• Keine Ergebnisse gefunden

3.6 Formulas for local error functions

3.6.4 Case (4)

Up to this point Theorem 3.30 has applied to all models we have had to consider. This is about to change and indeed we shall see that new compli-cations arise at once.

LetE denote the elliptic curve given by the Weierstrass equation Y2Z+h0f3Y +h1XY =X3+f2X2+f1f3X+f0f32 (3.26) and letE denote the elliptic curve given by the Weierstrass equation

Y2Z+h3f3Y +h2XY =X3+f4X2+f5f3X+f6f32. (3.27) According to Tate’s algorithm reproduced in [89,§IV.9] the reduction type of E is I0, since the reduction of the given equation is nonsingular, so in particular the valuation of its discriminant vanishes. Let K denote the Kodaira symbol for the reduction type ofE.

The curve C has reduction type [I0− K −l] for some l ≥0 and ∆(C) equals the discriminant of the given model ofE.

If we have l = 0, then the special fiber Cvmin is the same as the special fiber of reduction type K, but one of the rational curves of multiplicity 1

A C1 C2

C3 2

Figure 3.4: The special fiber of reduction type [I0−I0−0]

b b b

A

B I1 Il−1

C1 C2

Figure 3.5: The special fiber of reduction type [I0−IV −l]

is replaced by a curve A of genus 1. We denote the other components of multiplicity 1 in the special fiber of type K, if any exist, by C1, . . . , Ct, wheret+ 1∈ {1, . . . ,4}is the number of components ofK that have simple multiplicity. The case K=I0 is shown in Figure 3.4.

If l > 0, then Cvmin consists of the components making up the special fiber of type K, connected with a genus 1 component A of multiplicity 1 by a chain of l−1 curvesI1, . . . , Il−1 of genus 0. Here I1 intersects A and Il−1 intersects a component B of K of simple multiplicity and the other components of simple multiplicity are denoted C1, . . . , Ct as above. See Figure 3.5 for the case K=IV.

Lemma 3.59. The curve C has a model of the form (3.13) whose closure C has rational singularities if and only if l= 0. If l= 0, then this holds for the given model of C.

Proof. We have that the given model ofEisv-minimal if and only ifl= 0. If C −→ Cis a desingularization ofC, then it corresponds to a desingularization of the closure of the given model of E, where the strict transform of the nonsingular part is replaced by a curve of genus 1. Now we use Lemma 1.26 which tells us that having rational singularities only depends on the preimage of the singular locus. From Lemma 2.13 we get the second part of the lemma, since the given model isv-minimal ifl= 0.

If, on the other hand, l > 0, then the given equation of E is not v-minimal. In order to make itv-minimal, we need to apply the transformation

τ to the given model ofC that acts on affine points (ξ, η) by τ((ξ, η)) = (π−2lξ, π−3lη).

However, applying this transformation to the given equation of E results in a model of E that is not v-minimal and so we can again use Lemma 2.13.

If we havel= 0, then Theorem 3.30 and Lemma 3.59 imply thatεv and µv factor through the component group Φv of the N´eron model. Moreover, it is easy to see thatE cannot have multiplicative reduction, so the order of Φv is at most 4 and therefore the computation of µv becomes particularly easy.

There are several possible ways to do this computation. The most straightforward one consists in computingεv(P), εv(2P), εv(3P), εv(4P) un-til one of them equals zero and then using the definition of µv. However, the following approach, resembling the procedure used for elliptic curves first introduced in [87] (see Theorem 2.17) is faster. We use the multiplica-tion polynomials given by Uchida in [103] for models satisfyingH = 0 and generalized easily; more precisely the triplication function which we call

ψ3(x) = (ψ3,1(x), . . . , ψ3,4(x)), (3.28) satisfying:

• If xis a set of Kummer coordinates for P ∈J(kv), thenψ3(x) is a set of Kummer coordinates for 3P.

