1 Introduction
This paper is concerned with the Hasse principle for the family of projective cubic surfaces

where
$f,g\in \mathbb{Q}[u,v]$
are binary cubic forms. Our main result shows that a positive proportion of these surfaces, when ordered by height, possess a
$\mathbb{Q}$
-rational point. First, we discuss the question of local solubility. The following result will be addressed in §2.2.
Theorem 1.1. Approximately 99% of the cubic surfaces (1.1), when ordered by height, are everywhere locally soluble.
It has long been known that the Hasse principle does not always hold for cubic surfaces and so local solubility is not enough to ensure that the surface (1.1) has a
$\mathbb{Q}$
-rational point. In the special case that
$f$
and
$g$
are diagonal there are many known counter-examples to the Hasse principle, the most famous being the surface

that was discovered by Cassels and Guy [Reference Cassels and Guy6]. As explained by Manin [Reference Manin16], this example is accounted for by the Brauer–Manin obstruction. It has been conjectured by Colliot-Thélène and Sansuc [Reference Colliot-Thélène and Sansuc7, §V] that this obstruction explains all failures of the Hasse principle for smooth cubic surfaces. This conjecture remains wide open in general. However, work of Swinnerton-Dyer [Reference Swinnerton-Dyer21] confirms it for a special family of diagonal cubic surfaces, conditionally under the assumption that the Tate–Shafarevich group of elliptic curves is finite.
A recent investigation of Bright [Reference Bright4] has focused on families of varieties over number fields that have no Brauer–Manin obstruction to the Hasse principle. In §2.1 we shall check that the conditions of his main result are satisfied for the family of cubic surfaces (1.1), thereby leading to the following conclusion.
Theorem 1.2 (Bright).
Assume that the Brauer–Manin obstruction is the only obstruction to the Hasse principle for smooth cubic surfaces. Then 100% of the cubic surfaces (1.1), when ordered by height, satisfy the Hasse principle.
There are very few results which establish the existence of
$\mathbb{Q}$
-rational points on cubic surfaces unconditionally. Given coprime
$a,b\in \mathbb{Z}$
, Heath-Brown and Moroz [Reference Heath-Brown and Moroz12, Corollary 1.2] have shown that any cubic surface

has a
$\mathbb{Q}$
-rational point, provided that either
$a\equiv \{\pm 2,\pm 3\}\hspace{0.2em}{\rm mod}\hspace{0.2em}9$
, or
$b\equiv$
$\{\pm 2,\pm 3\}\hspace{0.2em}{\rm mod}\hspace{0.2em}9$
, or
$a\equiv \pm b\hspace{0.2em}{\rm mod}\hspace{0.2em}9$
. This result combines sophisticated sieve methods, which allow primes to be represented by binary cubic polynomials, with a result of Satgé [Reference Satgé17, Proposition 3.3], which demonstrates that the curve
$u^{3}+2v^{3}=pz^{3}$
has a
$\mathbb{Q}$
-rational point for any prime
$p\equiv 2\hspace{0.2em}{\rm mod}\hspace{0.2em}9$
.
Our main result provides a much denser set of smooth cubic surfaces which are in possession of a rational point.
Theorem 1.3. A positive proportion of the cubic surfaces (1.1), when ordered by height, possess a
$\mathbb{Q}$
-rational point.
Recall that the set of
$\mathbb{Q}$
-rational points is Zariski dense on the cubic surface (1.1) as soon as it is non-empty, as proved by Segre [Reference Segre18].
Theorem 1.3 will be established in §3. Our proof follows a strategy of Bhargava [Reference Bhargava1, §4], which was developed to show that a positive proportion of plane cubic curves have a
$\mathbb{Q}$
-rational point. The key input is recent work of Kriz and Li [Reference Kriz and Li15], concerning the family of Mordell curves

for
$k\in \mathbb{Z}$
. Using a Heegner point construction, Kriz and Li show that a positive proportion of these elliptic curves have rank one. The curves
$E_{k}$
have also been the focus of novel work by Bhargava et al [Reference Bhargava, Elkies and Shnidman3] and we shall draw heavily on their investigation. It follows from [Reference Bhargava, Elkies and Shnidman3, Theorem 8], in particular, that at least 41.1% of the curves
$E_{k}$
, for
$k\in \mathbb{Z}$
, have rank one, but this is only achieved under the assumption that the Tate–Shaferevich group is finite, whereas the work of Kriz and Li is unconditional.
Associated to
$E_{k}$
is an obvious
$3$
-isogeny
$\unicode[STIX]{x1D711}$
and we shall monopolize on the correspondence between elements of the
$\unicode[STIX]{x1D711}$
-Selmer group and locally soluble binary cubic forms. Let
$V(\mathbb{Z})$
be the lattice of (integer-matrix) binary cubic forms

for
$a,b,c,d\in \mathbb{Z}$
. Associated to each
$f\in V(\mathbb{Z})$
is the genus-one plane curve

To establish Theorem 1.3, it will suffice to show that there is a positive proportion of irreducible
$f\in V(\mathbb{Z})$
, when ordered by height, such that
$C_{f}(\mathbb{Q})\neq \emptyset$
. To see this we first note that, since
$f$
is irreducible, we must have
$z\neq 0$
in any rational point
$(u:v:z)\in C_{f}(\mathbb{Q})$
. Now if the binary cubic forms
$f,g\in V(\mathbb{Z})$
give rise to the rational points
$(u:v:z)\in C_{f}(\mathbb{Q})$
and
$(u^{\prime }:v^{\prime }:z^{\prime })\in C_{g}(\mathbb{Q})$
, with
$zz^{\prime }\neq 0$
, then it follows that the rational point
$(z^{\prime }u:z^{\prime }v:zu^{\prime }:zv^{\prime })\in \mathbb{P}^{3}(\mathbb{Q})$
lies on the cubic surface (1.1).
Remark 1.4. It is also possible to say something about appropriate K3 surfaces. Consider the family of Kummer surfaces with affine equation

where
$g_{1},g_{2}$
are irreducible quartic polynomials defined over
$\mathbb{Q}$
. Recent work of Harpaz and Skorobogatov [Reference Harpaz and Skorobogatov11] establishes the Hasse principle for these surfaces, provided that
$g_{1},g_{2}$
satisfy suitable genericity conditions, albeit under an assumption about the finiteness of certain associated Tate–Shaferevich groups. It has been shown by Bhargava [Reference Bhargava1, Theorem 11] that a positive proportion of the genus-one curves
$y^{2}=g(u,v)$
possess a
$\mathbb{Q}$
-rational point, when the binary quartic forms
$g\in \mathbb{Q}[u,v]$
are ordered by height. An immediate consequence of this is the unconditional statement that a positive proportion of the Kummer surfaces (1.3), when ordered by height, possess a
$\mathbb{Q}$
-rational point.
2 The family of cubic surfaces and local solubility
Let
$X\subset \mathbb{P}^{7}\times \mathbb{P}^{3}$
be the biprojective hypersurface defined by

