Elliptic Surface of Kodaira Dimension 1
Elliptic Surface of Kodaira Dimension 1
Elliptic Surface of Kodaira Dimension 1
Abstract
We consider elliptic surfaces E over a field k equipped with zero section O and another
section P of infinite order. If k has characteristic zero, we show there are only finitely
many points where O is tangent to a multiple of P. Equivalently, there is a finite list
of integers such that if n is not divisible by any of them, then n P is not tangent to
O. Such tangencies can be interpreted as unlikely intersections. If k has characteristic
zero or p > 3 and E is very general, then we show there are no tangencies between
O and n P. We apply these results to square-freeness of elliptic divisibility sequences
and to geography of surfaces. In particular, we construct mildly singular surfaces of
arbitrary fixed geometric genus with K ample and K 2 unbounded.
1 Introduction
Our aim in this paper is to study transversality properties of sections of elliptic sur-
faces and to deduce consequences for elliptic divisibility sequences and geography of
surfaces.
B Douglas Ulmer
[email protected]
Giancarlo Urzúa
[email protected]
0123456789().: V,-vol
25 Page 2 of 36 D. Ulmer, G. Urzúa
To state the first result, let k be a field of characteristic zero and let C be a smooth,
projective, geometrically irreducible curve over k. Let π : E → C be a relatively
minimal Jacobian elliptic surface over k (i.e., a smooth, projective elliptic surface
with a section O which will play the role of zero section), and let P be another
section. We write n P for the section induced by multiplication by n in the group law
of the fibers of E → C. Assume that P has infinite order, i.e., n P = O for all n = 0.
As we will see below, except in degenerate situations the intersection number (n P).O
grows like a constant times n 2 . Our first result says that the intersections are usually
transverse.
Theorem 1.1 The set
T = {t ∈ C | n P is tangent to O over t }
n=0
is finite.
Here and in the rest of the paper, we conflate the sections O : C → E and P : C → E
with their images O(C) ⊂ E and P(C) ⊂ E. Thus we say “P is tangent to O” rather
than “the image of P is tangent to the image of O.”
In [32], we give an explicit upper bound on the cardinality of T .
Remark 1.2 We note that a tangency between n P and O can be regarded as an “unlikely
intersection” as follows: Let TE be the tangent bundle of E and let PTE be the associated
projective bundle. Thus PTE → E is a P1 -bundle, and the total space PTE is a smooth,
projective threefold. If C ⊂ E is a smooth curve, then there is a canonical lift of C to
C̃ ⊂ PTE defined by sending a point t ∈ C to the class of its tangent line TC,t ⊂ TE ,t
in PTE . Two curves C1 and C2 in E that meet at y ∈ E are tangent there if and only
if their lifts meet at a point of PTE over y. Thus a tangency between C1 and C2 is
equivalent to the “unlikely” intersection of the two curves C̃1 and C̃2 in the threefold
PTE . We refer to [33] for a comprehensive account of work on unlikely intersections
up to 2012.
Remarks 1.3 (1) A result very similar to our Theorem 1.1 was communicated to us
by Corvaja, Demeio, Masser, and Zannier after we posted the first version of this
paper. Their methods are rather different, see [11]. They show more generally
that finiteness holds when the cyclic group {n P|n ∈ Z} is replaced by a finitely
generated, torsion-free group of sections.
(2) On the other hand, our methods lead to non-trivial results in families, and in partic-
ular we show that for “generic” data, the set T above is empty. (See Theorems 1.7
and 1.8 below.) This is crucial for our application to geography of surfaces.
(3) In the first version of this paper, we used a trivialization essentially equivalent to
the Betti foliation discussed in Sect. 3 of this version. Later, we learned of the
“Betti” terminology used by several authors, including in [12], and adopted it in
this paper.
We next reformulate Theorem 1.1 in analogy with the “elliptic divisibility sequence”
associated to an elliptic curve and a point. (See [28, Exers. 3.34-36, 9.4, 9.12] for
Transversality of sections on elliptic surfaces Page 3 of 36 25
definitions and examples, and [18] for more on the function field case.) Define a
sequence of effective divisors on C for n ≥ 1 by
Dn := O ∗ (n P),
i.e., Dn is the pull-back along the zero section of the divisor n P on E. (We will give
several other equivalent definitions in Sect. 2.)
The sequence Dn is a natural analogue of an elliptic divisibility sequence. In par-
ticular, we will see below that if m divides n, then Dm divides Dn (i.e., Dn − Dm is
effective), and that Möbius inversion gives a sequence of effective divisors Dm such
that
Dn = Dm .
m|n
where the ti are distinct closed points of C (i.e., each non-zero coefficient of D equals
1). This is an analogue of an integer being square-free.
Theorem 1.4 Given E and P as above, there is a finite set of integers M =
{m 1 , . . . , m k } such that
(1) O and n P intersect transversally if and only if n is not divisible by any element
of M.
(2) Dn is reduced if and only if n is not divisible by any element of M.
(3) Dm is reduced if and only if m ∈/ M.
Remark 1.5 Thereom 1.4 is much stronger than what one might predict from standard
conjectures. For simplicity, assume that C = P1 , let F = C(C), and let E/F be the
generic fiber of E → C. Choosing a coordinate t on P1 so that none of the Dn involve
the place at infinity, we may identify each Dn with a monic polynomial f n in t, and
to say that Dn is reduced is to say that gcd( f n , d f n /dt) = 1. Arguments similar to
those in [27] applied to a certain Buium jet space of E/F together with a function
field analogue of Vojta’s conjecture suggest that
?
deg gcd( f n , d f n /dt) = o(n 2 ) = o(h(n P))
where h(n P) is the canonical height of n P. (See Sect. 2.7 for definitions.) But Theo-
rem 1.4 shows that deg gcd( f n , d f n /dt) is in fact bounded!
Remark 1.6 We have no reason to believe that the analogues of Theorems 1.1 and 1.4
(with n restricted to be prime to the characteristic) are false in positive characteristic.
However, our proof uses analytic techniques and does not obviously carry over to the
arithmetic situation.
25 Page 4 of 36 D. Ulmer, G. Urzúa
The next two results hold for k a field of characteristic zero or sufficiently large p.
As before, C is a smooth, projective, geometrically irreducible curve over k. The next
result says roughly that if E → C is a very general Jacobian elliptic surface with an
additional section P, there are no tangencies between n P and O for n = 0. Recall
that a line bundle L on C is said to be globally generated (or base point free) if for
every t ∈ C, there is a global section of L which does not vanish at t.
V = H 0 (L 2 ⊕ L 3 ⊕ L 4 ).
E : y 2 + a3 y = x 3 + a2 x 2 + a4 x
equipped with the usual zero section O and the section P = (0, 0) has the following
properties:
(Here and elsewhere in the paper, L i means L ⊗i and not L ⊕i .) We will explain the
construction of the elliptic surface attached to a in Sect. 5.4 and the meaning of “very
general” in Sect. 6.
As with many results about “very general” points, Theorem 1.7 does not allow one
to deduce the existence of examples over “small” (countable) fields such as number
fields or global function fields. However, after relaxing condition (2) above, we can
write down such examples explicitly, at least when L is the square of a globally
generated line bundle.
