Elliptic Curves
Daniel Naylor

Contents

1Fermat’s Method of Infinite Descent
1.1A variant for polynomials
2Some Remarks on Plane Curves
2.1The degree of a morphism
3Weierstrass Equations
4The Group Law
5Isogenies
6The Invariant Differential
7Elliptic Curves over Finite Fields
7.1Zeta functions
8Formal Groups
9Elliptic Curves over Local Fields
10Elliptic Curves over Number Fields: The torsion subgroup
11Kummer Theory
12Elliptic Curves over Number Fields: The weak Mordell-Weil Theorem
13Heights
14Dual isogenies and the Weil pairing
15Galois Cohomology
16Descent by Cyclic Isogeny
16.1Descent by 2-isogeny
16.2Birch Swinnerton Dyer Conjecture
Index

1 Fermat’s Method of Infinite Descent

PIC

Definition (Rational, primitive triangle). A triangle is rational if a,b,c.

A triangle is primitive if a,b,c and are coprime.

Lemma 1.1. Assuming that:

Then Δ is of the form

PIC

for some integers u>v>0.

Proof. Without loss of generality a odd, b even (work modulo 4). This then forces c odd.

Then

(b2)2=c+a2ca2,

and note that all the fractions are integers. Also note that the product on the right hand side is a product of positive coprime integers.

Unique prime factorisation in gives that c+a2=u2, ca2=v2 for some u,v.

Then a=u2v2, b=2uv, c=u2+v2.

Definition (Congruent number). D>0 is a congruent number if there exists a rational triangle Δ with area(Δ)=D.

Note. It suffices to consider D>0 square-free.

Example. D=5,6 are congruent.

Lemma 1.2. Assuming that:

  • D>0

Then D is congruent if and only if Dy2=x3x for some x,y, y0.

Proof. Lemma 1.1 shows

D congruentDw2=uv(u2v2)

for some u,v,w, with w0. Put x=uv and y=wv2.

Fermat showed that 1 is not a congruent number.

Theorem 1.3. There is no solution to

w2=uv(u+v)(uv)(∗)

for u,v,w and w0.

Proof. Without loss of generality u,v are coprime, u>0, w>0. If v<0 then replace (u,v,w) by (v,u,w). If u,v both odd then replace (u,v,w) by (u+v2,uv2,w2).

Then u,v,u+v,uv are pairwise coprime positive integers with product a square.

Unique factorisation in gives

u=a2,v=b2,u+v=c2,uv=d2

for some a,b,c,d>0.

Since uv(mod2), both c and d are odd.

Then consider:

(c+d2)2+(cd2)2=c2+d22=u=a2.

PIC

This is a primitive triangle. The area is c2d28=v4=(b2)2.

Let w1=b2. Lemma 1.1 gives w12=u1v1(u1+v1)(u1v1) for u1,v1. Therefore we have a new solution to (). But 4w12=b2=v|w2 so w112w.

So by Fermat’s method of infinite descent, there is no solution to () .

1.1 A variant for polynomials

In Section 1, K is a field with charK2.

Write K¯ for the algebraic closure of K.

Lemma 1.4. Assuming that:

  • u,vK[t] coprime

  • αu+βv is a square for 4 distinct (α:β)1

Then u,vK.

Proof. Without loss of generality K=K¯.

Changing coordinates on 1, we may assume the ratios (α:β) are (1:0),(0:1),(1:1),(1:λ) for some λK{0,1} (Möbius map).

u=a2v=b2uv=(a+b)(ab)uλv=(a+μb)(aμb)

(where μ is a square root of λ). Unique factorisation in K[t] gives that a+b, ab, a+μb, aμb are squares. But

max( deg (a), deg (b))12max( deg (u), deg (v)).

So Fermat’s method of infinite descent, we get a contradiction, unless the degrees of u and v are zero. So u,vK.

Definition 1.5 (Elliptic curve (temporary definition)).

  • (i) An elliptive curve EK is the projective closure of the plane affine curve
    y2=f(x)

    where fK[x] is a monic cubic polynomial with distinct roots in K¯. We call this equation “a Weierstrass equation”.

  • (ii) For LK any field extension
    E(L)={(x,y)L2|y2=f(x)}{0}.

    0 is the “point at infinity” that we get because we take the projective closure.

Fact: E(L) is naturally an abelian group.

In this course, we study E(K) for K being a finite field, local field ([K:p]<) or number field ([K:]<).

Lemma 1.2 and Theorem 1.3 tells us that if E is y2=x3x, then

E()={0,(0,0),(±1,0)}.

Corollary 1.6. Let EK be an elliptic curve. Then E(K(t))=E(K).

Proof. Without loss of generality K=K¯. By a change of coordinates, we may assume

y2=x(x1)(xλ)

for some λK{0,1}. Suppose (x,y)E(K(t)). Put x=uv, where u,vK[t] are coprime.

Then w2=uv(uv)(uλv) for some wK[t].

Unique factorisation in K[t] gives that u,v,uv,uλv are squares. Hence by Section 1.1, u,vK, so xK, so yK.

2 Some Remarks on Plane Curves

Work over K=K¯.

Definition 2.1 (Rational plane affine curve). A plane affine curve C={f(x,y)=0}𝔸2 is rational if it has a rational parametrisation, i.e. ϕ(t),ψ(t)K(t) such that:

  • (i) 𝔸1𝔸2, t(ϕ(t),ψ(t)) is injective on 𝔸1{finite set}.
  • (ii) f(ϕ(t),ψ(t))=0.

Example 2.2.

Remark 2.3. The genus g(C)0 is an invariant of a smooth projective curve C.

Proposition 2.4. Assuming that:

  • K=K¯

  • C a smooth projective curve

Then
  • (i) C is rational (see Definition 2.1) if and only if g(C)=0.
  • (ii) C is an elliptic curve (see Definition 1.5) if and only if g(C)=1.

Proof.

Order of vanishing

C algebraic curve, function field K(C), PC smooth point.

We write ordp(f) for the order of vanishing of fK(C)× at P (negative of f has a pole).

Fact: ordp:K(C)× is a discrete valuation, i.e. ordp(f1f2)=ordp(f1)+ordp(f2) and ordp(f1+f2)min(ordp(f1),ordp(f2)).

Definition (Uniformiser). tK(C)× is a uniformiser at P if ordp(t)=1.

Example 2.5. C={g=0}𝔸2, gK[x,y] irreducible.

K(C)=FracK[x,y](g).

g=g0+g1(x,y)+g2(x,y)+

where gi are homogeneous of degree i.

Suppose P=(0,0)C is a smooth point, i.e. g0=0, g1(x,y)=αx+βy with α,β not both zero.

PIC

Fact: γx+δyK(C) is a uniformiser at P if and only if αδβγ0.

Example 2.6.

{y2=x(x1)(xλ)}𝔸2

where λ0,1. Projective closure (x=XZ, y=YZ):

{Y2Z=X(XZ)(XλZ)}2.

Let P=(0:1:0).

Aim: Compute ordp(x) and ordp(y).

Put w=ZY, t=XY. So

w=t(tw)(tλw)(∗)

Now Pis the point (t,w)=(0,0). This is a smooth point with

ordp(t)=ordp(tw)=ordp(tλw)=1.

() implies ordp(w)=3. Therefore ordp(x)=ordp(tw)=13=2 and ordp(y)=ordp(1w)=3.

Riemann Roch Spaces

Let C be a smooth projective curve.

Definition (Divisor). A divisor is a formal sum of points on C, say

D=PCnpP

where np and np=0 for all but finitely many PC.

We write degD=nP.

We say D is effective (written D0) if nP0 for all P.

If fK(C)×, then div(f)=PCordp(f)P.

The Riemann Roch space of DDiv(C) is

L(D)={fK(C)×|div(f)+D0}{0},

i.e. the K-vector space of rational functions on C with “poles no worse than specified by D”.

We quote: Riemann Roch for genus 1:

dimL(D)={ deg Dif  deg D>00 or 1if  deg D=00if  deg D<0

For example, in Example 2.6:

L(2P)=1,xL(3P)=1,x,y

Proposition 2.7. Assuming that:

  • K=K¯, charK2

  • C2 a smooth plane cubic

  • PC a point of inflection

Then we may change coordinates such that
C:Y2Z=X(XZ)(XλZ)

for some λ0,1 and P=(0:1:0).

Proof. We change coordinates such that P=(0:1:0), Tp(C)={Z=0}, and C:{F(X,Y,Z)=0}2.

PC part of inflection implies F(t,1,0)=constt3, i.e. F has no terms X2Y, XY2 or Y3.

Therefore

FY2Z,XYZ,YZ2,X3,X2Z,XZ2,Z3.

The Y2Z coefficient must be 0 otherwise PC is singular, and the coefficient of X3 is 0 otherwize Z|F.

We are free to rescale X, Y, Z and F. Then without loss of generality C is defined by

Y2Z+a1XYZ+a3YZ2=X3+a2XzZ+a4XZ2+a6Z3.

Weierstrass form.

Substituting YY12a1X12a3Z, we may suppose a1=a3=0. Now

Y2Z=Z3f(XZ)

for some monic cubic polynomial f.

C is smooth, so f has distinct roots. Without loss of generality say 0,1,λ. Then C is given by

Y2Z=X(XZ)(XλZ).

Remark. It may be shown that the points of inflection on a smooth plane curve

C={F(X1,X2,X3)=0}2

are given by

F=det(2FXiXj)Hessian=0,

2.1 The degree of a morphism

Let ϕ:C1C2 be a non-constant morphism of smooth projective curves.

Then ϕ:K(C2)K(C1), ffϕ.

Definition (Degree of a morphism). degϕ[K(C1):ϕK(C2)].

Definition (Separable morphism). ϕ is separable if K(C1)ϕK(C2) is a separable field extension.

Definition (Ramification index). Suppose PC1, QC2, ϕ:PQ. Let tK(C2) be a uniformiser at Q.

The ramification index of ϕ at P is

eϕ(P)=ordP(ϕt)

(always 1, independent of choice of t).

Theorem 2.8. Assuming that:

  • ϕ:C1C2 a non-constant morphism of smooth projective curves

Then
Pϕ1Qeϕ(P)= deg ϕQC2.

Moreover, if ϕ is separable then eϕ(P)=1 for all but finitely many PC1. In particular:

  • (i) ϕ is surjective (on K¯-points)
  • (ii) #ϕ1(Q)degϕ
  • (iii) If ϕ is separable then equality holds in (ii) for all but finitely many QC2.

Remark 2.9. Let C be an algebraic curve. A rational map is given C−−→n, P(f0(P):f1(P)::fn(P)) where f0,f1,,fnK(C) are not all zero.

Important Fact: If C is smooth then ϕ is a morphism.

3 Weierstrass Equations

In this section, we drop the assumption that K=K¯, but we instead assume that K is a perfect field.

Definition (Elliptic curve). An elliptic curve EK is a smooth projective curve of genus 1, defined over K with a specified K-rational point 0EE(K).

Example. {X3+pY3+p2Z3=0}2 is not an elliptic curve over , since it has no -rational points.

Theorem 3.1. Assuming that:

Then E is isomorphic over K to a curve in Weierstrass form via an isomorphism taking 0E to (0:1:0).