We obtain a flat surjective morphism
$\unicode[STIX]{x1D70B}:X\rightarrow \mathbb{P}^{7}$
by projecting to the point
$(a:b:\cdots :h)$
, with
$X$
smooth, projective and geometrically integral. We denote by
$X_{P}$
the fibre above any point
$P=(a:b:\cdots :h)\in \mathbb{P}^{7}(\mathbb{Q})$
. Let
$\unicode[STIX]{x1D702}:\operatorname{Spec}(K)\rightarrow \mathbb{P}^{7}$
denote the generic point and
$\overline{\unicode[STIX]{x1D702}}:\operatorname{Spec}(\overline{K})\rightarrow \mathbb{P}^{7}$
a geometric point above
$\unicode[STIX]{x1D702}$
, where
$K$
is the function field of
$\mathbb{P}^{7}$
and
$\overline{K}$
is an algebraic closure of
$K$
. The geometric generic fibre
$X_{\overline{\unicode[STIX]{x1D702}}}$
is a smooth cubic surface and so it is connected and has torsion-free Picard group. Denote by
$\overline{X}$
the base change of
$X$
to an algebraic closure
$\overline{\mathbb{Q}}$
of
$\mathbb{Q}$
and let
$\mathbf{A}_{\mathbb{Q}}$
denote the ring of adèles of
$\mathbb{Q}$
.
2.1 Proof of Theorem 1.2
The statement follows from [Reference Bright4, Theorem 1.4], provided that the following hypotheses are met:
-
(1)
$X(\mathbf{A}_{\mathbb{Q}})\neq \emptyset$ ;
-
(2) the fibre of
$\unicode[STIX]{x1D70B}$ at each codimension-one point of
$\mathbb{P}^{7}$ is geometrically integral;
-
(3) the fibre of
$\unicode[STIX]{x1D70B}$ at each codimension-two point of
$\mathbb{P}^{7}$ has a geometrically reduced component;
-
(4)
$\text{H}^{1}(\mathbb{Q},\operatorname{Pic}\overline{X})=0$ ;
-
(5)
$\operatorname{Br}\overline{X}$ is trivial;
-
(6)
$\text{H}^{2}(\mathbb{Q},\operatorname{Pic}\mathbb{P}^{7})\rightarrow \text{H}^{2}(\mathbb{Q},\operatorname{Pic}\overline{X})$ is injective;
-
(7)
$\operatorname{Br}X_{\overline{\unicode[STIX]{x1D702}}}=0$ .
Since
$X$
is
$\mathbb{Q}$
-rational the conditions (1), (4) and (5) are automatically met. The cubic surface
$X_{P}$
has equation
$f(x_{0},x_{1})=g(x_{2},x_{3})$
for suitable binary cubic forms
$f$
and
$g$
. If
$X_{P}$
fails to be geometrically integral then it must be singular, which can only happen when
$\operatorname{disc}(f)=\operatorname{disc}(g)=0$
. This is a codimension-two condition, which thereby establishes (2). The only way that (3) can fail is if
$f$
and
$g$
are proportional and both have repeated roots. But this is a codimension-three condition on the set of coefficients. Condition (6) follows from [Reference Bright, Browning and Loughran5, Proposition 5.17]. Finally,
$X_{\overline{\unicode[STIX]{x1D702}}}$
is a smooth cubic surface over an algebraically closed field of characteristic
$0$
. Thus it is rational and it follows that
$\operatorname{Br}X_{\overline{\unicode[STIX]{x1D702}}}=0$
, as required for condition (7). This completes the proof of Theorem 1.2.
Remark 2.1. Although we shall not need it in our work, it is possible to show that
$\operatorname{Br}X_{\unicode[STIX]{x1D702}}/\operatorname{Br}K=0$
for the family of cubic surfaces considered here, using an argument suggested to us by Jörg Jahnel. According to Swinnerton-Dyer [Reference Swinnerton-Dyer20], there is no Brauer–Manin obstruction whenever the action of
$\text{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})$
on the
$27$
lines is as large as possible. A binary cubic form splits into three linear forms, generically by an
$S_{3}$
-extension. Thus, generically, there is an
$S_{3}\times S_{3}$
-extension
$k/\mathbb{Q}$
, such that the surface may be written in the classical Cayley–Salmon form
$l_{1}l_{2}l_{3}=l_{4}l_{5}l_{6}$
, for binary linear forms
$l_{1},l_{2},l_{3}\in k[x_{0},x_{1}]$
and
$l_{4},l_{5},l_{6}\in k[x_{2},x_{3}]$
. There are nine obvious
$k$
-rational lines, given by
$l_{i}=l_{j}=0$
for
$i\in \{1,2,3\}$
and
$j\in \{4,5,6\}$
, and six obvious tritangent
$k$
-rational planes. The latter form a pair of Steiner trihedra, both of which are defined over
$\mathbb{Q}$
. The group
$S_{3}\times S_{3}\times S_{3}$
is the maximal Galois group stabilizing two such Steiner trihedra. Consider now the cubic surface
$S\subset \mathbb{P}^{3}$
with equation

A calculation in magma shows that the Galois group operating on the 27 lines is precisely the maximal possible group
$S_{3}\times S_{3}\times S_{3}$
. Since the Galois group has order 216, we may consult the appendix in work of Jahnel [Reference Jahnel13] to conclude that
$\operatorname{Br}S/\operatorname{Br}\mathbb{Q}=0.$
This suffices to show the claim.
2.2 Proof of Theorem 1.1: initial steps
Conditions (1) and (2) in §2.1 are sufficient to ensure that all of the hypotheses of [Reference Bright, Browning and Loughran5, Theorem 1.3] are met. Thus it follows that

where
$\unicode[STIX]{x1D717}_{v}$
is the density of cubic surfaces
$X_{P}$
which have
$\mathbb{Q}_{v}$
-points, for each place
$v$
of
$\mathbb{Q}$
. Clearly
$\unicode[STIX]{x1D717}_{\infty }=1$
. We shall show that
$\unicode[STIX]{x1D717}_{p}=1-\unicode[STIX]{x1D712}_{p}(1/p)$
for a suitable rational function
$\unicode[STIX]{x1D712}_{p}\in \mathbb{Q}(x)$
. Let

Then we shall show that

when
$p\equiv 1\hspace{0.2em}{\rm mod}\hspace{0.2em}3$
, while if
$p\not \equiv 1\hspace{0.2em}{\rm mod}\hspace{0.2em}3$
then

where

We note that
$1-\unicode[STIX]{x1D712}_{3}(1/3)=0.9965901\ldots \,$
. Inserting our expressions for
$\unicode[STIX]{x1D712}_{p}(1/p)$
into a computer and taking a product over primes
${\leqslant}10\,000$
we find that