• ψ3((0,0,0,1)) = (0,0,0,1).

• ψ3(x) has coefficients in Z[f0, . . . , f6, h0, . . . , h3].

Note that ourψ3,i isµ3,i in Uchida’s notation. Forx∈KA(kv) we set ωv(x) :=v(ψ3(x))−9v(x)

and notice that, similarly to εv, this function is well defined on K(kv) and moreover, if we compose it with the usual surjection from J onto K, on J(kv).

Furthermore, Proposition 3.24 implies

µv(3P) = 9µv(P)−ωv(P). (3.29) Let us assume that we are given a point P ∈ J(kv) and we know that the reduction type of C over Ov is of the form [I0− K −0], where K is some Kodaira type. We also assumeεv(P)6= 0.

If 2P ∈J0(kv), then

µv(P) = 1 4εv(P),

but on the other hand we haveµv(3P) =µv(P). Therefore (3.29) implies µv(P) = 1

v(P) and

ωv(P) = 2εv(P).

If 3P ∈J0(kv), then we find µv(P) = 1

v(P) = 1 9ωv(P) so that the relation

ωv(P) = 3εv(P) (3.30)

holds.

The final case is 2P,3P /∈J0(kv). We have 4P ∈J0(kv) and hence µv(P) = 1

v(P) = 1

v(P) + 1

16εv(2P).

We cannot compute µv(P) directly if we find that (3.30) holds. But if we take a closer look which reduction types are possible in this case, we see that we must have K ∈ {IV, IV} if 3P ∈ J0(kv), whereas the complementary case can only occur if K = In and n is odd. This means that, at least if v(6) = 0, we can tell which case we are in by checking the valuation of the discriminant: For IV, IV it is even, whereas for K = In it is odd if and only ifnis. If the residue characteristic is equal to 2, then we know at least that ifv(∆) is odd, then we have typeIn and hence 3P /∈J0(kv). Similarly, if the residue characteristic is 3, then we must have reduction type IV or IV and hence 3P ∈J0(kv) ifv(∆) is even. If none of these conditions are satisfied, we simply check εv(2P) and εv(P) for equality.

This leads to Algorithm 2, where P ∈ J(kv) and we assume that the reduction ofJ over Ov is of the form [I0− K −0].

Remark 3.60. What about the height constant βv? If #Φv < 4, then we can use Corollary 3.32 because of Lemma 3.59. If we are in case K = IV or K =IV, then we have #Φv = 3 and if a search for P ∈J(kv) of small height produces no nontrivial εv(P), then we may have cv = #Φv(kv) = 1 and thus βv = 0. See the discussion following Corollary 3.32.

If we have #Φv = 4, then we can proceed as follows: If we find cv <4, then we are in the situation discussed already. If this does not hold, then we can computeεv(P) forP of small naive height. However, we can only be certain that we have determined all possible values if we find three different values if Φv∼=Z/2Z⊕Z/2Z (in fact, ifεv(P)6= 06=εv(Q), thenεv(P+Q) will yield the third value), respectively two if Φv ∼=Z/4Z, taken on by εv – unless we can also show somehow that we have found at least one point

Algorithm 2Computation of µv(P) for reduction type [I0− K −0]

if ωv(P) = 2εv(P) then return 18ωv(P)

else if ωv(P)6= 3εv(P) then return 18ωv(P)

else if v(6) = 0 then

return 9−(v(∆) (mod 2))1 ωv(P) else if v(2)>0 and 2∤v(∆) then

return 19ωv(P)

else if v(3)>0 and 2|v(∆) then return 18ωv(P)

else if εv(2P) =εv(P)then return 19ωv(P)

