + log
0≤i≤n d
39
Here kv = max |cid0 ,...,dn |v . It is clear that, for all but finitely many v,
log kv = 0. As before, the combinatorial term is not necessary in the non-
archimedean case. The result follows by summing over v.
40
The numerator and denominator have degree 2 in x and degree 1 in y. So
“h (Φ(a)) ≤ dh(a) + O(1)” gives 3.
Now 2. The x-coordinate of 2P is given by
2 2
3x + a (3x2 + a)2 − 8xy 2
=
2y 4y 2
(3x2 + a)2 − 8x(x3 + ax + b) A(x)
= 3
= ,
4(x + ax + b) B(x)
likewise,
X
h(P ) = h (x) = nv log (max {|x|v , 1}) .
v
|A(x)|v = a0 x4 + a1 x3 + · · · + an v
= x4 v a0 + a1 x−1 + · · · + an x−4 v
≥ |x|4v |a0 |v − |x|−1 a1 + · · · + an x−3
v v
≥ |x|4v (|a0 |v − Cv ) ,
for some constant Cv . Thus, log (|A(x)|v ) ≥ 4 log(|x|v )+Cv0 . Note in the non-
archimedean case, we can take Cv = 1 using the strong triangle inequality.
Also |a0 |v is almost always 1 (that is, all but finitely many are 1). Likewise
for B, we have |B(x)|v ≤ |x|3v + Cv ; so for those v with |x|v > 1, we get
41
Proof. If p(x)|B(x) then p(x)|f (x). If also p(x)|A(x), then p(x)|f 0 (x). But
(f (x), f 0 (x)) = 1, since f (x) has distinct roots (as we are working on an
elliptic curve).
1 = |u(x)A(x) + v(x)B(x)|
≤ Cv00 max {|A(x)|v , |B(x)|v } Cv00 = 1 for almost all v,
Tate’s approach:
h(2n P )
ĥ(P ) = lim .
n→∞ 4n
We have h(2P ) ∼ 4h(P ), which implies h(2P
4
)
∼ h(P ); thus
h(2n P )
is a Cauchy sequence.
4n
Formally, you get ĥ(2P ) = 4ĥ(P ); one needs to prove it works for all values
of m.
Remark. Now everytime you have a positive-definite quadratic form on a
finitely generated abelian group Γ of rank r means that you can P
embed Γ in
Rr in such a way that the quadratic form is the restriction of x2i . This
leads to many geometric question about the resulting lattice.
Open Problems
1. What is the “shape” of the Lattice? (square, skew, etc.)
42
2. Finding lower bounds for the smallest positive value of ĥ(P ) for P ∈
A(K). (There is a conjecture of Lang in this direction.)
3. Are there elliptic curves E over Q with rank E(Q) of arbitrarily large
values? (The largest known is something like 28.)
4. How often is the rank large? More generally, what is the distribution
of ranks?
Remark. Some key facts about the proof of C(Q) = 16. Let Y1 (m) be the set
of isomorphism classes of pairs (E, P ), where E is an elliptic curve P ∈ E
is of order m. It turns out that Y1 (m) is an algebraic curve; It has a map
to A1 via the j-invariant which comes from (E, P ) 7→ E. One finds that
Y1 (m)(Q) is finite if the genus of Y1 (m) is at least 2. For m large, the set
is actually empty. Proving C(Q) = 16 is extremely difficult.
Let K be a global field. A/K will be an abelian variety of dimension
n, MK the set of places of K, and S the set of places of bad reduction and
archimedean places.
Suppose v is a non-archimedean place, with corresponding absolute value |·|v
and completion Kv . Let Ov denote the ring of integers {x ∈ Kv : |x|v ≤ 1}
and Mv the unique maximal ideal {x ∈ Kv : |x|v < 1}. Note that Ov /Mv is
a finite field; we will write qv for its cardinality. We can view the equations
defining A/K as equations over Kv ; for all but finitely many v’s, these
equations are in Ov . We can look at the same equations in Ov /Mv .
We say v is of good reduction if we get from A an abelian variety Av over
Ov /Mv .
Theorem (Weil). Fix v ∈
/ S. There exists a polynomial Pv (T ) ∈ Z[T ] such
that
2n
Y
Pv (T ) = (1 − αi T )
i=1
43
1
where the αi ∈ C have |αi | = qv2 , and such that over the finite field Fqvm , for
all m ≥ 1 we have
Y2n
#Av (Fqvm ) = (1 − αim ).
i=1
44
From the latter expression, if #Ep (Fp ) > p “often enough”, then L(E/Q, 1) =
0.