Remark. Proposition 2.7 treated the special case E is a smooth plane cubic and 0E is a point of inflection.

Fact: If DDiv(E) is defined over K (i.e. fixed by Gal(K¯K)) then L(D) has a basis in K(E) (not just K¯(E)).

Proof of Theorem 3.1. L(2.0E)L(3.0E). Pick basis 1,x for L(2.0E) and 1,x,y for L(3.0E).

Note: ord0E(x)=2, ord0E(y)=3.

The 7 elements 1,x,y,x2,xy,x3,y2 in the 6-dimensional space L(6.0E) must satisfy a dependence relation.

Leaving out x3 or y2 gives a basis for L(6.0E) since each term has a different order pole at 0E.

Therefore the coefficients of x3 and y2 are non-zero. Rescaling x and y (if necessary) we get

y2+a1xy+a3y=x3+a2x2+a4x+a6

for some aiK.

Let E be the curve defined by this equation (or rather its projective closure). There is a morphism

ϕ:EE2P(x(P):y(P):1)=(xy(P):1:1y(P))0E(0:1:0)

Then

[K(E):K(x)]=deg(Ex1)=ord0E(1x)=2[K(E):K(y)]=deg(Ey1)=ord0E(1y)=3

This gives us a diagram of field extensions:

           K (E)


           K (x,y)


K (x)                  K (y)
By the Tower Law (since 2,3 are coprime), we get that [K(E):K(x,y)]=1. Hence K(E)=K(x,y)=ϕK(E), so degϕ=1, so ϕ is birational.

If E is singular then E and E are rational, contradiction.

So E is smooth. Then Remark 2.9 implies that ϕ1 is a morphism. So ϕ is an isomorphism.

Proposition 3.2. Assuming that:

Then EE over K if and only if the equations are related by a change of variables x=u2x+ry=u3y+u2sx+t

where u,r,s,tK with u0.

Proof.

A Weierstrass form equation defines an elliptic curve if and only if it defines a smooth curve, which happens if and only if Δ(a1,,a6)0, where Δ[a1,,a6] is a certain polynomial.

If charK2,3, we may reduce to the case E:y2=x3+ax+b, with discriminant Δ=16(4a3+27b2).

Corollary 3.3. Assuming that:

  • charK2,3 and K is perfect

  • elliptic curves E:y2=x3+ax+b, E:y2=x3+ax+b

Then E and E are isomorphic over K if and only if a=u4a, b=u6b for some uK.

Proof. E and E are related by a substitution as in Proposition 3.2 with r=s=t=0.

Definition (j-invariant). The j-invariant is

j(E)=1728(4a3)4a3+27b2.

Corollary 3.4. Assuming that:

  • EE

Then j(E)=j(E). Moreover, the converse holds if K=K¯.

Proof.

EE{a=u4ab=u6b for some uK(a3:b2)=((a)3:(b)2)j(E)=j(E)

and the converse holds if K=K¯ (to go backwards on the step, we only need to take some kind of n-th root).

4 The Group Law

Let E2 be a smooth plane cubic, and 0EE(K).

PIC

E meets any line in 3 points counted with multiplicity.

Let S be the third point of intersection of PQ with E, and R be the third intersection point of 0ES and E.

Define PQ=R.

If PQ then take TPE instead of PQ etc.

This is called the “chord and tangent process”.

Theorem 4.1. (E,) is an abelian group.

Note. E here means E(K¯). As mentioned before, we only ever mean “over K” if it is explicitly mentioned (otherwise we are always working “over K¯”).

Proof.

Definition (Linearly equivalent). D1,D2Div(E) are linearly equivalent if there exists fK¯(E) such that div(f)=D1D2.

Write D1D2 and [D]={D:DD}.

Definition. Pic(E)=Div(E), Pic0(E)=Div0(E). where Div0(E)={DDivE| deg D=0}.

Proposition 4.2. Assuming that:

  • we define

    ψ:EPic0(E)P[(P)(0E)]

Then
  • (i) ψ(PQ)=ψ(P)+ψ(Q).
  • (ii) ψ is a bijection.

Proof.

Formulae for E in Weierstrass form
E:y2+a1xy+a3y=x3+a2x2+a4x+a6(∗)

0E=(0:1:0).

PIC

P1P2=P3. P1=(x1,(a1x+a3)y1).

Substituting y=λx+ν into () and looking at coefficient of x2 gives

λ2+a1λa2=x1+x2+x=x3.

Therefore

x3=λ2+a1λa2x1x2y3=(a1x+a3)y=(a1x3+a3)(λx3+ν)=(λ+a1)x3a3ν

It remains to find formulae for λ and ν.

Corollary 4.3. E(K) is an abelian group.

Proof. It is a subgroup of (E,). Identity 0EE(K) by definition.

Closure / inverses: see formulae above.

Associative / commutative: inherited.

Theorem 4.4. Elliptic curves are group varieties, i.e. [1]:EE; PP and :E×EE; (P,Q)PQ are morphisms of algebraic varieties.

Proof.

Statement of Results

The isomorphisms in (i), (ii), (iv) respect the relevant topologies.

Brief remarks on the case K=

Let

Λ={aω1+bω2:a,b},

where ω1,ω2 a basis for as an -vector space.

Then

{meromorphic functions on the Riemann surface Λ}{Λ-invariant meromorphic functions on }.

The function field of Λ is generated by

(z)=1z2+0λΛ(1(zλ)21λ2)(z)=2λΛ1(zλ)3

These satisfy

(z)2=4(z)3g2(z)g3

for some g2,g3 depending only on Λ.

One shows Λ=E() (isomorphism as groups and as Riemann surfaces) where E: y2=4x3g2xg3.

Theorem (Uniformisation Theorem). Every elliptic curve over arises this way (one proof uses modular forms).

Definition. For n, let [n]:EE be defined by

PPPPn copies

for n0, and [n]=[1][n].

Definition (n-torsion subgroup). The n-torsion subgroup of E is

E[n]=ker(E[n]E).

If K= then E()Λ. Therefore

{E[n](n)2(1)deg[n]=n2(2)

We’ll show (2) holds over any field K, and (1) holds if charKn.

Lemma 4.5. Assuming that:

  • charK2

  • E: y2=f(x)=(xe1)(xe2)(xe3), e1,e2,e3K¯

Then E[2]={0E,(e1,0),(e2,0),(e3,0)}(2)2.

Proof. Let P=(x,y)E. Then

[2]P=0P=P(x,y)=(x,y)y=0

5 Isogenies

Let E1,E2 be elliptic curves.

Definition (Isogeny). An isogeny ϕ:E1E2 is a nonconstant morphism with ϕ(0E1)=0E2 (by Theorem 2.8, a morphism is nonconstant if and only if surjective on K¯-points).

We say E1 and E2 are isogenous in this case.

Definition. Hom(E1,E2)={isogenies E1E2}{0}.

This is an abelian group under

(ϕ+ψ)(P)=ϕ(P)+ψ(P).

If E1ϕE2ψE3 are isogenies then ψϕ is an isogeny.

Tower Law implies deg(ψϕ)= deg (ϕ) deg (ψ).

Proposition 5.1. Assuming that:

  • 0n

Then [n]:EE is an isogeny.

Proof. [n] is a morphism by Theorem 4.4. We must show [n][0].

Assume charK2.

Case n=2: Lemma 4.5 implies E[2]E implies [2][0].

Case n odd: Lemma 4.5 implies 0TE[2]. Then nT=T0 which gives [n][0].

Now use [mn]=[m][n].

If charK=2, then could rpelace Lemma 4.5 with an explicit lemma about 3-torsion points.

Corollary 5.2. Hom(E1,E2) is a torsion free -module.

Theorem 5.3. Assuming that:

Then ϕ(P+Q)=ϕ(P)+ϕ(Q) for all P,QE1.

Proof (sketch). ϕ incudes

ϕ:Div0(E1)Div0(E2)PE1nPPPE1nPϕ(P)

Recall ϕ:K(E2)K(E1).

   K (E1)


normKm(aEp2)
Fact: If fK¯(E1) then

div(NK(E1)K(E2)f)=ϕ(divf).

So ϕ sends principal divisors to principal divisors.

Since ϕ(0E1)=0E2, the following diagram commutes:

      P              E1            E2             Q

                        0             0
ϕ∼=∼=ϕ∗    [(P )− (0E1)]    Pic (E1)      Pic(E2)        [(Q )− (0E2)]
ϕ a group homomorphism implies ϕ is a group homomorphism.

Lemma 5.4. Assuming that:

Then there exists a morphism ξ making the following diagram commute:
  E1      E2

ϕxxξ ℙ1      ℙ1
 12
(xi= x-coordinate on a Weierstrass equation for Ei). Moreover if ξ(t)=r(t)s(t), r,sK[t] coprime, then
degϕ= deg ξ=max( deg (r), deg (s)).

Proof. For i=1,2. K(Ei)K(xi) is a degree 2 Galois extension, with Galois group generated by [1].

Theorem 5.3 implies that ϕ[1]=[1]ϕ.

So if fK(x2) then

[1]ϕf=ϕ[1]f=ϕf.

Therefore ϕfK(x1).

                    K (E1) = K (x1,y1)

   K (x1)


                    K (E2) = K (x2,y2)


2dd2eeggKϕξ (x2)
In particular, ϕ=ξ(x1) for some rational function ξ.

Tower Law implies degϕ= deg ξ.

Now K(x2)K(x1), x2ξ(x1)=r(x1)s(x1), r,sK[t] coprime.

We claim the minimal polynomial of x1 over K(x2) is

F(t)=r(t)s(t)x2K(x2)[t].

Check:

Therefore

degξ=[K(x1):K(x2)]= deg F=max( deg r, deg s).

Lemma 5.5. deg[2]=4.

Proof. Assume charK2,3 (the result is true even in the case of charK{2,3}, but we will only prove the simpler case).

E: y2=x3+ax+b=f(x).

If P=(x,y)E, then

x(2P)=(3x2+a2y)22x=(3x2+a)28xf(x)4f(x)=x4+f(x)

The numerator and denominator are coprime. Indeed, otherwise there exists 𝜃K¯ with f(𝜃)=f(𝜃)=0 and hence f has a multiple root (contradiction).

Now Lemma 5.4 implies deg[2]=max(4,3)=4.

Definition (Quadratic form in an abelian group). Let A be an abelian group. We say q:A is a quadratic form if

  • (i) q(nx)=n2q(x) for all n, xA.
  • (ii) (x,y)q(x+y)q(x)q(y) is -bilinear.

Lemma 5.6. q:A is a quadratic form if and only if it satisfies the parallelogram law:

q(x+y)+q(xy)=2q(x)=2q(y)x,yA.

Proof.

Theorem 5.7. deg:Hom(E1,E2) is a quadratic form.

Note. We define deg(0)=0.

For the proof we assume charK2,3. Write E2: y2=x3+ax+b.

Let P,QE2 with P,Q,P+Q,PQ0. Let x1,,x4 be the x-coordinates of the 4 points.

Lemma 5.8. There exists W0,W1,W2[a,b][x1,x2] of degree in x1 and of degree 2 in x2 such that

(1:x3+x4:x3x4)=(W0:W1:W2).

Proof. Two methods.