Combining these, one verifies that
$\prod _{p}\unicode[STIX]{x1D717}_{p}$
has numerical value
$0.993782\ldots \,$
, which is satisfactory for Theorem 1.1.
Our calculation of
$\unicode[STIX]{x1D717}_{p}$
, for a given prime
$p$
, is based on work of Bhargava et al [Reference Bhargava, Cremona and Fisher2], who perform the same sort of calculation for the family of all ternary cubic forms. We shall say that a binary cubic form over
$\mathbb{F}_{p}$
is diagonal if it takes the shape
$au^{3}+bv^{3}$
for
$a,b\in \mathbb{F}_{p}^{\ast }$
. We begin by recording the following result.
Lemma 2.2.
-
(1) The probability that a random monic cubic over
$\mathbb{F}_{p}$ has a simple root in
$\mathbb{F}_{p}$ is
$\unicode[STIX]{x1D70E}_{1}={\textstyle \frac{2}{3}}(p^{2}-1)/p^{2}$ , and the probability that it has a triple root is
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$ .
-
(2) The probability that a random primitive binary cubic form over
$\mathbb{F}_{p}$ has a simple root in
$\mathbb{P}^{1}(\mathbb{F}_{p})$ is
$\unicode[STIX]{x1D70E}_{\ast }={\textstyle \frac{1}{3}}p(2p+1)/(p^{2}+1)$ , and the probability that it has a triple root is
$\unicode[STIX]{x1D70F}_{\ast }=1/(p^{2}+1)$ .
-
(3) The probability that a random diagonal binary cubic form over
$\mathbb{F}_{p}$ has a simple root in
$\mathbb{P}^{1}(\mathbb{F}_{p})$ is
$$\begin{eqnarray}\unicode[STIX]{x1D70E}_{2}=\left\{\begin{array}{@{}ll@{}}\frac{1}{3}\quad & \text{if }p\not \equiv 2\hspace{0.2em}{\rm mod}\hspace{0.2em}3,\\ 1\quad & \text{if }p\equiv 2\hspace{0.2em}{\rm mod}\hspace{0.2em}3.\end{array}\right.\end{eqnarray}$$
Proof. Part (1) is proved in [Reference Bhargava, Cremona and Fisher2, Corollary 4]. According to [Reference Bhargava, Cremona and Fisher2, Corollary 6], the probability that a random binary cubic form over
$\mathbb{F}_{p}$
has a simple root in
$\mathbb{P}^{1}(\mathbb{F}_{p})$
is
$\unicode[STIX]{x1D70E}={\textstyle \frac{1}{3}}(p^{2}-1)(2p+1)/p^{3}$
, and the probability that it has a triple root is
$\unicode[STIX]{x1D70F}=(p^{2}-1)/p^{4}$
. Part (2) therefore follows on noting that
$\unicode[STIX]{x1D70E}_{\ast }=\unicode[STIX]{x1D70E}p^{4}/(p^{4}-1)$
and
$\unicode[STIX]{x1D70F}_{\ast }=\unicode[STIX]{x1D70F}p^{4}/(p^{4}-1)$
. Consider now the special case of diagonal binary cubic forms over
$\mathbb{F}_{p}$
. The lemma follows on noting that a random element of
$\mathbb{Z}_{p}^{\ast }$
is a cube in
$\mathbb{Q}_{p}^{\ast }$
with probability
$\unicode[STIX]{x1D70E}_{2}$
.◻
We now proceed with our calculation of the density
$\unicode[STIX]{x1D717}_{p}$
for a given prime
$p$
. Over
$\mathbb{Q}_{p}$
, cubic surfaces of the shape (1.1) are determined by pairs
$(f,g)$
of binary cubic forms
$f\in \mathbb{Z}_{p}[u,v]$
and
$g\in \mathbb{Z}_{p}[x,y]$
, where

We say that the surface
$(f,g)$
is soluble over
$\mathbb{Q}_{p}$
if there exist
$u,v,x,y\in \mathbb{Z}_{p}$
, with
$\min \{v_{p}(u),v_{p}(v),v_{p}(x),v_{p}(y)\}=0$
, such that
$f(u,v)=g(x,y)$
. We say that a binary cubic form over
$\mathbb{Z}_{p}$
is primitive if not all of its coefficients are divisible by
$p$
. We are only interested in surfaces
$(f,g)$
defined over
$\mathbb{Z}_{p}$
for which
$f$
or
$g$
is primitive.
Mimicking [Reference Bhargava, Cremona and Fisher2], we shall say that the surface
$(f,g)$
has valuations

if we happen to know that
$v_{p}(c_{i})\geqslant \unicode[STIX]{x1D711}_{i}$
and
$v_{p}(d_{i})\geqslant \unicode[STIX]{x1D6FE}_{i}$
, for
$0\leqslant i\leqslant 3$
. Likewise, we shall replace the inequality symbol with equality if and only if we know the exact
$p$
-adic valuation of the relevant coefficient. We may clearly assume without loss of generality that
$\min _{0\leqslant i\leqslant 3}v_{p}(c_{i})=0$
(i.e.
$f$
is primitive). Through a change of variables, furthermore, we will also be able to restrict attention to the case where
$\min _{0\leqslant i\leqslant 3}v_{p}(d_{i})=\unicode[STIX]{x1D709}\in \{0,1,2\}$
, in which case we write
$g=p^{\unicode[STIX]{x1D709}}\tilde{g}$
for a primitive form
$\tilde{g}$
.
We shall examine the solubility of
$(f,g)=(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$
by considering its reduction modulo
$p$
. Let
$\overline{f}$
(respectively,
$\overline{g}$
) be the reduction of
$f$
(respectively,
$\tilde{g}$
) modulo
$p$
. Whenever
$\overline{f}$
has a simple root over
$\mathbb{F}_{p}$
, then this root can be lifted to a root over
$\mathbb{Z}_{p}$
and it follows that the cubic surface
$(f,g)$
is soluble over
$\mathbb{Q}_{p}$
. We claim the same is true when
$\overline{g}$
has a simple root. This is obvious when
$\unicode[STIX]{x1D709}=0$
and so we suppose that
$\unicode[STIX]{x1D709}>0$
. In this case we merely substitute
$pu,pv$
for
$u,v$
and divide by
$p^{\unicode[STIX]{x1D709}}$
to obtain a new cubic surface
$(p^{3-\unicode[STIX]{x1D709}}f,\tilde{g})$
, which is plainly soluble over
$\mathbb{Q}_{p}$
.
Suppose now that
$\overline{f}$
does not have any
$\mathbb{F}_{p}$
-roots and that
$\overline{g}$
does not have a simple root. Then, either
$\unicode[STIX]{x1D709}=0$
, in which case
$(\overline{f},\overline{g})$
is absolutely irreducible and it follows that the surface has a smooth
$\mathbb{F}_{p}$
-point which can be lifted, or else
$\unicode[STIX]{x1D709}>0$
. But in the latter case, substituting
$pu,pv$
for
$u,v$
and dividing by
$p^{\unicode[STIX]{x1D709}}$
, we obtain a new cubic surface
$(p^{3-\unicode[STIX]{x1D709}}f,\tilde{g})$
. If
$\overline{g}$
does not have any
$\mathbb{F}_{p}$
-roots then certainly
$(f,g)$
is not soluble over
$\mathbb{Q}_{p}$
. Next, suppose that
$\overline{f}$
has a triple root and
$\overline{g}$
has no
$\mathbb{F}_{p}$
-roots. If
$\unicode[STIX]{x1D709}=0$
then once again the surface
$(\overline{f},\overline{g})$
has a smooth
$\mathbb{F}_{p}$
-point which can be lifted to ensure solubility over
$\mathbb{Q}_{p}$
.
For each
$\unicode[STIX]{x1D709}\in \{0,1,2\}$
, we define the following probabilities:
-
∙
$\unicode[STIX]{x1D6FC}_{1}(\unicode[STIX]{x1D709})$ is the probability of solubility for a pair
$(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$ such that both
$\overline{f}$ and
$\overline{g}$ have a triple root;
-
∙
$\unicode[STIX]{x1D6FC}_{2}(\unicode[STIX]{x1D709})$ is the probability of solubility for a pair
$(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$ such that
$\overline{f}$ has a triple root and
$\overline{g}$ is insoluble;
-
∙
$\unicode[STIX]{x1D6FC}_{3}(\unicode[STIX]{x1D709})$ is the probability of solubility for a pair
$(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$ such that
$\overline{f}$ is insoluble and
$\overline{g}$ has a triple root;
-
∙
$\unicode[STIX]{x1D6FC}_{4}(\unicode[STIX]{x1D709})$ is the probability of solubility for a pair
$(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$ such that
$\overline{f}$ and
$\overline{g}$ are both insoluble.
Our work above already implies that