We now examine whether the existence of many rational points forces the
existence of many points mod p. Empirically, it has been observed that
Y #Ep (Fp )
−→ ∞ ⇔ E(Q) is infinite.
p
p≤x
Indeed, if this is the case, then the product grows proportionally to (log x)r ,
where r = rankZ E(Q).
R r
where the regulator R = vol( (A(K)/A(K) tor
) with respect to canonical
height, Sha = Sha(A/K) is the Tate-Shafarevich group, the cv depend on Ψ
above, A(K)tor is the subgroup of A(K) consisting of points of finite order,
and A∗ is the dual abelian variety to A, which we don’t define. For elliptic
curves, A∗ = A.
Statements (1) and (1 21 ) are known for elliptic curves over Q from Wiles’
proof of Fermat’s Last Theorem.
The statements (1) and (1 21 ) are known for abelian varieties over func-
tion fields. If K/Fq (t), then L(A/K, s) is a polynomial in q −s (Weil, Dwork,
Grothendieck).
For elliptic curves over Q, ords=1 L(A/Q, s) ≤ 1 implies statement (2) (Ru-
bin, Kolyvagin, Gross-Zagier, ... )
If K is a function field, then rankZ A(K) ≤ ords=1 L(A/K, s), with equality
iff Sha is finite. In this case, (2) implies (3) (Tate, . . . , Kato-Trihan).
We start with a theorem, which was conjectured by Mordell.
45
Theorem 29. If K is a global field and C/K is a curve of genus ≥ 2, then
C(K) is finite except if K is a function field over Fq and C is defined over
Fq .
Remark. In particular, if K is a number field, then C(K) is finite. The
number field case is solved by Faltings in 1983. The function field case was
done by Samuel in the 60’s.
Before we embark on the arithmetic of curves we need to do some
geoemtry. We begin by defining divisors and Jacobians.
Suppose F is algebraically closed and C/F is a smooth irreducible pro-
jective curve. In general C(F ) is not finite.
P
Definition 30. A divisor on C is a formal sum D := P ∈C(F ) nP P where
nP : C(F ) → Z is supported on finitely many points.
0 =
P P 0 0
Given two divisors D = P n P P, D P nP P define D + D :=
0
P
P (nP + nP )P . Denote the set of all divisors on C by Div(C). In other
words, Div(C) is a free abelian group generated by C(F ). P
We also define a map deg : Div(C) → Z by deg(D) := P nP where
P 0
D = P nP P . Set Div (C) := ker(deg).
Recall that the function field F (C) is defined as a field of fractions of
R := F [x1 , . . . , xn ]/(f1 , . . . , fm )
where ordP (h) is the order of zero or pole of h at P . We have the following
fact.
Proposition 31. deg((h)) = 0 for any h ∈ F (C), that is, any function has
as many zeros as poles with multiplicities.
Verify this for a couple of examples. If F = C, then C is a compact
Riemann surface. Take any function h ∈ F (C), that is, a meromorphic
function on the surface. Then there are only finitely many zeros and poles
46
on C. We can take a closed path γ on C enclosing a disk that does not
contain any zeros or poles of h. By the residue theorem we have that
Z
dh X
0= = ordP (h).
γ h P
A(x)
If C = P1 , then F (C) = F (x). For h = B(x) , we have ord∞ (h) = deg(B) −
deg(A) hence the proposition holds.
Definition 32. Prin(C) := {(h) : h ∈ F (C)}.
Remark. By proposition 1, Prin(C) ⊂ Div0 (C) ⊂ Div(C). Also for any
h1 , h2 ∈ F (C), (h1 h2 ) = (h1 ) + (h2 ).
We are now in position to define the Jacobian of C.
Definition 33. The Jacobian of C is the group Div0 (C)/Prin(C).
Definition 34. We say divisors D, D0 on C are linearly equivalent if D −
D0 ∈ Prin(C) and denote D ∼ D0 if those are equivalent. In other words,
D ∼ D0 if and only if there exists h ∈ F (C) so that D − D0 = (h).
Here are a couple of examples.
0 1 P
Suppose
P D ∈ Div (P ). Then D is a form of α∈F nα [α] + n∞ [∞]. Note
that nα + n∞ = 0. Define a function
Y
h := (x − α)nα .