We show that if ϕ,ψHom(E1,E2), then

deg(ϕ+ψ)+ deg (ϕψ)2 deg (ϕ)+2 deg (ψ).

We may assume ϕ,ψ,ϕ+ψ,ϕψ0 (otherwise trivial, or use deg[1]=1, deg[2]=4).

ϕ:(x,y)(ξ1(x),)ψ:(x,y)(ξ2(x),)ϕ+ψ:(x,y)(ξ3(x),)ϕψ:(x,y)(ξ4(x),)

Lemma 5.8 implies

(1:ξ3+ξ4:ξ3ξ4)=((ξ1ξ2)2:).

Put ξi=risi, ri,siK[t] coprime.

(s3s4:r3s4+r4s4:r3r4)coprime=((r1s2r2s1)2:).(∗)

Therefore

deg(ϕ+ψ)+ deg (ϕψ)=max( deg (r3), deg (s3))+max( deg (r4), deg (s4))=max( deg (s3s4), deg (r3s4+r3s3), deg (r3r4))2max( deg (r1), deg (s1))+2max( deg (r2), deg (s2))=2deg(ϕ)+2 deg (ψ)(1)

Now replace ϕ,ψ by ϕ+ψ,ϕψ to get

deg(2ϕ)+ deg (2ψ)2deg(ϕ+ψ)+2 deg (ϕψ)4deg(ϕ)+4 deg (ψ)2deg(ϕ+ψ)+2 deg (ϕψ)2deg(ϕ)+2 deg (ψ)deg(ϕ+ψ)+ deg (ϕψ)(2)

(1) and (2) give that deg satisfies the parallelogram law, hence deg is a quadratic form.

This proves Theorem 5.7.

Corollary 5.9. deg(nϕ)=n2 deg (ϕ) for all n, ϕHom(E1,E2). In particular, deg[n]=n2.

Example 5.10. Let EK be an elliptic curve. Suppose charK2. Let 0TE(K)[2].

Without loss of generality E: y2=x(x2+ax+b), a,bK, b(a24b)0, and T=(0,0).

If P=(x,y) and P=P+T=(x,y) then

x=(yx)2ax=x2+ax+bxax=bxy=(yx)x=byx

Let

ξ=x+x+a=(yx)2η=y+y=yx(xbx)η2=(yx)2[(x+bx)24b]=ξ[(ξa)24b]=ξ(ξ22aξ+a24b)

Let E: y2=x(x2+ax+b), where a=2a, b=a24b. There is an isogeny

ϕ:EE(x,y)((yx)2:y(x2b)x2:1)     2        3       0       ord0E()         1           0       3       +30E(0:1:0)=0E

To compute the degree: (yx)2=x2+ax+bx (coprime numerator and denominator as b0). Lemma 5.4 gives degϕ=2.

We say ϕ is a 2-isogeny.

6 The Invariant Differential

Let C be an algebraic curve over K=K¯.

Definition (Space of differentials). The space of differentials ΩC is the K(C)-vector space generated by df for fK(C), subject to the relations

  • (i) d(f+g)=df+dg
  • (ii) d(fg)=fdg+gdf
  • (iii) da=0  aK

Fact: ΩC is a 1-dimensional K(C)-vector space.

Let 0ωΩC. Let PC be a smooth point, and tK(C) a uniformiser at P.

Then ω=fdt for some fK(C). We define ordp(ω)=ordp(f) (independent of choice of t).

We assume C is a smooth projective curve.

Definition. div(ω)=PCordP(ω)PDiv(C).

Note. This is a divisor, i.e. ordp(ω)=0 for all but finitely many PC.

Definition (Regular differential). A differential ωΩC is regular if div(ω)0, i.e. it has no poles.

g(C)=dimK{ωΩC:div(ω)0}.

As a consequence of Riemann-Roch we have:

If 0¬ωΩC, then deg(divω)=2g2.

Fact: Suppose fK(C), ordp(f)=n0. If charKn, then ordp(df)=n1.

Lemma 6.1. Assuming that:

  • charK2

  • E: y2=(xe1)(xe2)(xe3), e1,e2,e3K distinct

Then ω=dxy is a differential on E with no zeroes or poles g(E)=1. In particular, the K-vector space of regular differentials on E is 1-dimensional, spanned by ω.

Proof. Let Ti=(ei,0), E[2]={0,T1,T2,T3}.

div(y)=(T1)+(T2)+(T3)3(0).(1)

For PE{0},

div(xxp)=(P)+(P)2(0).

If PEE[2], then ordp(xxp)=1ordp(dx)=0.

If PEE[2] then ordp(xxp)=1 ordp(dx)=0.

If P=Ti then ordp(xxp)=2 ordp(dx)=1.

If P=0 then ordp(x)=2 ordp(dx)=3.

Therefore

div(dx)=(T1)+(T2)+(T3)3(0).(2)

(1) and (2) implies div(dxy)=0.

Definition. For ϕ:C1C2 a non-constant morphism we define

ϕ:ΩC2ΩC1fdg(ϕf)d(ϕg)

Lemma 6.2. Assuming that:

  • PE,

    τP:EEXP+X

  • ω=dxy as above

Then τPω=ω (we say ω is the invariant differential).

Proof. τPω is a regular differential on E. So τPω=λpω.

The map E1, PλP is a morphism of smooth projective curves but not surjective (misses 0 and ). Therefore it is constant (by Theorem 2.8), i.e. there exists λK such that τPω=λω for all PE.

Taking P=0E shows λ=1.

Remark. If K=, ΛE(), z((z),(z)),

dxy=(z)dz(z)=dz.

(invariant under zz+(const)).

Lemma 6.3. Assuming that:

  • ϕ,ψHom(E1,E2)

  • ω an invariant differential on E2

Then (ϕ+ψ)ω=ϕω+ψω.

Proof. Write E=E2

E×EEμ:(P,Q)P+Qpr1:(P,Q)Ppr2:(P,Q)Q

Fact: ΩE×E is a 2-dimensional K(E×E)-vector space with basis pr1ω and pr2ω.

Therefore

μω=fpr1ω+gpr2ω(1)

for some f,gK(E×E).

For fixed QE, let

ιQ:EE×EP(P,Q)

Applying ιQ to (1) gives

(μιQ)ω=(ιQf)(pr1ιQ)ω+(ιQg)(pr2ιQ)ωτQω=ω=(ιQf)by Lemma 6.2ω+0

Therefore ιQf=1 for all QE, so f(P,Q)=1 for all P,QE.

Similarly g(P,Q)=1 for all P,QE.

(1)μω=pr1ω+pr2ω.

Now pull back by

E1E×EP(ϕ(P),ψ(P))

to get

(ϕ+ψ)ω=ϕω+ψω.

Lemma 6.4. Assuming that:

  • ϕ:C1C2 a non-constant morphism

Then ϕ is separable if and only if ϕ:ΩC1ΩC2 is non-zero

Proof. Omitted.

Example. 𝔾m=𝔸1{0} (multiplicative group).

ϕ:𝔾m𝔾mxxn

(n2 integer).

ϕ(dx)=d(xn)=nxn1dx. So if charKn then ϕ is separable.

Theorem 2.8 implies #ϕ1(Q)=degϕ for all but finitely many Q𝔾m.

But ϕ is a group homomorphism, so

#ϕ1(Q)=#ker(Q)

for all Q𝔾m. Thus #kerϕ= deg ϕ=n and hence K(=K¯) contains exactly n n-th roots of unity.

Theorem 6.5. Assuming that:

  • charKn

Then E[n](n)2.

Proof. Lemma 6.3 + induction gives [n]ω=nω.

charKn implies [n] is separable. So #[n]1(Q)=deg[n] for all but finitely many QE. But [n] is a group homomorphism, so [n]1(Q)=#E[n] for all QE.

Putting these two statements together gives

#E[n]=deg[n]=Corollary 5.9n2.

Group theory (structure theorem) gives that

E[n]d1××dt

for some 1<d1|d2||dt|n.

Let p be a prime with p|d1. Then E[p](p)t. But #E[p]=p2, so t=2. But #E[p]=p2, so t=2, i.e. E[n]d1×d2. Since d1|d2|n and d1d2=n2, we get d1=d2=n.

Thus E[n](n)2.

Remark. If charK=p then [p] is inseparable. It can be shown that:

either E[pr]pr r1“ordinary”or E[p]=0“supersingular”

Do not use this remark on Example Sheet 2!

7 Elliptic Curves over Finite Fields

Lemma 7.1. Assuming that:

  • A an abelian group

  • q:A a positive definit quadratic form

Then
|q(x+y)q(x)q(y)=x,y|2q(x)q(y)x,yA.

Proof. We may assume x0, otherwise the result is clear. So q(x)0.

Let m,n. Then

0q(mx+ny)=12mx+ny,mx+ny=m2q(x)+mnx,y+n2q(y)=q(x)(m+x,y2q(x)n)2+(q(y)x,y24q(x))n2

Take m=x,y, n=2q(x)0 to deduce

q(y)x,y24q(x)0

hence x,y24q(x)q(y) and hence

|x,y|2q(x)q(y).

Theorem 7.2 (Hasse’s Theorem). Assuming that:

Then
|#E(𝔽q)(q+1)|2q.

Proof. Recall Gal(𝔽qr𝔽q) is cyclic of order r and generated by Frobenius xxq.

Let E have Weierstrass equation with coefficients a1,,a6𝔽q (so aiq=ai). Define the Frobenius endomorphism

ϕ:EE(x,y)(xq,yq)

This is an isogeny of degree q.

Then

E(𝔽q)={PE|ϕ(P)=P}=ker(1ϕ).

ϕω=ϕ(dxy)=d(xq)yq=qxq1dxyq=0(q0(modp)) Lemma 6.3 tells us that
(1ϕ)ω=ωϕ(ω)=ω0.

Hence 1ϕ is separable.

By Theorem 2.8 and the fact that 1ϕ is a group homomorphism, we argue as in the proof of Theorem 6.5 that

# ker (1ϕ)=#E(𝔽q)= deg (1ϕ).

deg:Hom(E,E) is a positive definite quadratic form (Theorem 5.7, and positive definiteness is obvious since non-constant morphisms have positive degree).

Lemma 7.1 gives

|deg(1ϕ)1 deg ϕ|2 deg ϕ.

Hence

|#E(𝔽q)1q|2q.

Definition. For ϕ,ψEnd(E)=Hom(E,E), we put ϕ,ψ=deg(ϕ+ψ) deg ϕ deg ψ and tr(ϕ)=ϕ,1.

Corollary 7.3. Assuming that:

Then #E(𝔽q)=q+1tr(ϕ) and |tr(ϕ)|2q.

7.1 Zeta functions

For K a number field, let

ζK(s)=𝔞OK1(N𝔞)s=𝔭OKprime(11(N𝔭)s)1.

For K a function field, i.e. K=𝔽q(C) where C𝔽q is a smooth projective curve,

ζK(s)=x|C|(11(Nx)s)1,

where

|C|={closed points on C}

(closed points are orbits for action of Gal(𝔽q¯𝔽q) on C(𝔽q¯)) and Nx=qdegx, degx is the size of orbit.