Next we claim that
$\unicode[STIX]{x1D6FC}_{3}(\unicode[STIX]{x1D709})=\unicode[STIX]{x1D6FC}_{2}(3-\unicode[STIX]{x1D709})$
for
$\unicode[STIX]{x1D709}\in \{1,2\}$
. Indeed, over
$\mathbb{F}_{p}$
, the only solutions have
$u=v=0$
. Substituting
$pu$
for
$u$
and
$pv$
for
$v$
and dividing by
$p^{2}$
we are left with the cubic surface
$(p^{3-\unicode[STIX]{x1D709}}f,\tilde{g})$
. This establishes the claim, implying that

for any
$\unicode[STIX]{x1D709}\in \{1,2\}$
.
We now return to our calculation of
$\unicode[STIX]{x1D717}_{p}$
. For each
$1\leqslant i\leqslant 4$
, we let
$\unicode[STIX]{x1D6FD}_{i}(\unicode[STIX]{x1D709})$
be the probability that our pair of cubic forms
$(f,g)=(f,p^{\unicode[STIX]{x1D709}}\tilde{g})$
has the reduction type considered in the definition of
$\unicode[STIX]{x1D6FC}_{i}(\unicode[STIX]{x1D709})$
. Let
$S\subset \mathbb{Z}_{p}^{4}$
be a set of coefficients producing primitive binary cubic forms. Then for any
$\unicode[STIX]{x1D709}\in \{0,1,2\}$
, the probability that a binary cubic form takes the shape
$p^{k}\tilde{g}$
for
$k\equiv \unicode[STIX]{x1D709}\hspace{0.2em}{\rm mod}\hspace{0.2em}3$
, where
$\tilde{g}$
has coefficient vector in
$S$
, is

where
$\unicode[STIX]{x1D707}(S)$
is the probability that a primitive binary cubic form has coefficient vector in
$S$
. It now follows from Lemma 2.2 that
$\unicode[STIX]{x1D6FD}_{i}(\unicode[STIX]{x1D709})=p^{-4\unicode[STIX]{x1D709}}\unicode[STIX]{x1D6FD}_{i}/(1-1/p^{12})$
, for
$1\leqslant i\leqslant 4$
, where

Hence we deduce that the probability of insolubility of a cubic surface (1.1) with coefficients in
$\mathbb{Q}_{p}$
is equal to

It remains to analyse
$\unicode[STIX]{x1D6FC}_{1}(\unicode[STIX]{x1D709})$
for
$\unicode[STIX]{x1D709}\in \{0,1,2\}$
and
$\unicode[STIX]{x1D6FC}_{2}(1),\unicode[STIX]{x1D6FC}_{2}(2)$
. In doing so, when the reduction of the form has a triple point, the density will not depend on what this triple point is and so we shall always take it to be
$(1:0)$
.
In the special case
$\unicode[STIX]{x1D6FC}_{1}(1)$
, which corresponds to surfaces
$(f,p\tilde{g})$
, where
$\overline{f}$
and
$\overline{g}$
both have triple roots, we may assume that
$\overline{f}$
and
$\overline{g}$
both have the triple root
$(1:0)$
. Substituting
$pv$
for
$v$
and dividing by
$p$
, it is easily seen that
$\unicode[STIX]{x1D6FC}_{1}(1)$
is equal to the probability of solubility over
$\mathbb{Q}_{p}$
for a cubic surface with valuations

in which one is only interested in solutions with
$\min \{v_{p}(u),v_{p}(x)\}=0$
.
As in [Reference Bhargava, Cremona and Fisher2], we will develop recursions in order to solve for the densities of soluble cubics among those whose reductions produce the bad configurations outlined above.
2.3 Technical lemmas
The purpose of this section is to collect together definitions and calculations of particular probabilities that will feature in our analysis.
Lemma 2.3. Let
$\unicode[STIX]{x1D706}$
be the probability that a cubic surface
$(f,g)$
with valuations

is soluble over
$\mathbb{Z}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. Then

Proof. If
$v_{p}(d_{1})=0$
(which happens with probability
$1-1/p$
) there is a simple root which can be lifted to a
$\mathbb{Q}_{p}$
-point on the cubic surface with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0})=0$
and
$v_{p}(d_{1})\geqslant 1$
(probability
$(p-1)/p^{2}$
) then the equation is insoluble over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0}),v_{p}(d_{1})\geqslant 1$
(probability
$1/p^{2}$
) then we divide by
$p$
to get the valuations

With probability
$\unicode[STIX]{x1D70E}_{1}$
we obtain a simple root of
$g$
over
$\mathbb{F}_{p}$
which can be lifted. Note that if
$\overline{g}$
is insoluble (probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
) then the equation
$\overline{f}=\overline{g}$
is absolutely irreducible and so has a smooth
$\mathbb{F}_{p}$
-point which can be lifted. Finally, with probability
$\unicode[STIX]{x1D70F}_{1}$
there is a triple root over
$\mathbb{F}_{p}$
, which we may take to be
$(1:0)$
. Now with probability
$\unicode[STIX]{x1D70E}_{2}$
the surface is soluble over
$\mathbb{Z}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. With probability
$1-\unicode[STIX]{x1D70E}_{2}$
the diagonal cubic form is insoluble over
$\mathbb{F}_{p}$
and we replace
$v$
by
$pv$
and
$y$
by
$pv$
. Dividing by
$p$
leads us to the valuations