α∈F
(L) = P + Q + R − 3O
P + Q ∼ R + O.
47
This can be used iteratively on a divisor to trade
P two arbitrary
P points
by a pair of points
P one Pof which is O. Let D = n i P i − m i P i where
ni , mi > 0 and ni = mi . Then, applying this procedure gives:
X X
D ∼ Q+ ni − 1 O − R + mi − 1 O = Q − R
F (x)
Thus F (E) = {r(x) + ys(x) : r(x), s(x) ∈ F (x)} so we may write h(x, y) =
a(x)+yb(x)
c(x) , with gcd(a, b, c) = 1 (by getting a common denominator and
cancelling common factors).
We claim that if h has no poles in the affine part of E then c is a constant.
To see this, suppose deg(c) ≥ 1. Then ∃α ∈ F with c(α) = 0. Let β satisfy
β 2 = f (α) (where f (x) = x3 + ax + b) so the points (α, ±β) ∈ E. If h has
no pole at (α, ±β), the numerator must vanish; so a(α) ± b(α)β = 0. Since
it must hold for both (α, +β) and (α, −β), we can add both equations to
deduce a(α) = 0, and so b(α)β = 0.
If β 6= 0, then b(α) = 0; implying that (x − α) is a common factor to a,
b, and c, contradiction. If β = 0 then f (α) = 0. So y 2 = f (x) = (x − α)g(x).
We have ord(α,0) (x − α) = 2 and ord(α,0) y = 1, so a(x) = (x − α)r a1 (x) and
ord(α,0) (a(x)) = 2r ≥ 2 and similarly for b. In particular, their orders are
even. If ord(α,0) (h) ≥ 0 then ord(a + by) ≥ ord(c), so min {ord(a), ord(b)} ≥
ord(c). But ord(b) > 0 implies b(α) = 0, contradiction as above. So
ord(b(α)y) = ord(b) + ord(y) = ord(y) = 1. Thus we have 1 ≥ 2, con-
tradiction.
Thus we have h = a + yb for a, b ∈ F [x]. Now the ordO a = −2 deg a and
ordO by = −2 deg b − 3 (the latter follows since y 2 = f (x) the ordO x = −2
and ordO y = −3). Now −1 = ordO h = min {−2 deg a, −2 deg b − 3} for
a, b 6= 0. [If a = 0 or b = 0, we have a term not appearing in the minimum.]
We get a contradiction, and this shows injectivity.
48
Example. Let y 2 = f (x) with deg f (x) = 5 and f has distinct roots. It
is a fact that this gives a curve of genus 2. f (x) − (ax2 + bx + c) has 5
zeroes say α1 , . . . , α5 . Say βi = −(aαi2 + bαi + c) and Pi = (αi , βi ). Then
(h) = P1 + P2 + · · · + P5 − 5P∞ , where h(x, y) = y + (ax2 + bx + c).
Now given P1 , P2 , P3 , we can choose a, b, c so that P1 + P2 + P3 − 3P∞ ∼
2P∞ − (P4 + P5 ) is the relation given by h.
Using these kinds of relations, we can trade three points for two points
and prove that every divisor D, deg D = 0, is linear equivalent to a divisor
of the form P1 + P2 − 2P∞ .
Define C (2) = C × C/S2 (i.e. mod out by switching coordinates); this is
the set of unordered pairs. Hence we have a map Φ : C (2) → Jac(C) given
by {P1 , P2 } 7→ P1 + P2 − 2P∞ , which is surjective.
Unfortunately, it is not injective. The divisor of (x − α) is (α, β) +
(α, −β)−2P∞ (where β 2 = f (α)). Hence all the pairs of the form {(α, β), (α, −β)}
gives O under Φ. It turns out that this is the only source of non-injectivity.
The other points in C (2) uniquely represent a point in Jac(C). [In algebraic
geometry terms, Φ is a birational map and Jac(C) is the blow-down of C (2)
along a curve.]
For curves of genus 2, the Jacobian is a surface. In general the Jacobian
of a curve of genus g is an algebraic variety of dimension g. In fact, Jac(C)
is projective and, since it has a group law, it is an abelian variety. Moreover,
Jac(C) is birational to C (g) := C g /Sg , where {P1 , . . . , Pg } 7→ P1 + · · · + Pg −
gP0 for some fixed P0 . The sources of non-injectivity are more complicated,
but it is a collection of blow-downs.