Theorem 2.5 in Sect. 3. Section 4 discusses two moduli spaces which play a key role
in the proof of Theorem 1.7. Section 5 discusses a construction of elliptic surfaces
equipped with extra structure associated to elements in certain Riemann-Roch spaces.
We then prove Theorem 1.7 in Sect. 6. In Sect. 7, we give an explicit construction
of surfaces satisfying the requirements of Theorem 1.8. Finally, in Sect. 8, we prove
Theorem 1.9.
2.1 Multiplication by n
Let E sm denote the locus where π is smooth (i.e., the complement of the singular
points in the bad fibers). Then by [15, Prop. II.2.7], E sm is a commutative group
scheme over C. Let n be an integer not divisible by the characteristic of k. Consider
the homomorphism of group schemes given multiplication by n: [n] : E sm → E sm .
Clearly, [n] fixes the zero section O pointwise. If x ∈ O, the tangent space to E sm at
x splits canonically into the sum of two lines, namely the tangent space to O at x and
the tangent space to the fiber of π through x. Since [n] fixes O, [n] acts as the identity
on the former. A calculation in the formal group of E sm [28, Ch. IV] shows that [n]
acts as multiplication by n on the tangent space to the fiber of π through x. It follows
that [n] is étale at every point of O, and since [n] is a group scheme homomorphism
it is étale everywhere.
The morphism [n] is also quasi-finite: it has degree n 2 on the smooth geometric
fibers of π , and degree dividing n 2 on all geometric fibers [28, Ch. III and §VII.6]. It
is not in general finite if π has singular fibers.
If P : C → E is a section of π , P necessarily lands in the smooth locus E sm and
we may define a new section n P as the composition [n] ◦ P. This is the meaning of
the notation n P used in the introduction.
2.2 Torsion
With E as above and n > 0 and relatively prime to the characteristic of k, we define
i.e, E[n] is the inverse image of the zero section under [n]. Since [n] is étale, E[n] is
a reduced, closed subscheme of E sm of dimension 1, and in particular, locally closed
in E. Since [n] is quasi-finite, E[n] is étale and quasi-finite over C of generic degree
n 2 . It is in general not finite over C.
Transversality of sections on elliptic surfaces Page 7 of 36 25
The fiber of E[n] over a geometric point t of C consists of the points of π −1 (t) with
order divisible by n. We define
E[n] ⊂ E[n]
to be the subscheme whose fiber over t consists of the points of π −1 (t) of order exactly
n. If m divides n, then E[m] is a closed subscheme of E[n], and we have a disjoint
union
where m runs over the positive divisors of n. Each E[m] is a union of irreducible
components of E[n] and is étale and quasi-finite over C. Note that E[1] = E[1] = O.
We refer to the unions of irreducible components of E[n] as “torsion multisections”.
Dn = O ∗ (n P).
Dn = π∗ (n P ∩ O) = π∗ (n P ∩ E[1]) .
Note that n P meets O = E[1] over t if and only if P meets E[n] over t, and since
[n] is étale, the intersection multiplicity of n P and E[1] over t equals the intersection
multiplicity of P and E[n] over t. In other words, we have
Dn = π∗ (P ∩ E[n]) . (2.2)
Define
Dn = π∗ P ∩ E[n] .
Then the disjoint union (2.1) yields a decomposition of Dn into effective divisors:
Dn = Dm (2.3)
m|n
Note in particular that if t is a closed point of C and P(t) is a torsion point, say of
order exactly m, then t appears in Dn if and only if m divides n, and the multiplicity
of t in such Dn is equal to the multiplicity of t in Dm .
Remark 2.4 A section P can meet at most one torsion point over a given t ∈ C. This
and D have disjoint support. In particular, as soon
implies that if m 1 = m 2 , then Dm 1 m2
as Dn = 0, Dn has “primitive divisors’ i.e., points in its support which are not in the
support of Dm for m < n. The existence of primitive divisors for all sufficiently large
n is established in [18, §5] by showing that Dn = 0 for all sufficiently large n. The
key idea is an estimation of intersection numbers using heights as in Sect. 2.7 below.
Another simple proof of the existence of primitive divisors (suggested by a referee)
can be given using the fact that the “Betti coordinates” (as in Sect. 3) of a non-torsion
section P give rise to a locally defined, open map from C to R2 . The existence and
openness of this map can also be viewed as the key point in the simplest case (I0 ) of
the proof of Theorem 2.5.
We now state a result which implies Theorems 1.1 and 1.4:
Theorem 2.5 Let E → C be a relatively minimal Jacobian elliptic surface over the
complex numbers C, with zero section O and another section P which is not torsion.
Then the set
Ttor := {t ∈ C | P is tangent to E[n] over t }
n=0
is finite.
2.6 Proof that Theorem 2.5 implies Theorems 1.1 and 1.4
First we note that the general cases of Theorems 1.1 and 1.4 follow from the case
k = C. Indeed, since the hypotheses and conclusion of Theorems 1.1 and 1.4 are
insensitive to the ground field, we may replace k with a subfield k which is finitely
generated over Q (take the field generated by the coefficients defining C, E, π , and
P), then embed k in C. Thus it suffices to treat the case k = C.
Next, by the definition of Dn , to say that n P is tangent to O over t is to say that
t appears in Dn with multiplicity greater than 1. By the equality (2.2), to say that t
appears in Dn with multiplicity greater than 1 is to say that P is tangent to E[n] over t.
Thus the set Ttor of Theorem 2.5 is equal to the set T of Theorem 1.1, and Theorem 2.5
is equivalent to the case k = C of Theorem 1.1.
To finish, we show that Theorems 1.1 and 1.4 are equivalent. First note that points
(1) and (2) of Theorem 1.4 are equivalent by the definition of Dn . Moreover, Dm is
The set T of Theorem 1.1 is the image of the projection I → C and the set M of
Theorem 1.4 is the image of the projection I → Z>0 . The fibers of I → C are finite
(and in fact empty or singletons) because P meets E[m] for at most one value of m
and a fortiori can be tangent to at most one E[m] . The fibers of I → Z>0 are finite
because for a fixed m, P meets E[m] at only finitely many points, so a fortiori can
be tangent to E[m] at only finitely many points. This establishes that Theorem 1.1
and Theorem 1.4 are equivalent, and it completes the proof that Theorem 2.5 implies
Theorems 1.1 and 1.4
We will prove Theorem 2.5 in Sect. 3. First, we review material on heights used
later in the paper.
2.7 Heights
We refer to [13] or [25] or [26] or [31] for the basic assertions on heights in this section.
As usual, π : E → C is a relatively minimal Jacobian elliptic surface over a field k.
Given a section P of π , there is a unique Q-divisor C P supported on the non-
identity components of the fibers of π with the property that P − O + C P has zero
intersection multiplicity with every irreducible component of every fiber of π . There
is a simple recipe for C P that depends only on the components of the reducible fibers
met by P, and in particular, for a fixed E, there are only finitely many possibilities for
C P as P varies over all sections. If π has irreducible fibers, or more generally, if P
passes through the identity component of every fiber, then C P = 0.