If
$v_{p}(d_{0})=0$
(probability
$1-1/p$
) then we get solubility over
$\mathbb{Q}_{p}$
with probability
$\unicode[STIX]{x1D70E}_{2}$
, since when the diagonal cubic form is insoluble over
$\mathbb{F}_{p}$
we will not obtain solutions with
$x\in \mathbb{Z}_{p}^{\ast }$
. Alternatively, if
$v_{p}(d_{0})\geqslant 1$
(probability
$1/p$
) then we replace
$u$
by
$pu$
and divide once more by
$p$
to get the valuations

for which the probability of solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
is
$\unicode[STIX]{x1D706}$
. All in all it follows that

The lemma follows on rearranging and recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
.◻
Lemma 2.4. Let
$\unicode[STIX]{x1D71A}$
be the probability that a cubic surface
$(f,g)$
with valuations

is soluble over
$\mathbb{Z}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. Then

Proof. The first case to consider is when
$v_{p}(d_{0})=0$
(which happens with probability
$(p-1)/p$
). In this case the cubic surface is soluble over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
with probability
$\unicode[STIX]{x1D70E}_{2}$
, else it is insoluble over
$\mathbb{Q}_{p}$
. Suppose next that
$v_{p}(d_{0})\geqslant 1$
(probability
$1/p$
). Then we must have
$v_{p}(u)\geqslant 1$
. Replacing
$u$
by
$pu$
and simplifying, we get a cubic surface with valuations

When
$v_{p}(d_{1})=0$
(probability
$(p-1)/p$
) this is soluble over
$\mathbb{Q}_{p}$
. When
$v_{p}(d_{0})=0$
and
$v_{p}(d_{1})\geqslant 1$
(probability
$(p-1)/p^{2}$
) this is soluble over
$\mathbb{Q}_{p}$
with probability
$\unicode[STIX]{x1D70E}_{2}$
. Finally, we suppose that
$v_{p}(d_{0}),v_{p}(d_{1})\geqslant 1$
(probability
$1/p^{2}$
). In this case we must have
$v_{p}(v)\geqslant 1$
. Replacing
$v$
by
$pv$
and simplifying we get the valuations

With probability
$\unicode[STIX]{x1D70E}_{1}$
the second cubic form has a simple root which can be lifted to get solubility over
$\mathbb{Q}_{p}$
. With probability
$\unicode[STIX]{x1D70F}_{1}$
we get a triple root, which on moving to
$(1:0)$
leads us to replace
$y$
by
$py$
. On simplification, this gives a cubic surface with valuations whose probability of solubility is
$\unicode[STIX]{x1D71A}.$
All in all, it follows that

Rearranging this and recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
, the lemma follows.◻
Lemma 2.5. Let
$\unicode[STIX]{x1D714}$
be the probability that a cubic surface
$(f,g)$
with valuations

is soluble over
$\mathbb{Z}_{p}$
with
$\min \{v_{p}(u),v_{p}(x)\}=0$
. Then

Proof. With probability
$\unicode[STIX]{x1D70E}_{2}$
there is a solution over
$\mathbb{F}_{p}$
which can be lifted to ensure solubility over
$\mathbb{Q}_{p}$
. Otherwise, with probability
$1-\unicode[STIX]{x1D70E}_{2}$
, the only solution has
$v_{p}(u),v_{p}(y)\geqslant 1$
. Replacing
$u$
by
$pu$
,
$y$
by
$py$
and simplifying, we arrive at the valuations

in which we are interested in solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0})=0$
(probability
$(p-1)/p$
) then there are no solutions. If, on the other hand,
$v_{p}(d_{0})\geqslant 1$
then we simplify to get the valuations

Now if
$v_{p}(d_{1})=0$
(probability
$(p-1)/p$
) then there is a simple
$\mathbb{F}_{p}$
-root of
$g$
which can be lifted to get solubility over
$\mathbb{Q}_{p}$
. If
$v_{p}(d_{1})\geqslant 1$
and
$v_{p}(d_{0})=0$
(probability
$(p-1)/p^{2}$
) then the cubic surface is soluble over
$\mathbb{Q}_{p}$
with probability
$\unicode[STIX]{x1D70E}_{2}$
. If
$v_{p}(d_{0}),v_{p}(d_{1})\geqslant 1$
(probability
$1/p^{2}$
) then we replace
$v$
by
$pv$
and simplify, in order to get the valuations

With probability
$\unicode[STIX]{x1D70E}_{1}$
the second polynomial has a simple root over
$\mathbb{F}_{p}$
which can be lifted to get solubility over
$\mathbb{Q}_{p}$
. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
the form
$\overline{g}$
is irreducible over
$\mathbb{F}_{p}$
and we also get solubility over
$\mathbb{Q}_{p}$
. Finally, with probability
$\unicode[STIX]{x1D70F}_{1}$
, the form
$\overline{g}$
has a triple root. Without loss of generality we may assume that the root is
$(1:0)$
and the probability of solubility over
$\mathbb{Q}_{p}$
for a cubic surface with these valuations is
$\unicode[STIX]{x1D714}$
.
Putting everything together, we obtain

The lemma follows on rearranging and recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
.◻
Lemma 2.6. For
$i\in \{0,1\}$
let
$\unicode[STIX]{x1D6FE}_{i}$
be the probability that a cubic surface
$(f,g)$
is soluble over
$\mathbb{Z}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
, assuming that
$\overline{f}$
is irreducible over
$\mathbb{F}_{p}$
and the surface has valuations

Then

Proof. If
$v_{p}(d_{0})=0$
(probability
$1-1/p$
) then the cubic surface is soluble over
$\mathbb{Q}_{p}$
. If
$v_{p}(d_{0})\geqslant 1$
and
$v_{p}(d_{1})=0$
(probability
$(1-i)(1-1/p)/p$
) then we also get solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0}),v_{p}(d_{1})\geqslant 1$
(probability
$(1+i(p-1))/p^{2}$
) then we replace
$u$
by
$pu$
,
$v$
by
$pv$
and simplify to get the valuations

We seek solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
.
Suppose first that
$i=0$
. Then with probability
$\unicode[STIX]{x1D70E}_{1}$
the second cubic form
$g$
has a simple
$\mathbb{F}_{p}$
-root which can be lifted. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
it is not soluble over
$\mathbb{Q}_{p}$
and with probability
$\unicode[STIX]{x1D70F}_{1}$
the form
$g$
has a triple root, which we can move to
$(1:0).$
Replacing
$y$
by
$py$
we are led to the new valuations