We have ζK(s)=F(qs) for some F[[T]],

F(T)=x|C|(1T deg x)1logF(T)=x|C|m=11mTm deg xlog(1x)=m=1xmmTddT(logF(T))=x|C|m=1 deg xTm deg x=n=1(x|C| deg x|n deg x)Tnn=mdegx=n=1#C(𝔽qn)Tn

Therefore

F(T)=exp(n=1#C(𝔽qn)nTn).

Definition (Zeta function). The zeta function ZC(T) of a smooth projective curve C𝔽q is defined by

ZC(T)=exp(n=1#C(𝔽qn)nTn).

Theorem 7.4. Assuming that:

Then
ZE(T)=1aT+qT2(1T)(1qT).

Proof. Let ϕ:EE be the q power Frobenius map. By Corollary 7.3

#E(𝔽q)=q+1tr(ϕ).

Hence tr(ϕ)=a, deg(ϕ)=q.

Example Sheet 2, Q6(iii) implies ϕ2aϕ=q=0, hence ϕn=2aϕn+1+qϕn=0, so

tr(ϕn+2)atr(ϕn+1)+qtr(ϕn)=0.

This second order difference equation with initial conditions tr(1)=2, tr(ϕ)=a has solution

tr(ϕn)=αn+βn,

where α,β are roots of X2aX+q=0.

Again by Corollary 7.3,

#E(𝔽qn)=qn+1tr(ϕn)=1+qnαnβn

Therefore

ZE(T)=expn=1(Tnn+(qT)nn(αT)nn(βT)nn)=(1αT)(1βT)(1T)(1qT)=1aT+qT2(1T)(1qT)

Remark. Hasse’s Theorem tells us that |a|2q. α=β¯, and so |α|=|β|=q ().

Let K=𝔽q(E). ζK(s)=0, so ZE(qs)=0, so qs=α or β. Then qRe(s)=|α| or |β|, so by (), Re(s)=12.

“This is an analog of the Riemann hypothesis.”

8 Formal Groups

Definition (I-adic topology). Let R be a ring and IR an ideal. The I-adic topology on R has basis {r+In|rR,n1}.

Definition (Cauchy sequence). A sequence (xn) in R is Cauchy if k,N such that m,nN, xmxnIk.

Definition (Complete). R is complete if

  • (i) n0In={0}
  • (ii) every Cauchy sequence converges

Useful remark: if xI then

11x=1+x+x2+

so 1xR×.

Example. R=p, I=pp.

R=[[t]], I=(t).

Lemma 8.1 (Hensel’s Lemma). Assuming that:

  • R is complete with respect to an ideal I

  • FR[X], s1

  • aR satisfies F(a)0(modIs), F(a)R×

Then there exists a unique bR such that F(b)=0 and b0(modIs).

Proof. Let uR× with F(a)=u(modI) (e.g. we could take u=F(a)). Replacing F(X) by F(X+a)u, we may assume a=0, F(0)1(modI).

We put x0=0,

xn+1=xnF(xn)(1)

Easy induction gives

xn0(modIs)n(2)

Let

F(X)F(Y)=(XY)(F(0)+XG(X,Y)+YH(X,Y))(3)

for some G,HR[X,Y].

Claim: xn+1xn(modIn+s) for all n0.

Proof: By induction on n. Case n=0 is true.

Suppose xnxn=1(modIn+s1). Then

F(xn)F(xn1)=(xnxn1)(1+c)(modIn+s)

for some cI. Hence

F(xn)F(xn1)=xnxn1(modIn+s)

Hence

xnF(xn)xn1F(xn1)(modIn+s)

and hence xn+1xn(modIn+s).

This proves the claim.

Therefore (xn)n0 is Cauchy. Since R is complete, we have xnb as n for some bR.

Taking limit n in (1) gives b=bF(b), hence F(b)=0.

Taking limit n in (2) gives b0(modIs).

Uniqueness is proved using (3) and the “useful remark” (if xI then 1xR×).

E:Y2Z+a1XYZ+a3YZ2=X3+a2X2Z+a4XZ2+a6Z3.

Affine piece Y0. t=XY, w=ZY.

w=t3+a1tw+a2t2w+a3w2+a4tw2+a6w3=f(t,w).

We apply Lemma 8.1 with

R=[a1,,a6][[t]]I=(t)F(X)=Xf(t,X)R[X]

s=3, a=0.

Check:

F(0)=f(t,0)=t30(modI3)F(0)=1a1ta2t2R×

Hence there exists a unique w(t)R=[a1,,a6][[t]] such that

w(t)=f(t,w(t))w(t)0(modt3)

Remark. Taking u=1 in the proof of Lemma 8.1, w(t)=limnwn(t) where w0(t)=0, wn+1(t)=f(t,wn(t)).

In fact,

w(t)=t3(1+A1t+A2t2+)

where A1=a1, A2=a12+a2, A3=a13+2a1a2+2a3, …

Lemma 8.2. Assuming that:

  • R an integral domain which is complete with respect to an ideal I

  • a1,,a6R

  • K=Frac(R)

Then
Ê(I):={(t,w)E(K)|t,wI}

is a subgroup of E(K).

Note. By uniqueness in Hensel’s lemma:

Ê(I)={(t,w(t))E(K)|tI}.

Proof. Taking (t,w)=(0,0) show 0EÊ(I).

So it suffices to show that P1,P2Ê(I) then

P1P2=P3Ê(I).

PIC

P1,P2Ê(I) implies t1,t2I, w1=w(t1)I, w2=w(t2)I.

w(t)=n=2An2tn+1(A0=1)λ={w(t2)w(t1)t2t1t1t2w(t1)t1=t2=n=2An2(t1n+t1n1t2++t2n)Iν=w1λt1I

Substituting w=λt+ν into w=f(t,w) gives

λt+ν=t3+a1t(λt+ν)+a2t2(λt+ν)+a3(λt+ν)2+a4t(λt+ν)2+a6(λt+ν)3A=(coefficient of t3)=1+a2λ+a4λ2+a6λ3B=(coefficient of t2)=a1λ+a2ν+a3λ2+2a4λν+3a6λ2ν

We have AR× and BI. Hence t3=BAt1t2I, w3=λt3+νI.

Taking R=[a1,,a6][[t]], I=(t), then Lemma 8.2 gives that there exists ι[a1,,a6][[t]] with ι(0)=0 and

[1](t,w(t))=(ι(t),w(ι(t))).

Taking R=[a1,,a6][[t1,t2]], I=(t1,t2), Lemma 8.2 gives that there exists F[a1,,a6][[t1,t2]] with F(0,0)=0 and

(t1,w(t1))+(t2,w(t2))=(F(t1,t2),w(F(t1,t2)))

and

F(X,Y)=X+Ya1XYa2(X2Y+XY2)+

By properties of the group law, we deduce

Definition (Formal group). Let R be a ring. A formal group over R is a power series F(X,Y)R[[X,Y]] satisfying

  • (i)
    F(X,Y)=F(Y,X)
  • (ii)
    F(X,0)=X and F(0,Y)=Y
  • (iii)
    F(X,F(Y,Z))=F(F(X,Y),Z)

This looks like it would only define a monoid, but in fact inverses are guaranteed to exist in this context.

Exercise: Show that for any formal group, there exists a unique

ι(X)=X+R[[X]]

such that F(X,ι(X))=0.

Example.

Definition (Morphism / isomorphic (formal groups)). Let F and G be formal groups over R given by power series F and G.

  • (i)
    A morphism f:FG is a power series fR[[T]] such that f(0)=0 satisfying
    f(F(X,Y))=G(f(x),f(Y)).

  • (ii)
    FG if there exists FfG and GgF morphisms such that f(g(X))=g(f(X))=X.

Theorem 8.3 (All formal groups are isomorphic). Assuming that:

  • charR=0

Then any formal group F over R is isomorphic to 𝔾^a over R. More precisely
  • (i)
    There is a unique power series
    log(T)=T+a22T2+a33T3+

    with aiR such that

    log(F(X,Y))= log (X)+ log (Y).(∗)

  • (ii)
    There is a unique power series
    exp(T)=T+b22!+b33!T3+

    with biR such that

    exp( log (T))= log ( exp (T))=T.

Proof.

Lemma 8.4. Assuming that:

  • f(T)=aT+R[[T]] with aR×

Then there exists a unique g(T)=a1T+R[[T]] such that f(g(T))=g(f(T))=T.

Proof. We construct polynomials gn(T)R[T] such that

f(gn(T))T(modTn+1)gn+1(T)gn(T)(modTn+1)

Then g(T)=limngn(T) satisfies f(g(T))=T.

To start the induction, we set g1(T)=a1T. Now suppose n2 and gn1(T) exists. Then

f(gn1(T))T+bTn(modTn+1).

We put gn(T)=gn1(T)+λTn for some λR to be chosen later.

Then

f(gn(T))=f(gn1(T)+λTn)f(gn1(T))+λaTn(modTn+1)T+(b+λa)Tn(modTn+1)

We take λ=ba (aR×,bRλR).

This completes the induction step.

We get g(T)=a1T+R[[T]] such that f(g(T))=T. Applying the same construction to g gives h(T)=aT+R[[T]] such that g(h(T))=T.

Now note f(T)=f(g(h(T)))=h(T).

Theorem 8.3(ii) now follows by Lemma 8.4 and Q12 from Example Sheet 2.

Notation. Let F (e.g. 𝔾^a, 𝔾^m, Ê) be a formal group given by a power series FR[[X,Y]].

Suppose R is a ring complete with respect to ideal I. For x,yI, put

xFy=F(x,y)I.

Then F(I):=(I,F) is an abelian group.

Examples:

Corollary 8.5. Assuming that:

Then
  • (i) [n]:FF is an isomorphism of formal groups
  • (ii) If R is complete with respect to ideal I then F(I)×nF(I) is an isomorphism of groups. In particular, F(I) has no n-torsion.

Proof. We have

[1](T)=T[n](T)=F([n1](T),T)

(for n<0 use [1](T)=ι(T)).

Since

F(X,Y)=X+Y+XY()

we get

[2](T)=F(T,T)2T+R[[T]]

and by induction we get

[n](T)=nT+R[[T]].

Lemma 8.4 shows that if nR× then [n] is an isomorphism. This proves (i), and (ii) follows.

9 Elliptic Curves over Local Fields

Let K be a field, complete with respect to discrete valuation v:K.

Notation. Valuation ring (= ring of integers) will be denoted by

OK={xK|v(x)0}{0}.

Unit group will be denoted by

OK={xK|v(x)=0}.

The maximal ideal will be denoted by πOK where v(π)=1.

The residue field will be denoted by k=OKπOK.

We assume charK=0 and chark=p>0. For example, K=p, OK=p, k=𝔽p.

Let EK be an elliptic curve.

Definition (Integral / minimal Weierstrass equation). A Weierstrass equation for E with coefficients a1,,a6K is integral if a1,,a6OK and minimal if v(Δ) is minimal among all integral Weierstrass equations for E.

Remark.

Lemma 9.1. Assuming that:

Then either x,yOK or
{v(x)=2sv(y)=3s

for some s1.

(Compare with Q5 from Example Sheet 1)

Proof. Throughout this proof, LHS and RHS refer to the Weierstrass equation of the curve.

Case v(x)0:. If v(y)<0 then v(LHS)<0 and v(RHS)0. Therefore x,yOK.