There is a canonical Q-valued symmetric bilinear form on the group of sections of
E defined by
ht(P) 2
(n P).O = n + O(1).
2
ht(P) 2
(n P).O = n − d.
2
Proof It follows from the canonical bundle formula for elliptic surfaces, adjunction,
and the definition of d that every section P of π satisfies P 2 = −d. From the height
25 Page 10 of 36 D. Ulmer, G. Urzúa
formula, we find
and so
ht(P) 2 ht(P) 2
(n P).O = n − d + Cn P .(n P − O) = n + O(1).
2 2
If π has irreducible fibers, then C Q = 0 for all Q and we deduce the stated exact
formula.
In the following lemma we use square brackets to indicate the class of a curve in
NS(E), the Néron-Severi group of E. This allows us to distinguish between n[P] (n
times the class of P) and [n P] (the class of n P).
Lemma 2.9 Suppose that π : E → C has irreducible fibers, d = deg O ∗ (1E /C ), and
P is a section of π which does not meet O. Let F be a fiber of π . Then we have an
equality
in NS(E).
Proof We have an equality [n P] − [O] = n([P] − [O]) in the Picard group of the
generic fiber of E, so there is an equality of the form
in NS(E), and we just need to determine the coefficient of [F]. We do this by intersect-
ing with [O]. By assumption [P].[O] = 0, so ht(P) = 2d. By the previous lemma,
[n P].[O] = d(n 2 − 1) and solving for c yields c = d(n 2 − n).
We first note that Theorem 2.5 is a statement about intersections on an elliptic surface
over the complex numbers. To prove it, we may replace E and C with the corresponding
complex manifolds and make use of the classical topology, i.e., the topology induced
by the metric topology on C. For the rest of this section, we make this replacement,
although we will not change the notation.
Our strategy will be to show that the subset Ttor ⊂ C is closed and discrete (in the
classical topology). This implies Theorem 2.5 since C is compact. We note in passing
that the set of points of intersection of P and E[n] for varying n is usually everywhere
classically dense in P, so the discreteness that lies at the heart of the theorem is not
evident.
In fact, will consider a more general set of tangencies and prove that a certain subset
TBetti ⊂ C is closed and discrete (and thus finite) and contains all points of Ttor over
Transversality of sections on elliptic surfaces Page 11 of 36 25
which E has good reduction. Since the set of points of bad reduction is finite, this will
establish that Ttor is also finite.
We will establish the desired discreteness by using the complex analytic description
of E → C given by Kodaira in [20, §8]. Let C 0 ⊂ C be the maximal open subset over
which E has good reduction, and let E 0 = π −1 (C 0 ).
Let t ∈ C 0 . Write H for the upper half plane. Then there is a neighborhood U of t
biholomorphic to a disk and holomorphic functions τ : → H and w : → C
such that π −1 (U ) → U sits in a diagram
π −1 (U ) ( × C)/(Zτ + Z)
P|U [w] (3.1)
U
where the horizontal maps are biholomorphic, and ( × C)/(Zτ + Z) means the
quotient of × C by Z2 acting as
×C × R2
with inverse
The two vertical maps are the natural quotients, the middle horizontal map is induced
by the upper horizontal map, and the diagonal maps are the projections to the first
factor.
The top horizontal map is a real-analytic isomorphism which is R-linear on each
fiber of the projection to . The choice of this map is motivated by the fact that torsion
sections of E over U correspond the surfaces × (r , s) ⊂ × (R/Z)2 where r and
s are rational numbers. In other words, we have changed coordinates so that every
torsion section becomes a constant section. It will be of interest to consider all the
horizontal sections × (r , s) for arbitrary real numbers r and s.
Definition 3.1 With notation as above and (r , s) ∈ R2 , define the local Betti leaf
Lr ,s ⊂ π −1 (U ) as the image of the map U → ( × C)/(Zτ + Z) ∼= π −1 (U ) sending
z to the class of (z, r τ (z) + s).
This terminology is inspired by [12], where r and s are called “Betti coordinates”.
Clearly the assignment (r , s) → Lr ,s factors through (R/Z)2 . Each Lr ,s is a closed
holomorphic submanifold of π −1 (U ), and the set of Lr ,s as (r , s) runs through (R/Z)2
is a foliation of π −1 (U ). Although the indexing of Lr ,s by (r , s) depends on the choice
of period map τ , the submanifolds Lr ,s themselves are intrinsic (i.e., independent of
τ ). There is a corresponding global foliation of E 0 which we will not consider in this
paper, except implicitly in the following remark: If for some non-empty open U ⊂ C 0
and some (r , s), P(U ) = Lr ,s (i.e., P lands in a leaf of the local foliation), then by
analytic continuation, the same holds over every open. In this case, we say “P lies in
the Betti foliation”.
Note that if (r , s) ∈ Q2 , then each point of Lr ,s is a torsion point in its fiber. More
precisely, if n is the smallest positive integer such that (nr , ns) ∈ Z2 , then Lr ,s is a
connected component of E[n] ∩ π −1 (U ). On the other hand, if (r , s) ∈ R2 \ Q2 , then
Lr ,s is disjoint from every E[n]. Thus, if P lies in the Betti foliation, it is a torsion
section (which we have ruled out by hypothesis) or it meets no torsion sections over
C 0 and so is not tangent to any torsion section over C 0 . In the latter case, Theorem 2.5
is obviously true, so we may assume from now on that P does not lie in the Betti
foliation.
We now define
TBetti := t ∈ C 0 |P is tangent to some Lr ,s over t .
The preceding paragraph shows that Ttor ∩ C 0 ⊂ TBetti and so, as explained above,
to prove Theorem 2.5 it will suffice to prove that TBetti is a closed, discrete subset of
C. More formally:
To establish Claim 3.2, we will consider cases according to the reduction type of E
at t. We use the standard Kodaira notation (In , In∗ ,…) to index the cases.
Transversality of sections on elliptic surfaces Page 13 of 36 25
The case of I0 reduction: Using diagrams (3.1) and (3.2), we identify the section
P over U with the graph of a function φ : → R2 . Write (r0 , s0 ) = φ(0) for the
image of P over t. Since P is assumed not to be contained in Lr0 ,s0 , we may shrink
U so that P meets Lr0 ,s0 only over t, in other words, so that the only value of z with
φ(z) = (r0 , s0 ) is z = 0.
It is clear that P is tangent to some Lr ,s over t if and only if the derivative of
(x, y) → (r , s) (as a map of 2-manifolds) vanishes at the z corresponding to t .
To finish, we claim that after possibly shrinking U and , the derivative of φ does
not vanish away from 0 ∈ . To see this, apply the Lojasiewicz gradient inequality
([10,24]) to the components of (φ1 , φ2 ) of φ: That result says that after shrinking ,
there are constants C > 0 and 0 < θ < 1 such that
for all z ∈ . But if z ∈ \ {0}, φ(z) = φ(0) so one of the φi (z) = φi (0) which
implies that ∇φi (z) = 0 and so the derivative of φ is also non-zero.
This establishes Claim 3.2 at points of good reduction: if t is such a point, there
is an open neighborhood Ut of t in C such that P is not tangent to any Lr ,s over any
t ∈ Ut \ {t}. To finish the proof, we deal with tangencies near places of bad reduction.