If
$v_{p}(d_{0})=0$
then there are no solutions. If
$v_{p}(d_{0})\geqslant 1$
(probability
$1/p$
) then we simplify and observe that the probability of solubility is now
$\unicode[STIX]{x1D6FE}_{0}$
.
Suppose next that
$i=1$
. If
$v_{p}(d_{1})=0$
(probability
$1-1/p$
) then we get solubility over
$\mathbb{Q}_{p}$
. If
$v_{p}(d_{1})\geqslant 1$
and
$v_{p}(d_{0})=0$
then the surface is not soluble over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0}),v_{p}(d_{1})\geqslant 1$
(probability
$1/p^{2}$
) then we simplify to arrive at the valuations

With probability
$\unicode[STIX]{x1D70E}_{1}$
the second cubic form
$g$
has a simple
$\mathbb{F}_{p}$
-root which can be lifted to ensure the desired solubility over
$\mathbb{Q}_{p}$
. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
it is not soluble over
$\mathbb{Q}_{p}$
and with probability
$\unicode[STIX]{x1D70F}_{1}$
the form
$g$
has a triple root, which we can move to
$(1:0).$
Replacing
$y$
by
$py$
we are led to the a cubic surface for which the probability of solubility with
$x\in \mathbb{Z}_{p}^{\ast }$
is precisely
$\unicode[STIX]{x1D6FE}_{1}$
.
Thus, for
$i\in \{0,1\}$
, we have shown that

The lemma follows on rearranging and recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
.◻
Lemma 2.7. Let
$\unicode[STIX]{x1D716}$
be the probability that a cubic surface
$(f,g)$
with valuations

is soluble over
$\mathbb{Z}_{p}$
with
$\min \{v_{p}(u),v_{p}(x)\}=0$
. Then
$\unicode[STIX]{x1D716}=A+\unicode[STIX]{x1D6FC}_{1}(1)/p^{4}$
, where

Proof. If
$v_{p}(c_{1})=0$
(probability
$(p-1)/p$
) then the surface is soluble over
$\mathbb{Q}_{p}$
.
If
$v_{p}(c_{1})\geqslant 1$
and
$v_{p}(c_{0})=0$
(probability
$(p-1)/p^{2}$
) then with probability
$\unicode[STIX]{x1D70E}_{2}$
the surface is soluble over
$\mathbb{Q}_{p}$
. Alternatively (probability
$1-\unicode[STIX]{x1D70E}_{2}$
) we must have
$v_{p}(u),v_{p}(y)\geqslant 1$
in any solution. Replacing
$u$
by
$pu$
,
$v$
by
$pv$
and simplifying, we are led to the valuations

in which we are interested in solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
. If
$v_{p}(d_{0})=0$
(probability
$(p-1)/p$
) then this is soluble over
$\mathbb{Q}_{p}$
with probability
$\unicode[STIX]{x1D70E}_{2}$
. If
$v_{p}(d_{0})\geqslant 1$
(probability
$1/p$
) then we replace
$v$
by
$pv$
and simplify to get valuations for which the probability of solubility for such cubic surfaces is precisely
$\unicode[STIX]{x1D706}$
, in the notation of Lemma 2.3.
We now turn to the case
$v_{p}(c_{1}),v_{p}(c_{0})\geqslant 1$
(probability
$1/p^{2}$
) in the original equation. Replacing
$y$
by
$py$
and simplifying we arrive at the valuations

in which we are interested in solubility over
$\mathbb{Q}_{p}$
with
$\min \{v_{p}(u),v_{p}(x)\}=0$
. With probability
$\unicode[STIX]{x1D70E}_{1}$
the first cubic form has a simple root which can be lifted. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
the first cubic form is irreducible
$\mathbb{F}_{p}$
and the probability of solubility is equal to
$\unicode[STIX]{x1D6FE}_{1}$
, in the notation of Lemma 2.6. Finally, with probability
$\unicode[STIX]{x1D70F}_{1}$
it has a triple root, which we may move
$(1:0).$
In this final case, in view of our earlier discussions, the probability of solubility is equal to
$\unicode[STIX]{x1D6FC}_{1}(1)$
. The lemma easily follows on recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
.◻
2.4 Calculation of
$\unicode[STIX]{x1D6FC}_{2}(1)$
and
$\unicode[STIX]{x1D6FC}_{2}(2)$
For
$\unicode[STIX]{x1D709}\in \{1,2\}$
we proceed to calculate
$\unicode[STIX]{x1D6FC}_{2}(\unicode[STIX]{x1D709})$
. The density does not depend on the triple point that
$\overline{f}$
has and so we may take it to be
$(1:0)$
. Substituting
$pv$
for
$v$
and dividing by
$p$
we arrive at the cubic surface
$(f^{\prime },p^{\unicode[STIX]{x1D709}-1}\tilde{g})$
where the reduction
$\overline{g}$
of
$\tilde{g}$
modulo
$p$
is insoluble and the coefficients of
$f^{\prime }$
have valuations

If
$\unicode[STIX]{x1D709}=1$
then clearly
$\unicode[STIX]{x1D6FC}_{2}(1)=\unicode[STIX]{x1D6FE}_{1}$
, in the notation of Lemma 2.6. Suppose next that
$\unicode[STIX]{x1D709}=2$
. If
$v_{p}(c_{0})=0$
then we must have
$v_{p}(u)\geqslant 1$
and a change of variables from
$u$
to
$pu$
leads us to the conclusion that the surface is insoluble over
$\mathbb{Q}_{p}$
since
$\overline{g}$
is insoluble over
$\mathbb{F}_{p}$
. If
$v_{p}(c_{0})\geqslant 1$
(probability
$1/p$
) then we divide by
$p$
to get the cubic surface
$(f^{\prime \prime },\tilde{g})$
, in which the coefficients of
$f^{\prime \prime }$
have valuations

The probability of solubility in this case is precisely
$\unicode[STIX]{x1D6FE}_{0}$
, in the notation of Lemma 2.6. Thus we have shown that

2.5 Calculation of
$\unicode[STIX]{x1D6FC}_{1}(0)$
We may assume that the coefficients of
$(f,g)$
have valuations

With probability
$\unicode[STIX]{x1D70E}_{2}$
the equation
$c_{3}v^{3}=d_{3}y^{3}$
has a simple root over
$\mathbb{F}_{p}$
which can lifted, thereby ensuring solubility over
$\mathbb{Q}_{p}$
. With probability
$1-\unicode[STIX]{x1D70E}_{2}$
it is insoluble over
$\mathbb{F}_{p}$
. In the latter case we replace
$v$
by
$pv$
,
$y$
by
$py$
and simplify.
Suppose first that
$\min \{v_{p}(c_{0}),v_{p}(d_{0})\}\geqslant 2$
(which happens with probability
$1/p^{2}$
). Then we get a new pair of cubic forms with valuations