Example. If F = C and C/C you can choose ω1 , . . . , ωg linear independent
holomorphic differential forms on C with Jac(C) → Cg /L, where L is a
Pg R Pi R Pi
lattice, given by P1 + · · · + Pg − gP0 7→ i=1 P0 ω1 , . . . , P0 ωg ; the lattice
nR R o
is γ ω 1 , . . . , γ ω g : γ ∈ π 1 (C(C)) .
49
Definition. We say D is defined over K if σ(D) ∼ D for all σ ∈ Gal(F/K)
It can be shown that Jac(C)(K) is exactly the set of equivalence classes
of divisors of degree zero defined over K.
Example. For g = 1, Jac(E) = E over F . It turns out Jac(E)(K) = E(K),
since σ(P )−O = σ(P −O) ∼ P −O (that is using the fact that O is rational)
implies σP ∼ P . Which can be shown to imply σ(P ) = P . If this happens
for all σ ∈ Gal(F/K), we have P ∈ E(K).
If C is curve of genus 1 with no rational points, we get Jac(C) = E is
an elliptic curve E/K with E and C isomorphic over F .
Example. Consider g = 2. Let y 2 = f (x), where deg f = 5. Assume P0
is a rational point at ∞. Now Jac(C) = {[P1 + P2 − 2P0 ] : P1 , P2 ∈ C(F )};
note σ(P1 ) + σ(P2 ) − 2P0 = σ(P1 + P2 − 2P0 ) ∼ P1 + P2 − 2P0 “usually”
implies σ(P1 ) + σ(P2 ) = P1 + P2 . If σ(Pi ) = P2 for i = 1, 2 for all σ then
Pi ∈ C(K). Unfortunately this need not always happen. For example, we
could have σ(P1 ) = P2 and σ(P2 ) = P1 , which would occur if P1 ∈ C(L)
where [L : K] = 2 and P2 is the galois conjugate of P1 ; we have
[
Jac(C)(K) = {[P1 + P2 − 2P0 ] : P1 , P2 ∈ C(K)}∪ [P + P − 2P0 ] : P ∈ C(L)
[L:K]=2
α : C −→ J
P 7→ P − P0
Remark: We can use any divisor D0 with deg(D0 ) = 1 and get a map
α : C −→ J
P 7→ P − D0
So C(K) ⊆ J(K). Mordell-Weil implies that J(K) is a finitely generated
abelian group when K is a global field. Does it help to understand C(K)?
50
•Today:
Theorem(Chabauty ’38) If K is a number field and rankZ J(K) < g,
then C(K) is finite.
(We use p-adic analysis to prove it)
We can embed K in Qp for some prime p. Let’s study J(Qp ) first.
Lemma There exists a neighborhood of 0 ∈ J(Qp ) which is isomorphic
(as a p-adic analytic group) to Zgp .
More precisely, there are local coordinates t1 , ..., tg near 0 and power
series λ1 , ..., λg in t1 , ..., tg converging in some neighborhood U of 0 such that
P, Q ∈ U, P = (t1 (P ), ..., tg (P )), etc., then λi (t1 (P + Q), ..., tg (P + Q)) =
λi (t1 (P ), ..., tg (P )) + λi (t1 (Q), ..., tg (Q)).
Assume the Lemma for a while.
Proof of Theorem: Let P1 , ..., Pr be generators of the free part of
J(K), r < g. Replace Pi , if necessary, by pm Pi so that without loss of
generality Pi ∈ U . Consider the vectors
51
Now Q1 + ... + Qg − gP1 cover an open set of J(Qp ) as Q1 , ..., Qg vary. So
λ ≡ 0, a contradiction.
Idea of Proof of Lemma: Given P ∈ J, translation by P (Q 7→ Q+P )
gives a map τP : J −→ J such that 0 7→ P . The derivative
dτP : T0 J −→ TP J
is an isomorphism. Therefore the dual spaces (T0 J)∗ and (TP J)∗ are also
isomorphic.
Given an element of (T0 J)∗ say v we get a 1-form on J, given by ω =
(dτP )∗ (v) i.e., for each P an element of (TP J)∗ . We have τP∗ ω = ω. The
function Z P
λ : P 7→ ω
0
is linear in P . In fact,
Z P +Q Z P Z P +Q Z P Z Q Z P Z Q
ω= ω+ ω= ω+ τP∗ ω = ω+ ω
0 0 P 0 0 0 0
52
If X/K is a variety, X is the set of zeros of f1 , f2 , . . . , fm ∈ K[x1 , x2 , . . . , xn ].