The case of I1 reduction: We next consider the case of multiplicative reduction with
an irreducible special fiber. I.e., assume that E has reduction type I1 over t ∈ C. Let
→ C be a holomorphic parameterization of a neighborhood of t, where is the
unit disk and 0 ∈ maps to t. Again, over the course of the proof we will reduce the
radius of but not change the notation. Let X → be the pull-back of E → C to
, let = \ {0}, and let X → be the restriction of X → to . Note that
the special fiber
X \ X = nodal cubic ∼
= C× ∪ {q}
φ : × C× → X
25 Page 14 of 36 D. Ulmer, G. Urzúa
(Note that the logarithms appearing in the expression for r are evaluated at real num-
bers, so we use the standard real logarithm, and the class of s in R/Z is independent
of the choice of logarithm.)
Then we calculate that P is tangent to Lr ,s at z ∈ V if and only if
Note that the expression on the right is well defined independently of the choice of V
and the logarithm.
Now we assume that there is a sequence of tangencies accumulating at t and derive
a contradiction. More precisely, assume that there is a sequence z i ∈ tending to 0
such that for all i,
f (z i ) log | f (z i )|
f (z i ) = . (3.4)
zi log |z i |
for some positive constants B1 and B2 and all z ∈ . Shrinking again if necessary,
we may write f (z) = w0 (1 + h(z)) with
Taking z i close to zero and noting that g(z i ) = (log | f (z i )|)/(log |z i |) we get a
contradiction to the lower bound in Equation (3.5).
To finish, assume that |w0 | = 1. Then
Taking z i close to zero and noting that g(z i ) = (log | f (z i )|)/(log |z i |) we get a
contradiction to the upper bound in Equation (3.5).
This establishes that there is no accumulation of tangencies between P and local
Betti leaves at t when E has reduction of type I1 at t.
The case of Ib reduction: Now consider the case of multiplicative reduction of type Ib
over t ∈ C. This case is very similar to the I1 case, with some notational complications.
Let → C be a holomorphic parameterization of a neighborhood of t, where is
the unit disk and 0 ∈ maps to t. Again, over the course of the proof we will reduce
the radius of but not change the notation. Let X → be the pull-back of E → C
to , let = \ {0}, and let X → be the restriction of X → to . Then the
special fiber X \ X has the form
X \ X = chain of b copies of P1 ∼
= Ci× ∪ {qi }
i∈Z/bZ
X sm = X \ {q1 , . . . , qb }
For z ∈ and w ∈ C× , write (z, w)i for the class of (z, w) in Wi . Then X sm is
obtained by glueing the Wi according to the rule
× D → Wi → X
In this section, we discuss certain moduli spaces of elliptic curves with additional
structure. These spaces will be useful when we consider families of elliptic surfaces
in the following section. We work in more generality than needed in this paper, and
readers who are so inclined may replace the base ring R below with a field k of
characteristic = 2, 3 or even with C.
We begin by noting that there is a standard model for an elliptic curve E equipped
with a non-zero differential ω and a non-trivial point P: Given the data, choose a
Weierstrass model of E
such that ω = d x /(2y + a1 x + a3 ). Then there is a unique change of coordinates
x = x + r , y = y + sx + t such that P has coordinates (x, y) = (0, 0) and a1 = 0.
Thus there is a unique triple (a2 , a3 , a4 ) such that E is the elliptic curve defined by
y 2 + a3 y = x 3 + a2 x 2 + a4 x,
π :W→S
whose geometric fibers are reduced and irreducible curves of arithmetic genus 1
equipped with a section O : S → W whose image is contained in the locus where π
is smooth.
Let R = Z[1/6] and consider the stack M over Spec R whose value on an R-scheme
S is the set of triples (W → S, ω, P) where W → S is a curve of genus 1 over S as
defined above, ω is a nowhere vanishing section of O ∗ (1W /S ), and P : S → W is a
section disjoint from O. Two such triples (W → S, ω, P) and (W → S, ω , P ) are
isomorphic if there exists an S-isomorphism W → W carrying ω to ω and P to P .
Proposition 4.1 The stack M is represented by the affine scheme Spec R[a2 , a3 , a4 ].
The universal object over M is the projective family of plane cubics W →
Spec R[a2 , a3 , a4 ] defined by
y 2 + a3 y = x 3 + a2 x 2 + a4 x
where x, y, and z have weight 1. Then W = ProjSpec R[a2 ,a3 ,a4 ] (R).
Here and later in the paper, whenever we have elements a2 , a3 , a4 in some ring, we
set
We often omit the ai and simply write c4 , c6 , or . However, in the proof just below, we
do not omit the ai , i.e., we distinguish between the elements c4 , c6 generating a two-
variable polynomial ring R[c4 , c6 ] and the elements c4 (a2 , a3 , a4 ) and c6 (a2 , a4 , a6 )
in the ring R[a2 , a3 , a4 ].
Proof of Proposition 4.1 By [14, Prop. 2.5], the stack of pairs (W → S, ω) as above
is represented by the affine scheme Spec R[c4 , c6 ] with universal curve
c4 c6
y 2 = x 3 − x − 5 3
24 3 2 3
and universal differential d x /2y . (Deligne uses the more traditional coordinates
g2 = c4 /(22 3) and g3 = c6 /(23 33 ), but this is immaterial since 1/6 ∈ R.) Define a
morphism Spec R[a2 , a3 , a4 ] → Spec R[c4 , c6 ] by sending
Then pulling back the universal curve over Spec R[c4 , c6 ] to Spec R[a2 , a3 , a4 ] and
making the change of coordinates x = x + a2 /3, y = y + a3 /2 yields the curve and
differential mentioned in the statement of the theorem.
To finish the proof, one checks that the fibers of Spec R[a2 , a3 , a4 ] → Spec R[c4 , c6 ]
are the affine plane curves
c4 c6
y 2 = x 3 − 4
x − 5 3,
2 3 2 3
Transversality of sections on elliptic surfaces Page 19 of 36 25
i.e., Spec R[a2 , a3 , a4 ] is the universal curve over Spec R[c4 , c6 ] minus its zero section.
Indeed, the fiber over (c4 , c6 ) is
c4 = 16a22 − 48a4
c6 = 288a2 a4 − 64a23 − 216a32 .
a32 a23 c4 a2 c6
2
= 3
− 4 2 − 5 3.
2 3 2 3 2 3
In light of the proposition, from now we change notation and let M be defined as
the scheme Spec R[a2 , a3 , a4 ]. Also, we write Msm for the locus where = 0 and
Mn for the locus where = 0 and c4 = 0. Similarly, let N = Spec R[c4 , c6 ], N sm
the locus where c43 − c62 = 0, and N n the locus where c43 − c62 = 0 and c4 = 0.