If
$v_{p}(c_{1})=0$
or
$v_{p}(d_{1})=0$
(probability
$1-1/p^{2}$
) then there is a simple root over
$\mathbb{F}_{p}$
which can be lifted. If, on the other hand,
$v_{p}(c_{1}),v_{p}(d_{1})\geqslant 1$
(probability
$1/p^{2}$
) then the probability of solubility is
$\unicode[STIX]{x1D705}$
, where
$\unicode[STIX]{x1D705}$
denotes the probability that a pair of cubics with valuations

is soluble over
$\mathbb{Z}_{p}$
with
$\min \{v_{p}(u),v_{p}(x)\}=0$
. We claim that

where
$\unicode[STIX]{x1D705}_{1}$
is the probability of solubility for a cubic surface with valuations

with
$x\in \mathbb{Z}_{p}^{\ast }$
. To see this, if
$v_{p}(c_{0})=v_{p}(d_{0})=0$
(probability
$(1-1/p)^{2}$
), then with probability
$\unicode[STIX]{x1D70E}_{2}$
we get a simple root over
$\mathbb{F}_{p}$
which can be lifted (and it is insoluble over
$\mathbb{Q}_{p}$
otherwise). If
$v_{p}(c_{0}),v_{p}(d_{0})\geqslant 1$
(probability
$1/p^{2}$
) then we simplify and obtain the valuations

With probability
$\unicode[STIX]{x1D70E}_{1}$
the cubic form
$f$
has a simple
$\mathbb{F}_{p}$
-root which can be lifted to get solubility over
$\mathbb{Q}_{p}$
. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
the form
$f$
is irreducible over
$\mathbb{F}_{p}$
and we also get solubility over
$\mathbb{Q}_{p}$
. Finally, with probability
$\unicode[STIX]{x1D70F}_{1}$
the form
$f$
has a triple root. In this case, with probability
$1-\unicode[STIX]{x1D70F}_{1}$
we get solubility over
$\mathbb{Q}_{p}$
, since either
$g$
has a simple root over
$\mathbb{F}_{p}$
or it is irreducible over
$\mathbb{F}_{p}$
. Alternatively, with probability
$\unicode[STIX]{x1D70F}_{1}$
the form
$g$
also has a triple root over
$\mathbb{F}_{p}$
and the probability of solubility is precisely
$\unicode[STIX]{x1D6FC}_{1}(0)$
. The claim follows on recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
and noting that the probability of solubility is
$\unicode[STIX]{x1D705}_{1}$
when
$v_{p}(c_{0})=0,v_{p}(d_{0})\geqslant 1$
or
$v_{p}(c_{0})\geqslant 1,v_{p}(d_{0})=0$
.
Next we calculate
$\unicode[STIX]{x1D705}_{1}$
, noting that we must have
$v_{p}(u)\geqslant 1$
in any solution. Replacing
$u$
by
$pu$
and simplifying we arrive at the valuations

With probability
$1-\unicode[STIX]{x1D70F}_{1}$
this is soluble over
$\mathbb{Q}_{p}$
. With probability
$\unicode[STIX]{x1D70F}_{1}$
the second form
$g$
has a triple root over
$\mathbb{F}_{p}$
which we can move to
$(1:0)$
. But then with probability
$\unicode[STIX]{x1D70E}_{2}$
the cubic surface is soluble over
$\mathbb{Q}_{p}^{\ast }$
and with probability
$1\,-\,\unicode[STIX]{x1D70E}_{2}$
the only solution has
$v_{p}(v),v_{p}(y)\geqslant 1$
. Replacing
$v$
by
$pv$
,
$y$
by
$py$
and simplifying, we therefore obtain the valuations

If
$v_{p}(d_{0})=0$
there are no solutions over
$\mathbb{Q}_{p}^{\ast }$
, while if
$v_{p}(d_{0})\geqslant 1$
we may remove a factor of
$p$
from all of the coefficients. In the new pair of forms we get solubility if
$v_{p}(d_{1})=0$
(probability
$1-1/p$
), while we get solubility with probability
$\unicode[STIX]{x1D70E}_{2}$
if
$v_{p}(d_{1})\geqslant 1$
and
$v_{p}(d_{0})=0$
(probability
$(p-1)/p^{2}$
). Finally, the probability of solubility is
$\unicode[STIX]{x1D705}_{1}$
if
$v_{p}(d_{1}),v_{p}(d_{0})\geqslant 1$
(probability
$1/p^{2}$
). Recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
and rearranging, we easily conclude that

Next, suppose that
$v_{p}(c_{0})=1$
and
$v_{p}(d_{0})\geqslant 2$
(probability
$(p-1)/p^{2}$
). Then our cubic surface has valuations

Hence we replace
$u$
by
$pu$
and divide by
$p$
to get the new valuations

But the probability that a pair of cubics with these valuations is soluble over
$\mathbb{Z}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
is precisely equal to
$\unicode[STIX]{x1D706}$
in the notation of Lemma 2.3. The case in which
$v_{p}(c_{0})\geqslant 2$
and
$v_{p}(d_{0})=1$
yields the same conclusion.
Finally, we suppose that
$v_{p}(c_{0})=v_{p}(d_{0})=1$
(probability
$(p-1)^{2}/p^{2}$
). This leads to the valuations

With probability
$\unicode[STIX]{x1D70E}_{2}$
the equation
$c_{0}u^{3}=d_{0}x^{3}$
has a simple root over
$\mathbb{F}_{p}$
which can be lifted. The alternative case leads to insolubility over
$\mathbb{Q}_{p}$
.
Summarizing our argument so far, we have shown that

Our calculations show that
$\unicode[STIX]{x1D705}=B+\unicode[STIX]{x1D6FC}_{1}(0)/p^{6}$
, where

Substituting
$\unicode[STIX]{x1D705}$
and rearranging, we finally conclude that

2.6 Calculation of
$\unicode[STIX]{x1D6FC}_{1}(1)$
In this section we calculate the value of
$\unicode[STIX]{x1D6FC}_{1}(1)$
recorded at the close of §2.2. If
$v_{p}(c_{0})=0$
(probability
$1-1/p$
) then the probability of solubility is precisely
$\unicode[STIX]{x1D714}$
, in the notation of Lemma 2.5. Alternatively, with probability
$1/p$
, we have
$v_{p}(c_{0})\geqslant 1$
. Replacing
$y$
by
$py$
and simplifying, we deduce that

where
$\unicode[STIX]{x1D714}_{1}$
is the probability of solubility over
$\mathbb{Q}_{p}$
for a cubic surface with valuations

with
$\min \{v_{p}(u),v_{p}(x)\}=0$
.
It remains to calculate
$\unicode[STIX]{x1D714}_{1}$
. If
$v_{p}(c_{1})=0$
(probability
$(p-1)/p$
) then there is a simple root over
$\mathbb{F}_{p}$
which can lifted. If
$v_{p}(c_{1})\geqslant 1$
and
$v_{p}(c_{0})=0$
(probability
$(p-1)/p^{2}$
) then the probability of solubility over
$\mathbb{Q}_{p}$
is precisely
$\unicode[STIX]{x1D71A}$
, in the notation of Lemma 2.4. Suppose now that
$v_{p}(c_{1}),v_{p}(c_{0})\geqslant 1$
(probability
$1/p^{2}$
). If
$v_{p}(d_{0})=0$
(probability
$(p-1)/p$
) then we must replace
$x$
by
$px$
in order to arrive at the valuations