Case v(x)<0: v(LHS)min(2v(y),v(x)+v(y),v(y)) and v(RHS)=3v(x). We get 3 possible inequalities from this, and each of them gives v(y)<v(x).

Now

v(LHS)=2v(y)v(RHS)=3v(x)

so

{v(x)=2sv(y)=3s

for some s1.

If K is complete, then OK is complete with respect to πrOK (for any r1).

We fix a minimal Weierstrass equation for EK. Get formal group Ê over OK.

Taking I=πrOK (with r1) in Lemma 8.2 gives

Ê(I)={(x,y)E(K) | xy,1yπrOK}{0}={(x,y)E(K) | v(xy)r,v(1y)r}{0}={(x,y)E(K)|v(x)=2s,v(y)=3s for sr}{0}(using Lemma 9.1)={(x,y)E(K)|v(x)2r,v(y)3r}{0}(using Lemma 9.1)

By Lemma 8.2 this is a subgroup of E(K), say Er(K).

E2(K)E1(K)E(K).

More generally, for F a formal group over OK

F(π2OK)F(πOK).

We claim that

Reminder: charK=0, chark=p>0.

Theorem 9.2. Assuming that:

  • K a local field with charK=0, chark=p>0

  • F a formal group over OK

  • e=v(p)

  • r>ep1

Then
log:F(πrOK) 𝔾^a(πrOK)

is an isomorphism of groups with inverse

exp:𝔾^a(πrOK) F(πrOK).

Remark. 𝔾^a(πrOK)=(πrOK,+)(OK,+), xπrx.

Proof. For xπrOK we must show the power series log(x) and exp(x) converge to elements in πrOK.

Recall

exp(T)=T+b22!T2+b33!T3+

for some biOK.

Claim: vp(n!)n1p1.

Proof of claim:

vp(n!)=r=1npr<r=1npr=n1p11p=np1.

Therefore

(p1)vp(n!)<n(p1)vp(n!)n1

(we go from < to 1 by noting that the LHS is in ).

This proves the claim.

Now

v(bnxnn!)nre(n1p1)=(n1)(rep1)>0+r

This is always r and as n. Therefore exp(x) converges to an element in πrOK.

Same method works for log.

Lemma 9.3. F(πrOK)F(πr+1OK) (k,+) for all r1.

Proof. Definition of formal group gives

F(X,Y)=X+Y+XY().

So if x,yOK,

F(πrx,πry)πr(x+y)(modπr+1).

Therefore

F(πrOK)(k,+)πrxxmodπ

is a surjective group homomorphism with kernel F(πr+1OK).

Corollary. Assuming that:

  • |k|<

Then F(πOK) has a subgroup of finite index isomorphic to (OK,+).

Notation. Reduction modulo π

OKOKπOK=kxx~

Proposition 9.4. Assuming that:

Then the reduction mod π of any two minimal Weierstrass equations for E define isomorphic curves over k.

Proof. Say Weierstrass equations are related by [u;r,s,t], uK, r,s,tK. Then Δ1=u12Δ2. Both equations minimal gives us that v(Δ1)=v(Δ2), hence uOK.

Transformation formula for the ai and bi+OK is integrally closed, hence r,s,tOK.

The Weierstrass equations obtained by reducing mod π are now related by [ũ;r~,s~,t~], ũk, r~,s~,t~k.

Definition. The reduction k of EK is defined by the reduction of a minimal Weierstrass equation.

E has good reduction if is non-singular (and so an elliptic curve) otherwise has bad reduction.

For an integral Weierstrass equation,

  • If v(Δ)=0 then good reduction.

  • If 0<v(Δ)<12 then bad reduction.

  • If v(Δ)12 then beware that the equation might not be minimal.

There is a well-defined map

2(K)2(k)(x:y:z)(x~::z~)

(choose a representative with min(v(x),v(y),v(z))=0).

We restrict to give

E(K)(K)PP~

If P=(x,y)E(K) then by Lemma 9.1 either

Therefore

E1(K)=Ididntmanagetowritethisbeforeitwasrubbedoff

“kernel of reduction”.

Notation.

{if E has good reduction{singular point}if E has bad reduction

The chord and tangent process still defines a group law on ns.

In cases of bad reduction, ns𝔾m (over k or possibly a quadratic extension of k) or ns𝔾a (over k).

For simplicity we suppose chark2.

Then =y2=f(x), degf=3.

PIC

ns 𝔾a(x,y)xy(t2,t3)t(point at )0 Let P1,P2,P3 lie on the line ax+by=1. Write Pi=(xi,yi), ti=xiyi. Then xi3=yi2=yi2(axi+byi). So ti3atib=0. So t1,t2,t3 are the roots of T3aTb=0.

Looking at coefficient of T2 gives t1+t2+t3=0.

Definition (E0(K)). E0(K)={PE(K)|P~ns(k)}.

Proposition 9.5. E0(K) is a subgroup of E(K) and reduction modulo π is a surjective group homomorphism E0(K)ns(k).

Note. If EK has good reduction, then this is a surjective group homomorphism E(K)(k).

Proof. Group homomorphism: A line l in 2 defined over K has equation

l:aX+bY+cZ=0a,b,cK.

We may assume min(v(a),v(b),v(c))=0. Reduction modulo π gives a line

l~:ãX+b~Y+c~Z=0.

If P1,P2,P3E(K) with P1+P2+P3=0 then these points lie on a line l. So P~1,P~2,P~3 lie on the line l~. If P~1,P~2(~nsk) then P~3(~nsk).

So if P1,P2E0(K) then P3E0(K) and P~1+P~2+P~3=0.

[Exercise: check this still works if #{P~1,P~2,P~3}<]

Surjective: Let f(x,y)=y2+a1xy+a3y(x3+). Let P~ns(k){0}, say P~=(x~0,0) for some x0,y0OK.

Since P~ non-singular, either:

If (i) then put g(t)=f(t,y0)OK[t]. Then

{g(x0)0(modπ)g(x0)OK

Hensel’s lemma gives us that there exists bOK such that

{g(b)=0bx0(modπ)

Then (b,y0)E(K) has erduction P~. asdfadsf

Recall that for r1 we put

Er(K)={(x,y)|v(x)2r,v(y)3r}{0}.

If r>ep1, these give:

Er(K)(OK,+)E2(K)=Ê(π2OK)E1(K)=Ê(πOK)E0(K)E(K),

where for i1, each Ei+1(K)Ei(K) gives a quotient isomorphic to (k,+).

We have E0(K)E1(K)ns(k).

What about E(K)E0(K)?

Lemma 9.6. Assuming that:

  • |k|<

Then E0(K)E(K) has finite index.

Proof. |k|< implies that OKπrOK is finite for all r1.

Hence

OKlimr[]OKπrOK

is a profinite group, hence compact.

Then n(K) is the union of sets

{(a0::ai1:1:ai+1::an):ajOK}

and hence compact (for the π-adic topology).

Now note E(K)2(K) is a closed subset, hence compact.

So E(K) is a compact topological group.

If has a singular point (x~0,0) then

E(K)E0(K)={(x,y)E(K)|v(xx0)1,v(yy0)1}

is a closed subset of E(K) hence E0(K) is an open subgroup of E(K).

The cosets of E0(K) are an open cover of E(K). Hence [E(K):E0(K)]<.

Definition (Tamagawa number). cK(E)=[E(K):E0(K)] is called the Tamagawa number.

Remark.

We deduce:

Theorem 9.7. Assuming that:

  • [K:p]<

Then E(K) contains a subgroup of finite index isomorphic to (OK,+).

Let [K:p]< and LK a finite extension. Let the residue fields be k and k, and let f=[k:k].

    ∗
   K       ℤ

V⊂×Ve L∗      ℤ
 KL

Facts:

Definition (Unramified). LK is unramified if e=1.

Fact: For each m1

These extensions are Galois, with cyclic Galois groups.

Definition (Maximal unramified extension). Knr=m1Km (inside K¯). “maximal unramified extension”

Theorem 9.8. Assuming that:

Then K([n]1P)K is unramified.

Notation.

[n]1(P)={QE(K¯):nQ=P},
K({Q1,,Qr})=K(x1,,xr,y1,,yr),

where Qi=(xi,yi).

Proof. For each m1 there is a short exact sequence

0E1(Km)E(Km)(Km)0.

Taking m1 gives a commutative diagram with exact rows:

                                     --
0          E1(Knr)     E (Knr)     ˜E(k)    0


0          E1(Knr)     E (Knr)     ˜E(k)    0

PIC

An isomorphism by Corollary 8.5 applied over each Km (using pn here).

Snake lemma gives

{E(Knr)[n](n)2E(Knr)nE(Knr)=0

So if PE(K) then there exists QE(Knr) such that nQ=P and

[n]1P={Q+T:TE[n]}E(Knr).

Hence K([n]1P)Knr and so K([n]1P)K is unramified.

10 Elliptic Curves over Number Fields: The torsion subgroup

[K:]<, EK an elliptic curve.

Notation. 𝔭 a prime of K (i.e. a prime) ideal in OK).

K𝔭 is the 𝔭-adic completion of K, valuation ring O𝔭.

k𝔭=OK𝔭 residue field.

Definition (Good reduction (prime)). 𝔭 is a prime of good reduction for EK if EK𝔭 has good reduction.

Lemma 10.1. EK has only finitely many primes of bad reduction.

Proof. Take a Weierstrass equation for E with a1,,a6OK. E non-singular implies that 0ΔOK.

Write (Δ)=𝔭α1𝔭rαr (factorisation into prime ideals).

Let S={𝔭1,,𝔭r}. If 𝔭S then v𝔭(Δ)=0. Hence EK𝔭 has good reduction.

Therefore {bad primes for E}S, hence is finite.

Remark. If K has class number 1 (e.g. K=) then we can always find a Weierstrass equation for E with a1,,a6OK which is minimal at all primes 𝔭.

Basic group theory: If A is a finitely generatead abelian group then A{finite group}×r. We call r the “rank”, and {finite subgroup} is the torsion subgroup.

Lemma 10.2. E(K)tors is finite.

Proof. Take any prime 𝔭. We saw that E(K𝔭) has a finite index subgroup A (say) with A(O𝔭,+).

In particular, A is torsion free

E(K)torsE(K𝔭)torsE(K𝔭)Afinite.

Lemma 10.3. Assuming that:

Then reduction modulo 𝔭 gives an injective group homomorphism
E(K)[n](k𝔭).

Proof. Proposition 9.5 gives that E(K𝔭)(k𝔭) is a group homomorphism, with kernel E1(K𝔭).

Corollary 8.5 and 𝔭n gives that E1(K𝔭) has no n-torsion.

Example. E: y2+y=x3x, Δ=11.

E has good reduction at all p11.

p 2 3 5 7 11 13







#(𝔽p) 5 5 5 5 10

Lemma 10.3 gives:

{#()tors|52afor some a0#E()tors|53bfor some b0

Hence #()tors|5.

Let T=(0,0)E(). Calculation gives 5T=0. Therefore E()tors5.

Example. E: y2+y=x3+x2, Δ=43

p 2 3 5 7 11 13







#(𝔽p) 5 6 10 8 9 19

Lemma 10.3 gives:

{#E()tors|52afor some a0#E()tors|911bfor some b0

Therefore E()tors={0}.