else

return 18ωv(P) end if

for each component. But we can get a small improvement even without computing cv or any εv just from knowing the exponent of Φv. In case Φv ∼=Z/2Z⊕Z/2Z we haveµv(P) = 14εv(P), so that any bound B for the maximum value γv of |εv| yields a bound βvB4 (as opposed to B3). In particular we could use the bound v(24disc(F)) from Proposition 3.11 or one of its improvements discussed in [92,§7]. If we have Φv ∼=Z/4Z, then a similar argument shows thatβv165B, whereB is any upper bound forγv. Iflis positive andE has additive reduction, the order of the component group is still at most 4. However, according to Lemma 3.59 the closure of the given model of C does not have rational singularities and there is no way to repair this. But because the implication of Theorem 3.30 has not been shown to be an equivalence, this does not necessarily mean thatεv and µv cannot factor through Φv and there is some hope left. Yet consider the following example.

Example 3.61. Let p be an odd prime and let C be the smooth projective model ofy2 = (x2+ 1)(x3 +p5x+p8) over Qp. Let P1 = (0, p4) ∈C and P2 =−P1. We have reduction type [I0−III−1] and hence #Φv = 2. It turns out that bothP1 andP2 map to the componentC1 (see the beginning of this section) and so we have P ∈ J0(kv). The image on the Kummer surface is of the form (x1,0,0, x4), where v(x4) −v(x1) = 2 = 2l. We get εv(P) = εv(2P) = 6 and, in accordance with Theorem 3.62 below, µv(P) =µv(2P) = 2 = 2l.

Hence the computation ofµv(P) becomes more involved. Still, below we prove a simple formula for µv(P) when P ∈ J0(kv) under some additional conditions which can always be ensured to hold after a simple

transforma-tion.

Theorem 3.62. Suppose C has reduction type [I0 − K −l], where K is not a multiplicative Kodaira type and l ≥ 0. Furthermore, suppose that v(h0) ≥ 3l, v(h1) ≥ l, v(f0) ≥ 6l, v(f1) ≥ 4l, v(f2) ≥ 2l and that the only transformation required when applying Tate’s algorithm to E is (ξ, η) 7→

−2lξ, π−3lη). Let x= (x1, x2, x3, x4) be a set of v-integral Kummer coor-dinates for P ∈ J0(kv) with v(x3) >0, v(x4) >0 and either v(x1) = 0 or v(x2) = 0. Then we have

µv(P) = min{v(x3), v(x4),2l}.

Proof. See Appendix A.8.

Remark 3.63. The condition on Tate’s algorithm basically amounts to re-quiring the coefficientsh0, h1, f0, f1, f2 to have valuation as large as possible simultaneously. In order to satisfy it, we simply apply Tate’s algorithm to the given model ofE, record the transformations needed and apply them to C. Having done so, we can apply Theorem 3.62 to the resulting Jacobian.

Remark 3.64. We can safely leave out the case K = In, n ≥ 1, because if we have such a curve, then we can apply a transformation producing a model that falls into case (5) and this will be dealt with in Section 3.6.5.

This reduction type is easily distinguishable, for example by applying Tate’s algorithm to the model of E.

Remark 3.65. The preceding theorem enables us to computeµv(P) for arbi-traryP using the fact that we can always find somen∈ {1, . . . ,4}such that nP ∈ J0(kv). The number n can be determined quite easily once we have applied the transformations necessary to use the theorem. Finally we note that the case n= 4 is generally not very common, so mostly an application ofδ orψ3 (see (3.28)) suffices. This is in contrast with the fact that we may have to go up to quite large multiples of our point if we want to ensure that εv vanishes for this multiple, see the conjecture below.

The proof of Theorem 3.62 given in Appendix A.8 is very elementary but also lengthy. The simplicity of the formulas hints at the existence of a more clever or at least more enlightening proof, possibly using N´eron’s interpretation of canonical local height pairings and Lang’s interpretation of N´eron functions as intersection multiplicities that we have already used in the proof of Theorem 3.30.