4.2 Torsion
ψ2 = a3 ,
ψ3 = a2 a32 − a42 ,
ψ4 = 2a2 a33 a4 − 2a3 a43 − a35 ,
25 Page 20 of 36 D. Ulmer, G. Urzúa
E a : y 2 + a3 y = x 3 + a2 x 2 + a4 x
over k is nodal. We further assume that (a3 , a4 ) = (0, 0) so that P = (0, 0) and the
node, call it Q, are distinct. Let Gm be the multiplicative group over k. Then, possibly
after extending k quadratically, there is a group isomorphism
E a \ {Q} → Gm
−c6 2
y 2 = x 3 + x
4c4
where as usual c6 = 288a2 a4 −64a23 −216a32 . Letting γ be a square root of −c6 /(4c4 ),
the map to Gm is
y − γ x
(x , y ) →
y + γ x
Transversality of sections on elliptic surfaces Page 21 of 36 25
a3 c4 − γ (16a2 a4 − 36a32 )
(4.1)
a3 c4 + γ (16a2 a4 − 36a32 )
5 From E/K to E → C
We remind the reader how to go from an elliptic curve over a function field to an elliptic
surface. Although this is not strictly necessary for our main purposes, it suggests a
fruitful point of view on finite-dimensional families of elliptic surfaces parameterized
by certain Riemann-Roch spaces.
y 2 + a3 y = x 3 + a2 x 2 + a4 x,
WU → U
of curves of genus 1 (in the sense used before Proposition 4.1) with a section PU
disjoint from O, and the general fiber of WU → U is E/K equipped with P. If {U j }
is an open cover with trivializations φ j of L |U j , there is a unique way to glue over
the intersections compatible with the identification of the generic fiber of WU j → U j
with E/K , and the result is a global family W → C of curves of genus 1 equipped
with a section which we again denote by P. Writing P for the P2 bundle over C given
by
P = PC L 2 ⊕ L 3 ⊕ OC
25 Page 22 of 36 D. Ulmer, G. Urzúa
(with coordinates [x, y, z] on the fibers), we see that W is the closed subset of P
defined by the equation
y 2 + a3 y = x 3 + a2 x 2 + a4 x
There is a subtle point hiding in the last step of this construction: The section P of
W → C is disjoint from O, yet a section of E → C may very well meet O. Therefore,
there may be some blowing down in the last step to force such an intersection. We
make a few more comments about this situation and then give an example.
The underlying issue is that the local models WU → U are in a sense minimal
with respect to pairs “elliptic fibration + nowhere zero section,” but they may not be
minimal if we forget the section. We can quantify this as follows: Given E/K and P,
choosing ω leads to coefficients ai ∈ K and to invariants
Recall that D was defined as the smallest divisor on C such that div(ai ) + i D ≥ 0 for
i = 2, 3, 4. Similarly, let D be the smallest divisor on C such that div(c j ) + j D ≥ 0
for j = 2, 4. Then it is clear that D ≥ D and the points entering into D − D are
exactly those where the model W → C is not minimal (in the sense of [28, p. 816]).
Moreover, while W sits naturally as a divisor in
P = PC L 2 ⊕ L 3 ⊕ OC ,
5.3 An example
w2 = z 3 + t 2 z − 1
Transversality of sections on elliptic surfaces Page 23 of 36 25
y 2 + 2t −3 y = x 3 + 3t −2 x 2 + (3t −4 + t 2 )x
D = 0 + ∞ and D = ∞.
y 2 + 2y = x 3 + 3x 2 + (3 + t 6 )x.
The fiber over t = 0 is a cubic with cusp at t = 0, x = y = −1, and the surface
WA1 is singular at this point. Resolving the singularity requires blowing up once and
normalizing, and a further blow down removes a (−1)-curve in the fiber. This last
blow down brings the section P into contact with the zero section O.
We take the following point of view on constructing elliptic surfaces over C: Start with
a line bundle L on C. Then for each
a = (a2 , a3 , a4 ) ∈ H 0 (C, L 2 ⊕ L 3 ⊕ L 4 )
y 2 z + a3 yz 2 = x 3 + a2 x 2 z + a4 x z 2
in
P = PC L 2 ⊕ L 3 ⊕ OC .
Theorem 6.1 Let L be a globally generated line bundle on C of degree d > 0, and set
V = H 0 (L 2 ⊕ L 3 ⊕ L 4 ).
E: y 2 + a3 y = x 3 + a2 x 2 + a4 x
Lemma 6.2 The subset V ⊂ V consisting of a such that (a) has 12d distinct zeroes
(as a section of L 12 ) is Zariski open and not empty. There are a ∈ V whose zeroes
are disjoint from any given finite subset of points of C.
Transversality of sections on elliptic surfaces Page 25 of 36 25
Proof It is clear that the locus of a ∈ V where (a) has distinct zeroes is Zariski open.
To prove the lemma, we need to check that V is not empty. We do this constructively.
First assume C = P1 and L = OP1 (1). Set a2 = 0, a3 = c ∈ k, and a4 = t 4 . Then
= −27c4 − 64t 12 which has distinct zeroes as a section of OP1 (12) if c = 0.
Moreover, varying c, we can arrange for the zeros to avoid any finite subset of P1 .
In the general case, choose a morphism f : C → P1 such that L = f ∗ (OP1 (1)).
Let S ⊂ P1 be the branch locus of f . Then setting a2 = 0, a3 = c, and a4 = f ∗ (t 4 ),
where c is chosen so that the zeroes of −27c4 − 64t 12 are disjoint from S, yields an
explicit a with the required properties. Varying c allows us to avoid any finite subset
of C.
Lemma 6.3 For every n ≥ 1, there is a non-empty, Zariski open subset Vn of V such
that if a ∈ Vn , then the section P of Ea → C does not intersect any singular fiber in
a point of order exactly n.
Proof It is clear that the locus of a where P has the stated property is open, and our task
is to show it is non-empty. Since the bad fibers are all of type I1 , if k has characteristic
p > 0 and n is divisible by p, there are no points of order exactly n in the fiber, so we
may take Vn = V .
Now assume that n is not divisible by the characteristic of k. We check constructively
that there is a non-empty set as described in the statement. As in the previous lemma,
we may reduce to the case C = P1 and L = OP1 (1). Take a2 = 0, a3 = c, a4 = t 4 .
Then the bad fibers are at the roots of t 12 = (−27/64)c4 and at each such root,
the coordinate in Gm of P was given at (4.1). For the data we are considering, the
coordinate is
4t 4 − 3cγ
where γ = (−9c3 /2)1/2 t −2 .
4t 4 + 3cγ
Then for each n, there are only finitely many values of c such that for some root t of
t 12 = (−27/64)c4 , the displayed quantity is an n-th root of unity. This proves that Vn
is non-empty for each n.
Remark 6.4 Over an uncountable field, intersecting the opens in the theorem gives a
non-empty set. We can do a bit better over C: There is an everywhere dense classical
open set in V such that P meets each singular fiber away from the unit circle S 1 ⊂ C× .
Lemma 6.5 If the characteristic of k is p > 3, then for all a ∈ V and any n divisible
by p, P does not have order exactly n.
Ea → C is non-isotrivial) and not a p-th power. Then [30, Prop I.7.3] implies that
Ea → C has no p-torsion. (In [30], the ground field is finite, but the argument there
works over any field of positive characteristic.)
Proposition 6.6 For every n not divisible by the characteristic of k, there is a non-
empty, Zariski open subset Wn ⊂ V such that if a ∈ Wn , then n P = 0 and P is
transverse to Ea [n].