Therefore P=(0,0) is a point of infinite order.

In particular, E() is infinite.

Example. ED: y2=x3D2x=f(x). D square-free, Δ=26D6. If p2D, then

#D(𝔽p)=1+x𝔽p((f(x)p)+1).

If p3(mod4) then since f is odd,

(f(x)p)=(f(x)p)=(1p)(f(x)p)=(f(x)p).

Hence

#D(𝔽p)=p+1.
ED()tors{0,(0,0),(±D,0)}(2)2.

Let m=#ED()tors.

We have 4|m|p+1 forr all sufficiently large (p2mD) primes p with p3(mod4). Hence m=4 (otherwise get contradiction to Dirichlet’s theorem on primes on arithmetic progressions).

So

rankED()1x,y,y0y2=x3D2xLecture 1D is a congruent number

Lemma 10.4. Assuming that:

Then
  • (i) 4x,8y
  • (ii) If 2|a1 or 2T0 then x,y

Proof.

Example. y2+xy=x3+4x+1, (14,18)E()[2].

Theorem 10.5 (Lutz Nagell). Assuming that:

  • E: y2=x3+ax+b, a,b

  • 0T=(x,y)E()tors

Then x,y and either y=0 or y2|(4a3+27b2).

Proof. Lemma 10.4 gives that x,y.

If 2T=0 then y=0. Otherwise 02T=(x2,y2)E()tors.

Lemma 10.4 gives x2,y2. But

x2=(f(x)2y)22x

hence y|f(x).

E non-singular gives that f(X) adn f(X) are coprime, so f(X) and f(X)2 are coprime. Therefore there exists g,h[X] satisfying

g(X)f(X)+h(X)f(X)2=1.

Doing this and clearing denominators gives

(3X2+4a)f(X)227(X3+aXb)f(X)=4a3+27b2.

Since y|f(x) and y2=f(x), we get y2|(4a3+27b2).

Remark. Mazur showed that if E is an elliptic curve then

E()tors{n1n12,n112×2n1n4

Moreover, all 15 possibilities occur.

11 Kummer Theory

Let K be a field and charKn. Assume μnK.

Lemma 11.1. Assuming that:

  • ΔK(K)n a finite subgroup

  • L=K(Δn)

Then LK is Galois and Gal(LK)Hom(Δ,μn).

Proof. μnK gives LK normal, and charKn gives that LK is separable. So LK is Galois.

Define the Kummer pairing

,:Gal(LK)×Δμn(σ,x)σ(xn)xn

Well-defined: Suppose α,βL with αn=βn=x. Then (αβ)n=1, so αβμnK. Then σ(αβ)=αβ for all σGal(LK), hence σ(α)α=σ(β)β for all σGal(LK).

Bilinear:

στ,x=σ(τxn)τxnτxnxn=σ,xτ,xσ,xy=σxynxyn=σxnxnσynyn=σ,xσ,y

Non-degenerate: Let σGal(LK). If σ,x=1 for all xΔ, then

σxn=xnxΔ

adn hence σ fixes L pointwise, i.e. σ=1.

Let x(K)nΔ. If σ,x=1 for all σGal(LK), then

σxn=xnσGal(LK).

So xnK, so x(K)n, i.e. x(K)nΔ is the identity element.

We get injective group homomorphisms

(i) implies Gal(LK) is abelian and of exponent dividing n.

Fact: If G is a finite abelian group of exponent dividing n then Hom(G,μn)G (non canonically).

So

|Gal(LK)|(i)|Δ|(ii)|Gal(LK)|.

Therefore (i) and (ii) are isomorphisms.

Example. Gal((2,3,5))(2)3.

Definition (Abelian extension). We say LK is abelian if it is Galois, and has abelian Galois group.

Similarly for other group terminology (e.g. we can say that LK has exponent dividing n to mean that it is Galois, with Galois group having exponent dividing n).

Theorem 11.2. There is a bijection

{finite subgroupsΔK(K)n}{finite abelian extension LKof exponent dividing n}ΔK(Δn)(L)nK(K)nL

Proof.

Proposition 11.3. Assuming: - K a number field - μnK - S a finite set of promes of K Then there are only finitely many extensions LK such that

  • (i) LK is a finite abelian extension of exponent dividing n
  • (ii) LK is unramified at all 𝔭S

Proof. Theorem 11.2 gives L=K(Δn) for some ΔK(K)n a finite subgroup.

Let 𝔭 be a prime of K.

𝔭OL=P1e1Prer

for P1,,Pr some distinct primes in OL.

If xK represents an element of Δ then

nvPi(xn)=vPi(x)=eiv𝔭(x).

If 𝔭S then all ei=1, so v𝔭(x)0(modn). Therefore ΔK(S,n) where

K(S,n)={xK(K)n|v𝔭(x)0(modn) 𝔭S}.

The proof is completed by the next lemma.

Lemma 11.4. K(S,n) is finite.

Proof. The map

K(S,n)(n)|S|x(v𝔭(x)(modn))𝔭S

is a group homomorphism with kernel K(,n).

Since |S|<, it suffices to prove the lemma with S=.

If xK represents an element of K(,n) then (x)=𝔞n for some fractional ideal 𝔞. There is a short exact sequence

0OK(OK)nK(,n)ClK0x(K)n[𝔞]

|ClK|< and OK being a finitely generated abelian group (Dirichlet’s unit theorem) gives us that K(ϕ,n) is finite.

12 Elliptic Curves over Number Fields: The weak Mordell-Weil Theorem

Theorem 12.1. Assuming that:

Then the natural map E(K)nE(K)E(L)nE(L) has finite kernel.

Proof. For each element in the kernel we pick a coset representative PE(K) and then QE(L) such that nQ=P.

For any σGal(LK) we have

n(σQQ)=σPP=0.

So σQQE[n].

Since Gal(LK) and E[n] are finite, there are only finitely many possibilities for the map

Gal(LK)E[n]σσQQ

(even without requiring it to be a group homomorphism!).

So we have a map

ker(E(K)nE(K)E(L)nE(L))()Maps(Gal(LL),E[n])P+nE(K)(σσQQ)where nQ=P

It remains to show () is injective.

So suppose P1,P2E(K), Pi=nQi for i=1,2, and suppose σQ1Q1=σQ2Q2 for all σGal(LK). Then σ(Q1Q2)=Q1Q2 for all σGal(LK), hence Q1Q2E(K), so P1P2nE(K). Hence P1+nE(K)=P2+nE(K) as desired.

Theorem (Weak Mordell-Weil Theorem). Assuming that:

Then E(K)nE(K) is finite.

Proof. Theorem 12.1 tells us that we may replace K by a finite Galois extension.

So without loss of generality μnK and E[n]E(K). Let

S={𝔭|n}{primes of bad reduction for EK}.

For each PE(K), the extension K([n]1P)K is unramified outisde S by Theorem 9.8. Since Gal(K¯K) acts on [n]1P, it follows that Gal(K¯K([n]1P)) is a normal subgroup of Gal(K¯K) and hence K([n]1P)K is a Galois extension.

Let Q[n]1P. Since E[n]E(K), we have K(Q)=K([n]1P). Consider

Gal(K(Q)K)E[n](n)2σσQQ

Group homomorphism: στQQ=σ(τQQ)+(σQQ).

Injective: If σQ=Q then σ fixes K(Q) pointwise, i.e. σ=1.

Therefore K(Q)K is an abelian extension of exponent n, unramified outside S.

Proposition 11.3 shows that as we vary PE(K) there are only finitely many possibilities for K(Q).

Let L be the composite of all such extensions of K. Then LK is finite and Galois, and E(K)nE(K)E(L)nE(L) is the zero map. Theorem 12.1 gives |E(K)nE(K)|<.

Remark. If K= or , or [K:p]< then |E(K)nE(K)|<, yet E(K) is uncountable, so not finitely generated.

Fact: If K is a number field, then there exists a quadratic form (= canonical height) ĥ:E(K)0 with the property that for any B0,

{PE(K)|ĥ(P)B}(∗)

is finite.

Theorem (Mordell-Weil Theorem). Assuming that:

Then E(K) is a finitely generated abelian group.

Proof. Fix an integer n2.

Weak Mordell-Weil Theorem implies that |E(K)nE(K)|. Pick coset representatives P1,P2,,Pm. Let

Σ={PE(K)|(P)^max1imĥ(Pi)}.

Claim: Σ generates E(K).

If not, then there exists PE(K)(subgroup generated by Σ) of minimal height (exists by ()). Then P=Pi+nQ for some i and QE(K). Note that QE(K)(subgroup generated by Σ).

Minimal choice of P gives

4ĥ(P)4ĥ(Q)n2ĥ(Q)=ĥ(nQ)=ĥ(PPi)ĥ(PPi)+ĥ(P+Pi)=2ĥ(P)+2ĥ(Pi)parallelogram law

Therefore ĥ(P)ĥ(Pi). Hence PΣ (by definition of Σ), which contradicts the choice of P.

This proves the claim.

By (), Σ is finite.

13 Heights

For simplicity, take K=. Write Pn(Q) as P=(a0:a1::an) where a0,,an and gcd(a0,an)=1.

Definition (Height of a point). H(P)=max0in|ai|.

Lemma 13.1. Assuming that:

  • f1,f2[X1,X2] coprime homogeneous polynomials of degree d

  • let

    F:11(x1:x2)(f1(x1,x2):f2(x1,x2))

Then there exist c1,c2>0 such that for all P1(),
c1H(P)dH(F(P))c2H(P)d.

Proof. Without loss of generality f1,f2[X1,X2].

Upper bound: Write P=(a1:a2), a1,a2 coprime.

H(F(P))max(|f1(a1,a2)|,|f2(a1,a2)|)c2max(|a1|d,|a2|d)

where

c2=maxi=1,2(sum of absolute values of coefficients of fi).

Therefore H(F(P))c2H(P)d.

Lower bound: We claim that there exists gij[X1,X2] homogeneous of degree d1 and κ>0 such that

j=12gijfj=κXi2d1,i=1,2.(†)

Indeed, running Euclid’s algorithm on f1(X,1) and f2(X,1) gives r,s[X] of degree <d such that

r(X)f1(X,1)+s(X)f2(X,1)=1.

Homogenising and clearing denominators gives () for i=2. Likewise for i=1.

Write P=(a1:a2) with a1,a2 coprime. () gives

j=12gij(a1,a2)fj(a1,a2)=κai2d1,i=1,2.

Therefore gcd(f1(a1,a2),f2(a1,a2)) divides gcd(κa12d1,κa22d1)=κ.

|κai2d1|maxj=1,2|fj(a1,a2)|κH(F(P))j=12|gij(a1,a2)|γiH(D)d1

where

γi=j=12(sum of absolute values of coefficients of gij).

Therefore

κ|ai|2d1κH(F(P))γiH(P)d1

so

H(P)2d1max(γ1,γ2)H(F(P))H(P)d1

so

1max(γ1,γ2)=c1H(P)dH(F(P)).

Notation. For x, H(X)=H((x:1))=max(|u|,|v|), where x=uv, u,v coprime.

Definition (Height of a point). Let E be an elliptic curve, y2=x3+ax+b.