Recall that if H = 0 and char(kv) 6= 2, then J0(kv) consists of the elements ofJ(kv) that are nonsingular (when viewed as elements of the given model ofJ that is defined by 72 quadrics in P15 and that is determined by F) and map to the connected component of the identity of the special fiber of the given model. The group J0(kv) depends on the given model and

by Proposition 3.12 equals the group Uv of points on whichεv vanishes, at least for residue characteristic not equal to 2. Therefore we see that in the present caseJ0(kv) strictly containsJ0(kv). This phenomenon has also been discussed in a different context by Bruin and Stoll in [13, Remark 5.16]. On a side note, it should be possible to prove Proposition 3.12 in the situation where char(kv) = 2 using the explicit embedding of the Jacobian in P15 for fields of characteristic 2 (andh3 = 0) from [33], but we have not attempted this.

It is natural to ask about the index of J0(kv) in J0(kv). Experimental data suggests the following:

Conjecture 3.66. Suppose that k is a global field and that the given model of C is v-minimal with reduction type [I0− K −l], where K is an additive Kodaira type. Furthermore, suppose that v(h0) ≥ 3l, v(h1) ≥ l, v(f0) ≥ 6l, v(f1) ≥ 4l, v(f2) ≥ 2l and that applying Tate’s algorithm to the given equation of E yields no transformations except for (ξ, η) 7→ (π−2lξ, π−3lη).

Then we haveεv(plP) = 0 for all P ∈J0(kv).

Note that having reduction type [I0− K −l] means that there is a chain of l projective lines connecting the genus 1 curve coming from E and the Kodaira type. Each of these P1s contributes p new points to the special fiber of the minimal proper regular model ofC when compared to a curve of reduction type [I0− K −(l−1)]. We need the field to be global because for non-global one-dimensional function fields the quotient might not be finite.

We have not been able to give a proof in the general case, but we can show the following:

Proposition 3.67. If the residue characteristic is equal to 2, then Conjec-ture 3.66 holds.

Proof. The proof follows from the proof of Lemma A.3, see Appendix A.8.

The lemma is stated there for the situation

v(x1) = 0, v(x2)>0, 0≤v(x3)≤4l,

which impliesv(x4)≤ 12v(x3). For everyQ∈J(kv) letx(Q) denote a set of integral Kummer coordinates forQsuch that one of thex(Q)iis a unit. Then the proof of Lemma A.3 shows that we have v(x(2n+1P)4)−v(x(2nP)4)≥ 2v(2) as long asv(x(2nP)4)≥2v(2). The upper bound 2lforv(x4) implies thatv(x(2lP)4) = 0 and thusεv(2lP) vanishes.

Using Appendix A.8 it is easy to see that the same principle applies as long as we havev(x3)v(x4)>0 andK 6=I0. But ifv(x3)v(x4) = 0, then we haveεv(P) = 0 already, and so we are done.

The proof suggests that if k is a number field and l is divisible by the ramification indexeofkv overQv, then we might haveεv(pl/eP) = 0 for any

P ∈ J0(kv) and indeed this was the case in our examples. However, since most of our experiments dealt with either Qv or unramified extensions, we have not dared to include this into the statement of the conjecture. Anyway, for practical purposes – at least for our intended applications – Conjecture 3.66 does not seem to have much relevance. IfK=I0, then there are several counterexamples to the statement of the conjecture.

Example 3.68. LetC be given by

y2= (x2+ 1)(x3+p6),

where p >2 is a prime less than 100, and let P = [(0, p3)− ∞]. Then we have εv(nP)6= 0 for n∈ {1, . . . ,5}, butεv(6P) = 0.

Remark 3.69. According to Lemma 3.59 we cannot use Corollary 3.32 to bound the height constant. Yet if we can show that there are no P ∈ J(kv)\J0(kv), then we get the bound

βv ≤2l,

which is certainly a significant improvement over the bound we get by ap-plying Proposition 3.11. This condition is satisfied when K ∈ {I0, II, II} and more generally when the Tamagawa numbercv is equal to 1.

Im Dokument Computing canonical heights on Jacobians (Seite 114-121)