Proof Again, it is clear that the set of a with the desired properties is open. Unfortu-
nately, it seems hopeless to give a constructive proof that it is non-empty, so we have
to do something more sophisticated.
Recall the moduli space M of Sect. 4. We write Mk for
and Mk [n] for the locally closed, smooth, codimension 1 locus parameterizing triples
(E, ω, P) where P has order n.
Recall also that V = H 0 (C, L 2 ⊕ L 3 ⊕ L 4 ) and V is the open subset consisting of
a such that (a) has distinct zeroes. Choose an open subset U ⊂ C and a trivialization
of L over U . Then for a = (a2 , a3 , a4 ) the ai may be regarded as functions on U ,
and we get a morphism f a : U → Mk . To say that n P = O is to say that f a (U ) is
contained in Mk [n]. To say that P is tangent to Ea [n] over x ∈ U is to say that f a (U )
is tangent to Mk [n] at f a (x). We will show that these conditions do not hold for most
a.
Consider the morphism
and let
We will use the global generation of L to show that Dn is a smooth, locally closed subset
of codimension 1 in V × U , and that there is a non-empty open subset WU ,n ⊂ V
such that the projection Dn → V is étale over WU ,n . This means that if a ∈ WU ,n ,
then {a} × U is transverse to Dn , i.e., that P meets the n-torsion multisection of Ea
transversally over U . Taking a finite cover {U j } of C and setting Wn = ∩ j WU j ,n will
complete the proof.
Since L is globally generated, so are its powers L i for i = 2, 3, 4. This means that
for every t ∈ U , there are global sections a2 , a3 , a4 not vanishing at t, and for all but
finitely many t there are global sections s2 , s3 , s4 which vanish to order 1 at t. (Since
L i is globally generated, there are sections of L i inducing a morphism C → P1 . If t
is not in the ramification locus, a section si as above can be obtained by pulling back
a section of OP1 (1) vanishing simply at the point of P1 under t.)
For each t ∈ U , the restriction
Ft : V × {t} → Mk
Transversality of sections on elliptic surfaces Page 27 of 36 25
is a linear map, and since L is globally generated, it is surjective. Thus the fibers are
all affine spaces of dimension h − 3 where h = dim V . Therefore, F is surjective and
smooth (smooth because it is submersive, i.e., it has a surjective differential at every
point). Moreover, the fibers of F are Ah−3 -bundles over U , and in particular, they are
all irreducible of dimension h − 2. It follows that each irreducible component Dn,i of
Dn is smooth and locally closed in V × U of codimension 1 and has the form
is étale. The image of Dn,i \ Dn,i in V is contained in a proper closed subset, and
o
removing these subsets for all i yields an open subset WU ,n over which Dn → V is
étale. Covering C with finitely many U j and setting Wn = ∩ j WU j ,n yields an open
subset of V such that if a ∈ Wn , then P does not have order n and is transverse to
Ea [n]. This completes the proof of the proposition.
The preceding lemmas show that if a ∈ V , then the corresponding Ea has the proper-
ties asserted in the Theorem. Indeed, since a ∈ V , (a) as 12d distinct zeroes, and so
Ea has 12d bad fibers of type I1 and no other bad fibers. Since a ∈ ∩n Vn , Lemma 6.3
shows that P does not meet a bad fiber in a torsion point. Since a ∈ ∩n Wn , Lemma 6.5
25 Page 28 of 36 D. Ulmer, G. Urzúa
and Proposition 6.6 show that P has infinite order, and Proposition 6.6 shows that if
n is prime to the characteristic, then P is transverse to Ea [n]. This establishes points
(1) through (4) of the Theorem.
The transversality in point (5) is equivalent to that in (4), so to finish we just need
to calculate the intersection multiplicity (n P).O. For this, we first note that P.O = 0
by construction, and as explained in the proof of Lemma 2.8, O 2 = P 2 = −d. Thus
ht(P) = 2d and Lemma 2.8 implies that (n P).O = d(n 2 − 1), as required.
This completes the proof of the theorem.
Remark 6.7 When C = P1 , every line bundle of non-negative degree is globally gener-
ated. Thus, starting from data (E, P) over P1 , we can find a deformation (E , P ) with
the same base C and bundle L such that P is transverse to all torsion multisections.
For a general C, if we do not assume any positivity for L = O ∗ (1E /C ), it may be
impossible to produce deformations with fixed C and L. Here are two alternatives:
First, we may embed L → L where L is globally generated, and deform a non-
minimal model of E (lying in PC (L 2 ⊕ L 3 ⊕ OC )). Second, it seems likely that the
ideas of Moishezon [23], as explained in [16, Thm. I.4.8] would allow one to find a
deformation of E where the base curve is also allowed to vary (i.e., deform to E → C
and section P ) with the desired transversality.
Remark 6.8 Suppose that k has characteristic zero and that π : E → C and P satisfy
the conclusions of Theorem 6.1. If n 1 and n 2 are two distinct integers, then n 1 P ∪ n 2 P
is a normal crossings divisor on E. More generally, if N ⊂ Z is a non-empty finite set,
then
D= nP
n∈N
In this section, we show by explicit construction that there are pairs (E, P) with P
transverse to torsion multisections over fields k such as number fields and global func-
tion fields. The precise statement is Theorem 1.8 in the introduction. For simplicity,
we assume throughout that the characteristic of k is not 2. We begin by constructing
examples of height 2 over P1 .
Proposition 7.1 Let k be a field of characteristic = 2. Then there exist Jacobian elliptic
surfaces E → P1 over k equipped with a section P such that
(1) P has infinite order.
(2) The singular fibers of E → P1 are of Kodaira type I0∗ .
(3) P meets each singular fiber in a non-torsion point.
Transversality of sections on elliptic surfaces Page 29 of 36 25
points.
(5) The height of E is 2, i.e., O ∗ (1E /P1 ) ∼
= OP1 (2).
Proof We will construct one such E → P1 for every elliptic curve E over k. Suppose
that f ∈ k[x] is a monic polynomial of degree 3 such that E is defined by y 2 = f (x).
Form the product E ×k E, and let {±1} ⊂ Aut(E) act diagonally. The quotient
(E ×k E)/(±1) is a singular (Kummer) surface, and projection to the first factor
induces a morphism
(E ×k E)/(±1) → E/(±1) ∼
= P1 .
E ×k E (E ×k E)/(±1) E
n nP
E E/(±1) ∼
= P1 P1 .
over k(t), and the point P has coordinates (x, y) = (t f (t), f 2 (t)). Indeed, if the two
factors of E × E are v 2 = f (u) and s 2 = f (r ), then the field of invariants of ±1 is
generated by u, r , and z = vs, and these satisfy the equation
z 2 = f (u) f (r ).
Setting u = t and z = y/ f (u), and r = x/ f (u) yields the equation and point above.