Define the height

H:E()1P{H(x)if P=(x,y)1if P=0E

Alsdefine logarithmic height

h:E()0PlogH(P)

Lemma 13.2. Assuming that:

Then there exists c>0 such that
|h(ϕ(P))(degϕ)h(P)|cPE().

Note. c depends on E and E, but not P.

Proof. Recall (Lemma 5.4)

  E       E ′


ϕxxξ ℙ1      ℙ1
degϕ= deg ξ (=d say). Lemma 13.1 tells us that there exists c1,c2>0 such that

c1H(P)dH(ϕ(P))c2H(P)dPE().

Taking logs gives

|h(ϕ(P))dh(P)|max( log c2 log c1)=c

Example. ϕ=[2]:EE. Then there exists c>0 such that

|h(2P)4h(P)|cPE().

Definition (Canonical height of a point). The canonical height is

ĥ(P)=limn14nh(2nP).

We check convergence:

Let mn. Then

|14mh(2mP)14nh(2nP)|r=nm1|14r+1h(2r+1P)14rh(2rP)|=r=nm114r+1|h(2(2rP))4h(2rP)|<cr=n14r+1=c4n+11114=c3×4n0

as n. So the sequence is Cauchy, ĥ(P) exists.

Lemma 13.3. |h(P)ĥ(P)| is bounded for PE().

Proof. Put n=0 in above calculation to get

|14mh(2mP)h(P)|<c3.

Take limit m.

Lemma 13.4. Assuming that:

  • B>0

Then
#{PE()|ĥ(P)B}<.

Proof. ĥ(P) bounded means we have a bound on h(P) (by Lemma 13.3). So only finitely many possibilities for x. Each x gives 2 choices for y.

Lemma 13.5. Assuming that:

  • ϕ:EE an isogeny defined over

Then
ĥ(ϕP)=(degϕ)ĥ(P)PE().

Proof. By ?? 66, there exists c>0 such that

|h(ϕP)(degϕ)h(P)|<cPE().

Replace P by 2nP, divide by 4n and take limit n.

Remark.

Lemma 13.6. Assuming that:

Then there exists c>0 such that for all P,QE() with P,Q,P+Q,PQ0, we have
H(P+Q)H(PQ)cH(P)2H(Q)2.

Proof. Let E have Weierstrass equation y2=x3+ax+b, a,b. Let P,Q,P+Q,PQ have x coordinates x1,,x4. By Lemma 5.8, there exists W0,W1,W2[x1,x2] of degree 2 in x2 such that

(1:x3+x4:x3x4)=(W0=(x1x2)2:W1:W2).

Write xi=risi with ri,si coprime.

(s3s4:r3s4+r4s3:r3r4)gcd=1=((r1s2r2s1)2:)H(P+Q)H(PQ)=max(|r3|,|s3|)max(|r4|,|s4|)2max(|s3s4|,|r3s4+r4s3|,|r3r4|)2max(|r1s2r2s1|2,)cmax(|r1|2,|s1|)2max(|r2|,|s2|)2=cH(P)2H(Q)2

where c depends on E, but not on P and Q.

Theorem 13.7. ĥ:E()0 is a quadratic form.

Proof. Lemma 13.6 and |h(2P)4h(P)| bounded gives that there exists c such that

h(P+Q)+h(PQ)2h(P)+2h(Q)+cP,QE().

Replacing P, Q by 2nP, 2nQ, dividing by 4n and taking the limit n gives

ĥ(P+Q)+ĥ(PQ)2ĥ(P)+2ĥ(Q).

Replacing P,Q by P+Q, PQ and ĥ(2P)=4ĥ(P) gives the reverse inequality. Therefore ĥ satisfies the parallelogram law, and hence ĥ is a quadratic form.

Remark. For K a number field and P=(a0:a1::an)n(K), define

H(P)=vmax0in|ai|v

where the product is over all places v, and the absolute values are normalised such that

v|λ|v=1λK.

Using this definition, all results in this section generalise when is replaced by a number field K.

14 Dual isogenies and the Weil pairing

Let K be a perfect field and EK an elliptic curve.

Proposition 14.1. Assuming that:

  • ΦE(K¯) be a finite Gal(K¯K)-stable subgroup

Then there exists an elliptic curve EK and a separable isogeny ϕ:EE defined over K, with kernel Φ, such that every isogeny ψ:EE with Φψ factors uniquely via ϕ:
 E                E′′

           ′
ψϕ∃ unique  E

Proof. Omitted (see Silverman, Chapter III, Proposition 4.12).

Proposition 14.2. Assuming that:

  • ϕ:EE an isogeny of degree n

Then there exists a unique isogeny ϕ^:EE such that ϕ^ϕ=[n].

Proof. Case ϕ is separable: We have |kerϕ|=n, hence kerϕE[n]. Apply Proposition 14.1 with ψ=[n].

Case ϕ is inseparable: omitted.

Uniqueness: Suppose ψ1ϕ=ψ2ϕ=[n]. Then (ψ1ψ2)ϕ=0, so deg(ψ1ψ2) deg ϕ=0. Then deg(ψ1ψ2)=0, hence ψ1=ψ2.

Remark.

Lemma 14.3. Assuming that:

  • ϕ,ψHom(E,E)

Then ϕ+ψ^=ϕ^+ψ^.

Proof.

Remark. In Silverman’s book he proves Lemma 14.3first, and uses this to show deg:Hom(E,E) is a quadratic form.

Notation.

sum:Div(E)Enp(P)formal sumnP(P)addusinggrouplaw

Recall

E Pic0(E)P[(P)(0)]

Therefore sumD[D] for all DDiv0(E).

We deduce:

Lemma 14.4. Assuming that:

  • DDiv(E)

Then D0 if and only if degD=0 and sumD=0E.

We will now discuss Weil pairing.

Let ϕ:EE be an isogeny of degree n, with Dual isogeny ϕ^:EE. Assume charKn (hence ϕ, ϕ¯ separable).

We define the Weil pairing

eϕ:E[ϕ]×E[ϕ^]μn.

Let TE[ϕ]. Then nT=0, so there exists fK¯(E) such that

div(f)=n(T)n(0).

Pick T0E(K¯) with ϕ(T0)=T. Then

ϕ(T)ϕ(0)=PE[ϕ](P+T0)PE[ϕ](P)

has sum

nT0=ϕ^ϕT0=ϕ^T=0.

So there exists gK¯(E) such that

div(g)=ϕ(T)ϕ(0).

Now

div(ϕf)=ϕ(divf)=n(ϕ(T)ϕ(0))=div(gn)

Therefore ϕf=cgn for some cK¯.

Rescaling f, we can say without loss of generality c=1, i.e. ϕf=gn.

For SE[ϕ] we get τS(divg)=(divg) so τSg=ζg for some ζK¯, i.e. ζ=g(X+S)g(X) is independent of choice of XE(K¯).

Now

ζn=g(X+S)ng(X)n=f(ϕ(X+S))f(ϕ(X))=1

since SE[ϕ]. Hence ζμn.

We define

eϕ(S,T)=g(X+S)g(X).

Proposition 14.5. e is bilinear and non-degenerate.

Proof.

We’ve shown E[ϕ]Hom(E[ϕ],μn). It is an isomorphism since #E[ϕ]=#E[ϕ^]=n.

Remark.

Corollary 14.6. Assuming that:

  • E[n]E(K)

Then μnK.

Proof. Let TE[n] have order n. Non degeneracy of en implies that there exists SE[n] such that en(S,T) is a primitive n-th root of unity, say ζn.

Then

σ(ζn)=σ(en(S,T))=en(σS,σT)=ζn

for all σGal(K¯K). So ζnK.

Example. There does not exist E with E()tors(3)2.

Remark. In fact, en is alternating, i.e. en(T,T)=1 for all TE[n].

(This implies en(S,T)=en(T,S)1).

15 Galois Cohomology

Let G be a group and A be a G-module (an abelian group with an action of G via group homomorphisms). G-module means exactly the same thing as [G]-module.

Definition (H0). Define

H0(G,A)=AG={aA|σ(a)=a σG}.

Definition (Cochains). Define

C1(G,A)={maps GA}

(called “cochains”).

Definition (Cocycles). Define

Z1(G,A)={(aσ)σG|aστ=σ(aτ)+aσ σ,τG}

(called “cocycles”).

Definition (Coboundaries). Define

B1(G,A)={(σbb)σG|bA}

(called “coboundaries”).

Note. C1(G,A)Z1(G,A)B1(G,A).

Then we can define:

Definition (H1). Define

H1(G,A)=Z1(G,A)B1(G,A).

Remark. If G acts trivially on A, then

H1(G,A)=Hom(G,A).

Theorem 15.1. Assuming that:

  • we have a short exact sequence of G-modules

    0AϕBψC0.

Then it gives rise to a long exact sequence of abelian groups:
0AGϕBGψCGδH1(G,A)ϕH1(G,B)ψH1(G,C).

Proof. Omitted.

Definition (delta). Let cCG. Then bB such that ψ(b)=c. Then

ψ(σbb)=σcc=0σG

so σbb=ϕ(aσ) for some aσA.

Can check (aσ)σGZ1(G,A).

Then δ(c)= class of (aσ)σG in H1(G,A).

Theorem 15.2. Assuming that:

  • A is a G-module

  • HG a normal subgroup

Then there is an inflation restriction exact sequence
0H1(GH,AH)infH1(G,A)resH1(H,A)

Proof. Omitted.

Let K be a perfect field. Then Gal(K¯K) is a topological group with basis of open subgroups being the Gal(K¯L) for [L:K]<.

If G=Gal(K¯K) then we modify the definition of H1(G,A) by insisting:

Then

H1(Gal(K¯K),A)=limLLK finite GaloisH1(Gal(LK),AGal(K¯L)).

(direct limit is with respect to inflation maps).

Theorem (Hilbert’s Theorem 90). Assuming that:

  • LK a finite Galois extension

Then
H1(Gal(LK),L)=0.

Proof. Let G=Gal(LK). Let (aσ)σGZ1(G,L). Distinct automorphisms are linearly independent, so there exists yL such that

τGaτ1τ(y)=x0.

Then

σ(x)=τGσ(aτ)1στ(y)=aστGaστ1στ(y)=x

Then aσ=σ(x)x for all σG. (aστ=σ(aτ)aσ σ(aτ)1=aσaστ1).

So (aσ)σGB1(G,L). Therefore H1(G,L)=0.

Corollary. H1(Gal(K¯K),K¯)=0.

Application: Assume charKn. There is a short exact sequence of Gal(K¯K)-modules

0μnK¯K¯0xxn

Long exact sequence:

KKH1(Gal(K¯K),μn)H1(Gal(K¯K),K¯)=0 by Hilbert 90xxn

Therefore H1(Gal(K¯K),μn)K(K)n.

If μnK then

Homcts(Gal(K¯K),μn)K(K)n

Finite subgroups of Homcts(Gal(K¯K),μn) are of the form Hom(Gal(LK),μn) for LK a finite abelian extension of K of exponent dividing n.

This gives another proof of Theorem 11.2.

Notation. H1(K,) means H1(Gal(K¯K),).