(p) (p) (p)
Define X (p) to be the set of common zeroes of f1 , f2 , . . . , fm . There is
a map F : X → X (p) given by (x1 , x2 , . . . , xn ) 7→ (xp1 , xp2 , . . . , xpn ) because
(f (x1 , x2 , . . . , xn ))p = f (p) (xp1 , xp2 , . . . , xpn ). (Follows from the fact that (x +
y)p = xp + y p in characteristic p).
Remark. If X is defined over Fp , then X (p) = X. If X is defined over
2 m
Fpm , then, if we define X (p ) = (X (p) )(p) , etc. then X (p ) = X, and we get
a map F m : X → X via
2) m)
X → X (p) → X (p → · · · → X (p
.
Let A/K be an abelian variety. Then we have the map F : A → A(p) .
We also have the map [p] : A → A defined by P 7→ pP . Fact: There is a
map V : A(p) → A such that [p] = V ◦ F .
Remark. If Φ : X → Y is an onto map of varieties over K, we have an
injection of function fields K(Y ) ,→ K(X). Then the map Φ is separable if
K(X)/K(Y ) is separable.
Definition. An abelian variety A is called ordinary if V is separable.
Theorem 35. If K is a global field of characteristic p > 0 and C/K is a
curve of genus g ≥ 2 such that C is not defined over K p and J = Jac(C) is
ordinary, then C(K) is finite.
Remark. The natural hypothesis for the Mordell conjecture is that C is not
defined over Fq . It can be shown that this is equivalent to C not defined over
53
m m
Kp Kp =
T
for some m ≥ 1: If K is a global field of characteristic p,
m≥1
Fq . The proof can be adapted to deal with this more general situation.
What happens over finite fields: Consider C/Fq , with q = pf . Let K
be a global field of characteristic p wth constant field Fq . Suppose P ∈
C(K) − C(Fq ). We can consider F f m (P ) ∈ C(K), and we get infinitely
many points with m = 1, 2, . . ..
Example. C : y 2 = x5 + 1 over F3 . Take K = F3 (t, s) where s2 = t5 + 1.
m m
Then (t3 , s3 ) ∈ C(K).
In place of J(K), we could have used any subgroup Γ ⊆ J(K sep ) with
the property that Γ/pΓ is finite (e.g., one can take Γ to be the group of
prime to p torsion points in J(K sep ). A special case of the Manin-Mumford
conjecture says that C ∩ Jtor is finite.
Another example that one can take is to embed J(K) in J(Kv ) for a
completion Kv of K and take J(K) in J(K). (In the case of number fields,
the Chabauty argument proves that J(K) ∩ C is finite if J(K) ⊆ J(Kv )
and rk(J(K)) < genus(C)). In characteristic p, J(K) ⊆ J(Kv ) is always
“small”, where “big” means that it “contains a neighborhood of 0,” and
“small” means not “big.” In characteristic 0, J(K) ⊆ J(Kv ) is “small” if
v is non-Archimedean and rk(J(K)) < genus(C). Otherwise, it is usually
“big.”
Conjecture. Let K be a global field and A/K an Abelian variety. If X ⊂ A
is a closed algebraic set defined over K, then
Y
X(Kv ) ∩ A(K) = X(K)
v∈MK
54
We’ll sketch the proof of this conjecture in characteristic p under the
following assumptions:
Recall that in this case the Mordell conjecture was proved by writing
55
Let’s now take a look at the zero characteristic case and assume the
condition rank(A(K)) < dim A of Chabauty’s theorem. In this case we
construct λ : A(Kv ) −→ Kv analytic with A(K) ⊂ λ−1 (0), but X ∩ λ−1 (0)
is finite. Zv = X ∩ λ−1 (0) has the property (*). Unfortunately Zv depends
on v in characteristic zero, so we cannot reduce the 1-dimensional case to
the 0-dimensional one.
Suppose instead that rank(A(K)) < dim A − 1, then you can construct
λ1 , λ2 : A(Kv ) −→ linearly independents, with A(K) ⊂ λ−1 i (0). Define
−1 −1
Zv = X ∩ λ1 ∩ λ2 .
Conjecture (Stoll). In this situation, there is a finite set Z independent of
v with Zv ⊂ Z, for all v
If this is true then the reduction of the 1-dimensional case to the 0-
dimensional case works in zero characteristic.
56