25 Page 30 of 36 D. Ulmer, G. Urzúa
We now verify the cases n = 1 and n = 2 of the proposition. Since P has polynomial
coefficients, it does not meet O over any finite value of t, and since its x and y
coordinates have degrees 4 and 6, and E has height 2, P also does not meet O over
t = ∞. In summary, P meets O nowhere, as claimed. For later use, we note that at the
roots of f , P specializes to (0, 0), i.e., to a singular point of the fiber of (E ×k E)/(±1),
so P lands on a non-identity component of the fiber of E. At t = ∞, P specializes to
(1, 1), a non-singular, finite point of the fiber (i.e., a point not on O).
A tedious but straightforward calculation (or an algebra package …) shows that 2P
has coordinates ((1/4)t 4 + · · · , (1/8)t 6 + · · · ) where · · · indicates terms of lower
degree in t. The argument of the previous paragraph shows that 2P meets O nowhere,
as claimed. For later use, we note that 2P passes through a finite point of the identity
component in each of the bad fibers.
Now consider n > 2. It is clear that n meets E × {0} exactly at the points of E of
order n, and each of these intersections is transverse. If ( p, 0) is such a point which is
not of order 2, then the quotient map
E ×k E → (E ×k E)/(±1)
E → (E ×k E)/(±1)
such values of t.
It remains to consider what happens over the roots of f (t) and t = ∞. But we
checked above that P meets a non-trivial point of the identity component at t = ∞
and such a point is either of infinite order or of order p when k has characteristic p. So,
for n prime to the characteristic of k, n P does not meet O over t = ∞. Similarly, over
the roots of f (t), P passes through the non-identity component and 2P passes through
a non-trivial point of the identity component, so n P does not meet O when n is prime
to the characteristic. We have thus identified all points where n P and O intersect,
the intersections are transverse, and their number is as stated in the proposition. This
completes the proof of the proposition.
Remark 7.2 As a check, we compute the intersection number (n P).O using heights
as in Lemma 2.8. We have O 2 = P 2 = −2 and P.O = 0. Since P passes through
a non-identity component of the fibers over roots of f (t) and through the identity
component at t = ∞, the “correction term” is −C P .(P − O) = −3. (See table 1.19
in [13].) Using the formula (2.4) for the height pairing yields ht(P) = 1.
Transversality of sections on elliptic surfaces Page 31 of 36 25
Similarly, for any odd n, −Cn P .(n P − O) = −3 and using that ht(n P) = n 2 and
calculating as in Lemma 2.8 we find (n P).O = (n 2 − 1)/2.
On the other hand, for even n, n P passes through the identity component in all bad
fibers, so −Cn P .(n P − O) = 0 and we find that (n P).O = (n 2 − 4)/2.
This confirms that the intersections we saw above are all transverse.
Proof of Theorem 1.8 Proposition 7.1 implies the case of the Theorem where C = P1
and L = OP1 (2), and we get infinitely many examples because k is infinite. Indeed,
for each j ∈ k, there is an elliptic curve E with j-invariant j, and elliptic curves with
distinct j-invariants give rise to non-isomorphic E → P1 since the non-singular fibers
are twists of the chosen E.
We deduce the general case by a pull-back construction. Write E → P1 for one of
the surfaces constructed in Proposition 7.1. Let f : C → P1 be a non-constant mor-
phism defined by sections of the globally generated line bundle F, so F = f ∗ OP1 (1)
and L = f ∗ OP1 (2). The conclusions of the theorem will hold for E := f ∗ E → C if
the branch locus of f is disjoint from the set of points of P1 over which E has bad
reduction or n P meets O. From the construction of E , we see that the set to be avoided
is precisely the set of x coordinates of torsion points of the elliptic curve y 2 = f (x)
used to construct E . Although this set is infinite, we will see that it is sparse in k.
We divide into two cases according to the characteristic of k, starting with the case
of characteristic zero. Choose an elliptic curve E over Q, and an auxiliary prime such
that equations defining E are -integral and E has good reduction modulo . Then
[28, VIII.7.1] implies that the x-coordinate of a torsion point Q (defined over some
number field K and taken with respect to an -integral model) is “almost integral,” i.e.,
it satisfies 2 x(Q) is integral at all primes of K over . Construct E → P1Q using E as
in Proposition 7.1. Then choose any non-constant morphism f : C → P1k defined by
sections of F. Composing φ with a linear fractional transformation, we may arrange
that the branch locus of f consists of points with finite, non-zero coordinates, and
that any of those coordinates which lie in a number field have large denominators at
primes over . They are thus distinct from the x-coordinates of torsion points of E,
and E = f ∗ E satisfies the requirements of the theorem.
When k has characteristic p > 2, the argument is similar, but simpler: Choose
an embedding F p (t) → k, an elliptic curve E over F p (t), and a place v of F p (t)
where E has good reduction. Then by [29, §4], the coordinates of any torsion point
Q of E (defined over some algebraic extension K of F p (t) and taken with respect
to an integral model) are integral at places of K over v. Use E to construct E as in
Proposition 7.1. Then choose any non-constant morphism f : C → P1k defined by
sections of F. Composing φ with a linear fractional transformation, we may arrange
that the branch locus of f consists of points with finite, non-zero coordinates, and that
any of those coordinates which are algebraic over F p (t) are not integral at places over
v. They are thus distinct from the x-coordinates of torsion points of E, and E = f ∗ E
satisfies the requirements of the theorem.
Remark 7.3 It seems likely that when k is a number field or a global function field, the
construction in Proposition 7.1 gives rise to elliptic divisibility sequences Dn whose
“new parts” Dn are often irreducible, i.e., prime divisors.
25 Page 32 of 36 D. Ulmer, G. Urzúa
In this section, we will prove Theorem 1.9. Let k = C, C = P1 , and L = OP1 (d) where
d = g +1, which by assumption satisfies d ≥ 1. Theorem 1.7 guarantees the existence
of an elliptic surface π : E → P1 of height d (i.e., such that O ∗ (1E /P1 ) = L) with a
section P such that for all n, n P meets O transversally in d(n 2 − 1) points. Moreover,
π has irreducible fibers. Let F be the class of a fiber of π . We have O 2 = P 2 = −d,
F 2 = 0, and the canonical divisor of E is
K E = (d − 2)F.
Y \ (C0 ∪ Cn ) ∼
= X \ {x}.
1 a2 − 1
a− = .
a a
(See [8, Ch. 3].) In particular, it follows that X is Q-Gorenstein and K X is Q-Cartier.
We next compute the discrepancy of x (as defined for example in [19]) and verify
that x is log-terminal. Since C j is smooth and rational with self-intersection −a, we
Transversality of sections on elliptic surfaces Page 33 of 36 25
K Y = f ∗ K X + α0 C0 + αn Cn
0 = ( f ∗ C0 ).K X
= C0 . f ∗ K Y
= (a − 2) + α0 a − αn
and similarly,
0 = (a − 2) − α0 + αn a.
We find that
a−2
α0 = αn = − > −1.
a−1
This confirms that x is a log-terminal singularity, and we have
a−2
f ∗ K X = (d − 2) F̃ + Ei + (C0 + Cn ) .
a−1
i
K X2 = ( f ∗ K X )2
2
= (d − 2) F̃ + E i + b(C0 + Cn )
i
= dn 2 (4b − 2b2 − 1) + d + 1 + 4b2 − 12b.
f ∗ C = D + m 0 C0 + m n Cn
C j . f ∗ K X = f ∗ (C j ).K X = 0
25 Page 34 of 36 D. Ulmer, G. Urzúa
for j = 0, n.