Let ϕ:EE be an isogeny of elliptic curves over K. Short exact sequence of Gal(K¯K)-modules

0E[ϕ]EϕE0

has long exact seqeucne

E(K)ϕE(K)δH1(K,E[ϕ])H1(K,E)ϕH1(K,E).

We get a short exact sequence

              ′
 0           EϕE(K(K-))            H1 (K,E [ϕ])          H1 (K, E)[ϕ∗]  0


            ∏   E′(Kv)-        ∏    1               ∏    1        ∗
δrrδ0esesVvv           v ϕE(Kv)           vH  (Kv,E[ϕ])        vH  (Kv, E)[ϕ0]
Now take K a number field.

For each place v, fix an embedding K¯K¯v. Then Gal(K¯vKv)Gal(K¯K).

Definition (Selmer-group). We define the ϕ-Selmer group

S(ϕ)(EK)=ker(H1(K,E[ϕ])vH1(Kv,E))={αH1(K,E[ϕ])|resv(α)Im(δv) v}

(the map H1(K,E[ϕ])vH1(Kv,E) is as in the commutative diagram above).

Definition (Tate-Shafarevich group). The Tate-Shafarevich group is

TS(EK)=ker(H1(K,E)vH1(Kv,E)).

We get a short exact sequence

0E(K)ϕE(K)S(ϕ)(EK)TS(EK)[ϕK]0.

Taking ϕ=[n] gives

0E(K)nE(K)S(n)(EK)TS(EK)[n]0.

Reorganising the proof of Mordell-Weil gives

Theorem 15.3. S(n)(EK) is finite.

Proof. For LK a finite Galois extension,

                          finite
                  ◜-1------◞◟--------◝
 0                H  (Gal(L∕K ),E(L)[n ]) H1 (K,E [n])      H1(L,E [n])


inr⊃⊃fes                                     S (n)(E ∕K )       S(n)(E∕L )
Therefore by extending our field, we may assume E[n]E(K) and hence by the Weil pairing μnK.

Therefore E[n]μn×μn as a Gal(K¯K)-module. Then

H1(K,E[n])H1(K,μn)×H1(K,μn)K(K)n×K(K)n

Let

S={primes of bad reduction for E}{v|n}

(a finite set of places).

Define the subgroup of H1(K,A) unramified outside of S as

H1(K,A;S)=ker(H1(K,A)vSH1(Kvnr,A)).

There is a commutative diagram with exact rows:

    E (Kv )       E(Kv )          H1(Kv,E [n])


×δr×0nvesn  E (Knrv )      E(Kvnr )        H1(Knrv ,E [n])
The bottom map ×n is surjective vS (see Theorem 9.8).

Therefore

S(n)(EK)H1(K,E[n];S)H1(K,μn;S)×H1(K,μn;S)

But

H1(K,μn;S)=ker(K(K)nvS(Kvnr)(Kvnr)n)K(S,n)

which is finite by Lemma 11.4.

Remark. S(n)(EK) is finite and effectively computable.

It is conjectured that |TS(EK)|<. This would imply that rankE(K) is effectively computable.

16 Descent by Cyclic Isogeny

Let E,E be elliptic curves over a number field K, and ϕ:EE an isogeny of degree n.

Suppose E[ϕ^]n generated by TE(K). Then E[ϕ]μn as a Gal(K¯K)-module

Se(S,T).

Have a short exact sequence of Gal(K¯K)-modules

0μnEϕE0.

Get long exact sequence

   E (K )      E ′(K )        H1(K, μn)     H1 (K,E )


ϕδα∼=                           K∗∕(K ∗)n
(the is by Hilbert 90).

Theorem 16.1. Assuming that:

  • fK(E) and gK(E)

  • div(f)=n(T)n(0)

  • ϕf=gn

Then α(P)=f(P)(mod(K)n) for all PE;(K){0,T}.

Proof. Let Qϕ1P. Then δ(P)H1(K,μn) is represented by σσQQE[ϕ]μn.

e(σQQ,T)=g(σQQ+X)g(X)for any XEavoiding zeroes and poles of g=g(σQ)g(Q)pick X=Q=σ(g(Q))g(Q)=σf(P)nfn(P)

(ϕf=gn, f(P)=g(Q)n).

But

H1(K,μn)K(K)n(σxnxn)x

Hence α(P)=f(P)(mod(K)n).

16.1 Descent by 2-isogeny

E: y2=x(x2+ax+b)E: y2=x(x2+ax+b) b(a24b)0, a=2a, b=a24b. ϕ:EE(x,y)((yx)2,y(x2b)x2)ϕ^:EE(x,y)(14(yx)2,y(x2b)8x2)

E[ϕ]={0,T}, T=(0,0)E(K).

E[ϕ^]={0,T}, T=(0,0)E(K).

Proposition 16.2. There is a group homomorphism

E(K)K(K)2(x,y){xmod(K)2if x0bmod(K)2if x=0

with kernel ϕE(K).

Proof. Either: Apply Theorem 16.1 with f=xK(E), g=yxK(E).

Or: direct calculation – see Example Sheet 4.

αE:E(K)ϕ^E(K)K(K)2αE:E(K)ϕE(K)K(K)2

Lemma 16.3. 2rankE(K)=|ImαE||ImαE|4.

Proof. If AfBgC are homomorphisms of abelian groups, then there is an exact sequence

 0         ker(f)        ker(gf )       ker(g)


fg          coker(f)      coker(gf)      coker(g)   0
Since ϕ^ϕ=[2]E, we get an exact sequence

0E(K)[ϕ]2E(K)[2]ϕE(K)[ϕ]2E(K)ϕE(K)Im(αE)ϕ^E(K)2E(K)E(K)ϕ^E(K)Im(αE)0.

Therefore

|E(L)2E(K)||E(K)[2]|=|ImαE||ImαE|4.(1)

Mordell-Weil: E(K)Δ×r, where Δ is a finite group and r=rankE(K).

E(K)2E(K)Δ2Δ×(2)r.

E(K)[2]Δ[2].

Since Δ is finite, we have that Δ2Δ and Δ[2] have the same order, and therefore

|E(K)2E(K)||E(K)[2]|=2r.(2)

So we are done, by using (1) and (2).

Lemma 16.4. Assuming that:

  • K is a number field

  • a,bOK

Then Im(αE)K(S,2), where S={𝔭|b}.

Proof. We must show that x,yK, y2=x(x2+ax+b) and v𝔭(b)=0 then v𝔭(x)0(mod2).

Case v𝔭(x)<0: Lemma 9.1 gives that for some r1, v𝔭(x)=2r and v𝔭(y)=3r, so done.

Case v𝔭(x)>0: Then v𝔭(x2+ax+b)=0. So v𝔭(x)=v𝔭(y2)=2v𝔭(y).

Lemma 16.5. Assuming that:

  • b1b2=b

Then b1(K)Im(αE) if and only if
w2=b1u4+au2v2+b2v4(∗)

is soluble for u,v,wK not all zero.

Proof. If b1(K)2 or b2(K)2 then both conditions are satisfied. So we may assume b1,b2(K)2.

Now note b1(K)2Im(αE) if and only if there exists (x,y)E(K) such that x=b1t2 for some tK. This implies

y2=(b1t2)(b12t2+ab1t2+b)

hence

(yb1t)2=b1t4+at2+bb1=b2.

So () has solution (u,v,w)=(t,1,yb1t).

Conversely, if (u,v,w) is a solution to () then uv0 and (b1(uv)2,b1uwv3)E(K).

Now take K=.

Example. E: y2=x3x (a=0, b=1).

Im(αE)=1()2.

E: y2=x3+4x.

Im(αE)1,2()2.

b1=1w2=u44v4b1=2w2=2u4+2v4b1=2w2=2u42v4 The first and last lines are insoluble over (squares are non-negative). The middle line does have a solution: (u,v,w)=(1,1,2).

Therefore Im(αE)=2()2.

Hence 2rankE()=224=1, so rankE()=0.

So 1 is not a congruent number.

Example. E: y2=x3+px, p a prime which is 5 modulo 8.

b1=1w2=u4pv4

This is insoluble over , hence Im(αE)=p()2.

E: y2=x34px.

Im(αE)1,2,p()2.

Note: αE(T)=(4p)()2=(p)()2.

b1=2w2=2u42pv4b1=2w2=2u4+2pv4b1=pw2=pu44v4

Suppose the first line is soluble. Then without loss of generality u,v,w with gcd(u,v)=1. If p|u, then p|w and then p|v, contradiction. Therefore w22u40(modp). So (2p)=+1, which contradicts p5(mod8).

Likewise the second line has no solution since (2p)=1.

TODO

Example (Lindemann). E: y2=x3+17x.

Im(αE)=17()2.

E: y2=x368x.

Im(αE)1,2,17()2.

b1=2, w2=2u434v4.

Replacing w by 2w and dividing by 2 gives

C: 2w2=u417v4.

Notation. C(K)={(u,v,w)K3{0}satisfying equation above}, where (u,v,w)(λu,λv,λ2w) for all λK.

C(2) since 17(2)4.

C(17) since 2(17)2.

C() since 2.

Therefore C(v) for all places v of .

Suppose (u,v,w)C(). Without loss of generality say u,v,w, gcd(u,,v)=1 and w>0.

If 17|w then 17|u and then 17|v, contradiction

So if p|w then p17 and (17p)=+1. Therefore (p17)=(17p)=1 (using quadratic reciprocity). (If p is odd, also (217)=+1).

Therefore (w17)=+1. But 2w2u4(mod17), so 2(𝔽17)4={±1,±4}, contradiction.

So C()=, i.e. C is a counterexample to the Hasse Principle. It represents a non-trivial element of TS(E).

16.2 Birch Swinnerton Dyer Conjecture

Let E be an elliptic curve.

Definition (L(E,s)). Define

L(E,s)=pLp(E,s)

where

Lp(E,s)={(1apps+p12s)1if good reduction(1ps)1if split multiplicative reduction(1+ps)1if nonsplit multiplicative reduction1if additive reduction

where #(𝔽p)=1+pap.

Hasse’s Theorem implies |ap|2p, which shows that L(E,s) converges for Re(s)>32.

Theorem (Wiles, Breuil, Conrad, Diamon, Taylor). L(E,s) is the L-function of a weight 2 modular form, and hence has an analytic continuation to all of (and a function equation relating L(E,s)=L(E,2s)).

Conjecture (Weak Birch Swinnerton-Dyer Conjecture). ords=1L(E,s)=rankE().

(=r say).

Conjecture (Strong Birch Swinnerton-Dyer Conjecture). ords=1L(E,s)=rankE(), which we shall call r, and

lims11(s1)rL(E,s)=ΩERegE()|TS(E)|pcp(E)|E()tors|2,

where

Theorem (Kolyvagin). If ords=1L(E,s) is 0 or 1, then Weak Birch Swinnerton-Dyer holds, and also |TS(E)|<.

˙

Index

Tamagawa number

bad reduction

Cauchy

cochain

complete

congruent

elliptic curve

formal group

morphism

good reduction

Hilbert 90

integral

isogeny

isogenous

j-invariant

linearly equivalent

minimal

Mordell-Weil

bad reduction

good reduction

primitive

quadratic form

rational

rational

separable

elliptic curve

unramified

Weierstrass equation

Weierstrass form

Weak Mordell-Weil Theorem