For the former assertion, we make a case by case analysis of the possibilities for
D. They are:
• the strict transform F̃ of a general fiber of π , for which we have
If d > 2, we have
which is > e for all n ≥ 4, and this shows that Q̃. f ∗ K X > 0. If Q = P, then
and we find that Q̃. f ∗ K X > 0 for all n ≥ 5. (When Q = P,we can also calculate
directly that Q̃. f ∗ K X = d − 2 + b(dn 2 − 2dn) which goes to infinity with n.)
This completes the check that Q̃. f ∗ K X > 0 for all irreducible multisections Q̃
not equal to C0 or Cn .
Transversality of sections on elliptic surfaces Page 35 of 36 25
The itemized list completes the verification that K X is ample, and this finishes the
proof of the theorem.
The first-named author thanks Seoyoung Kim, Nicole Looper, and Joe Silverman for
helpful conversations at the 2019 AMS Mathematics Research Community meeting
in Whispering Pines, Rhode Island, and for pointing out [17] and its antecedents. He
also thanks the Simons Foundation for partial support in the form of Collaboration
Grant 359573. The second-named author thanks FONDECYT for support from grant
1190066. Both authors thank Matthias Schütt for comments and corrections, and they
thank Pietro Corvaja, Brian Lawrence, and Umberto Zannier for their comments on an
earlier version of this paper and their pointers to related literature, notably the preprint
[11].
References
1. Alekseev, V.A., Liu, V.: On accumulation points of volumes of log surfaces. Izv. Ross. Akad. Nauk
Ser. Mat. 83, 5–25 (2019)
2. Alexeev, V., Liu, W.: Log surfaces of Picard rank one from four lines in the plane. Eur. J. Math. 5,
622–639 (2019)
3. Alexeev, V., Liu, W.: Open surfaces of small volume. Algebr. Geom. 6, 312–327 (2019)
4. Alexeev, V.: Boundedness and K 2 for log surfaces. Int. J. Math. 5, 779–810 (1994)
5. Alexeev, V., Mori, S.: Bounding singular surfaces of general type, Algebra, arithmetic and geometry
with applications (West Lafayette, IN, 2000), 2004, pp 143–174
6. Artin, M.: Some numerical criteria for contractability of curves on algebraic surfaces. Am. J. Math.
84, 485–496 (1962)
7. Bădescu, L.: Algebraic surfaces, Universitext. Springer, New York (2001)
8. Barth, W.P., Hulek, K., Peters, C.A.M., Van de Ven, A.: Compact complex surfaces, Second, Ergebnisse
der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, vol.
4, Springer, Berlin (2004)
9. Bedford, E., Lyubich, M., Smillie, J.: Polynomial diffeomorphisms of C 2 . IV. The measure of maximal
entropy and laminar currents. Invent. Math. 112, 77–125 (1993)
10. Bierstone, E., Milman, P.D.: Semianalytic and subanalytic sets. Inst. Hautes Études Sci. Publ. Math.
67, 5–42 (1988)
11. Corvaja, P., Demeio, J., Masser, D., Zannier, U.: On the torsion values for sections of an elliptic scheme,
2019. Preprint arxiv:1909.01253
12. Corvaja, P., Masser, D., Zannier, U.: Torsion hypersurfaces on abelian schemes and Betti coordinates.
Math. Ann. 371, 1013–1045 (2018)
13. Cox, D.A., Zucker, S.: Intersection numbers of sections of elliptic surfaces. Invent. Math. 53, 1–44
(1979)
14. Deligne, P.: Courbes elliptiques: formulaire d’après J. Tate, Modular functions of one variable, IV
(Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), 1975, pp. 53–73. Lecture Notes
inMath., Vol. 476
15. Deligne, P., Rapoport, M.: Les schémas de modules de courbes elliptiques (1973), 143–316. Lecture
Notes in Math., Vol. 349
16. Friedman, R., Morgan, J.W.: Smooth Four-Manifolds and Complex Surfaces, Ergebnisse derMathe-
matik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 27. Springer,
Berlin (1994)
17. Ghioca, D., Hsia, L.-C., Tucker, T.: A variant of a theorem by Ailon-Rudnick for elliptic curves. Pacific
J. Math. 295, 1–15 (2018)
18. Ingram, P., Mahé, V., Silverman, J.H., Stange, K.E., Streng, M.: Algebraic divisibility sequences over
function fields. J. Aust. Math. Soc. 92, 99–126 (2012)
19. Kawamata, Y., Matsuda, K., Matsuki, K.: Introduction to the minimal model problem. Algeb. Geomet.
Sendai 1987, 283–360 (1985)
20. Kodaira, K.: On compact analytic surfaces II. Ann. Math. 77, 563–626 (1963)
25 Page 36 of 36 D. Ulmer, G. Urzúa
21. Kollár, J., Shepherd-Barron, N.I.: Threefolds and deformations of surface singularities. Invent. Math.
91, 299–338 (1988)
22. Liu, W.: The minimal volume of log surfaces of general type with positive geometric genus. Preprint
arxiv:1706.03716
23. Moishezon, B.: Complex surfaces and connected sums of complex projective planes, Lecture Notes
inMathematics, Vol. 603, Springer, Berlin, 1977.With an appendix by R. Livne
24. Łojasiewicz, S.: Ensembles semi-analytiques, 1964. IHES Preprint
25. Shioda, T.: On the Mordell-Weil lattices. Comment. Math. Univ. St. Paul. 39, 211–240 (1990)
26. Shioda, T.: Mordell-Weil lattices for higher genus fibration over a curve, New trends in algebraic
geometry (Warwick, 1996), 1999, pp. 359–373
27. Silverman, J.H.: Generalized greatest common divisors, divisibility sequences, and Vojtas conjecture
for blowups. Monatsh. Math. 145, 333–350 (2005)
28. Silverman, J.H.: The Arithmetic of Elliptic Curves, Second, Graduate Texts in Mathematics, vol. 106.
Springer, Dordrecht (2009)
29. Tate, J.T.: Algorithm for determining the type of a singular fiber in an elliptic pencil, Modular functions
of one variable, iv (Proc. Internat. Summer School, Univ. Antwerp, Antwerp, 1972), 1975, pp. 33–52.
Lecture Notes in Math., Vol. 476
30. Ulmer, D.: Elliptic curves over function fields, Arithmetic of L-functions (Park City, UT, 2009), 2011,
pp. 211–280
31. Ulmer, D.: On Mordell-Weil groups of Jacobians over function fields. J. Inst. Math. Jussieu 12, 1–29
(2013)
32. Ulmer, D., Urzúa, G.: Bounding tangencies of sections on elliptic surfaces. Int. Math. Res. Not. IMRN
6, 4768–4802 (2021)
33. Zannier, U.: Some problems of unlikely intersections in arithmetic and geometry, Annals of Mathe-
matics Studies, vol. 181, Princeton University Press, Princeton, NJ, 2012.With appendixes by David
Masser
Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
and institutional affiliations.