Arguing as in the close of the proof of Lemma 2.4, the probability of solubility over
$\mathbb{Z}_{p}$
with
$u\in \mathbb{Z}_{p}^{\ast }$
is found to be
$\unicode[STIX]{x1D70E}_{1}+\unicode[STIX]{x1D70F}_{1}\unicode[STIX]{x1D71A}$
.
If
$v_{p}(d_{0})\geqslant 1$
(probability
$1/p$
), we simplify to get the valuations

We are interested in solubility over
$\mathbb{Q}_{p}$
with
$\min \{v_{p}(u),v_{p}(x)\}=0$
. With probability
$\unicode[STIX]{x1D70E}_{1}$
the first form has a simple
$\mathbb{F}_{p}$
-root which can be lifted. With probability
$1-\unicode[STIX]{x1D70E}_{1}-\unicode[STIX]{x1D70F}_{1}$
the first form is irreducible over
$\mathbb{F}_{p}$
and the overall probability of solubility is equal to
$\unicode[STIX]{x1D6FE}_{0}$
, in the notation of Lemma 2.6. Finally, with probability
$\unicode[STIX]{x1D70F}_{1}$
the first form has a triple root which can be moved to
$(1:0)$
. In this case the probability of solubility is equal to
$\unicode[STIX]{x1D716}=A+\unicode[STIX]{x1D6FC}_{1}(1)/p^{4}$
, in the notation of Lemma 2.7.
Recalling that
$\unicode[STIX]{x1D70F}_{1}=1/p^{2}$
, we have shown that
$\unicode[STIX]{x1D714}_{1}=C+\unicode[STIX]{x1D6FC}_{1}(1)/p^{9}$
, where

We now substitute this into our earlier expression and rearrange things to finally conclude that

2.7 Calculation of
$\unicode[STIX]{x1D6FC}_{1}(2)$
It remains to study the probability of solubility over
$\mathbb{Q}_{p}$
for cubic surfaces with valuations

We claim that

in the notation of Lemmas 2.4 and 2.7. To begin with, if
$v_{p}(c_{0})=1$
(with probability
$(p-1)/p$
) then we must have
$v_{p}(u),v_{p}(v),v_{p}(y)\geqslant 1$
in any solution. Making the obvious transformations and simplifying, the probability of solubility over
$\mathbb{Q}_{p}$
with
$x\in \mathbb{Z}_{p}^{\ast }$
is precisely
$\unicode[STIX]{x1D71A}$
. If, on the other hand,
$v_{p}(c_{0})\geqslant 2$
(probability
$1/p$
) then we replace
$v$
by
$pv$
to arrive at valuations for which the probability of solubility over
$\mathbb{Q}_{p}$
is
$\unicode[STIX]{x1D716}$
, in the notation of Lemma 2.7. The claim follows.
2.8 Conclusion
Returning to (2.1), we substitute the values of
$\unicode[STIX]{x1D70E}_{\ast },\unicode[STIX]{x1D70F}_{\ast }$
from part (2) of Lemma 2.2. This leads to the expression

where

Using a computer algebra package, it remains to substitute our various expressions for
$\unicode[STIX]{x1D6FC}_{1}(0),\unicode[STIX]{x1D6FC}_{1}(1),\unicode[STIX]{x1D6FC}_{1}(2)$
and
$\unicode[STIX]{x1D6FC}_{2}(1),\unicode[STIX]{x1D6FC}_{2}(2)$
into this, before simplifying the expression. This ultimately leads to the expressions for
$\unicode[STIX]{x1D712}_{p}(x)$
recorded in §2.2, with
$x=1/p$
.
3 Global solubility
We have seen that in order to prove Theorem 1.3 it suffices to prove that there is a positive proportion of irreducible
$f\in V(\mathbb{Z})$
, when ordered by height, such that
$C_{f}(\mathbb{Q})\neq \emptyset$
, where
$C_{f}$
is the associated genus-one curve (1.2). Our template for proving this is the argument developed by Bhargava [Reference Bhargava1, §4] to show that a positive proportion of plane cubic curves have a
$\mathbb{Q}$
-point.
3.1 Selmer groups of Mordell curves
Our work relies on the arithmetic of Mordell curves

for
$k\in \mathbb{Z}$
. These elliptic curves have
$j$
-invariant
$0$
and (potential) complex multiplication by the ring of integers of
$\mathbb{Q}(\sqrt{-3})$
. There is a
$3$
-isogeny
$\unicode[STIX]{x1D711}=\unicode[STIX]{x1D711}_{k}:E_{k}\rightarrow E_{-27k}$
and, by duality, a
$3$
-isogeny
$\widehat{\unicode[STIX]{x1D711}}=\unicode[STIX]{x1D711}_{-27k}:E_{-27k}\rightarrow E_{k}$
. As
$k$
varies over the integers, Bhargava et al [Reference Bhargava, Elkies and Shnidman3] have undertaken an extensive investigation into the average size of the
$\unicode[STIX]{x1D711}$
-Selmer groups
$\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{k})\subset \text{H}^{1}(G_{\mathbb{Q}},E_{k}[\unicode[STIX]{x1D711}])$
and
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})\subset \text{H}^{1}(G_{\mathbb{Q}},E_{-27k}[\widehat{\unicode[STIX]{x1D711}}])$
, where
$G_{\mathbb{Q}}=\text{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})$
, consisting of “locally soluble” cohomology classes. Let

While it does not play a role in our work, when the elliptic curves
$E_{k}$
are ordered by height, it follows from [Reference Bhargava, Elkies and Shnidman3, Theorem 1] that the average of
$\#\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{k})$
is
$1+\unicode[STIX]{x1D71A}$
(respectively
$1+\unicode[STIX]{x1D71A}/3$
) if
$k<0$
(respectively
$k>0$
).
We shall be interested in soluble elements of
$\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{k})$
and
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
According to the fundamental short exact sequence

the soluble elements of
$\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{k})$
are precisely those that come from the group
$E_{-27k}(\mathbb{Q})/\unicode[STIX]{x1D711}(E_{k}(\mathbb{Q}))$
. Similarly, soluble elements of
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
are precisely those that come from
$E_{k}(\mathbb{Q})/\widehat{\unicode[STIX]{x1D711}}(E_{-27}(\mathbb{Q}))$
. Noting that
$\widehat{\unicode[STIX]{x1D711}}\circ \unicode[STIX]{x1D711}=[3]$
, we may now turn to the exact sequence

which is recorded in [Reference Silverman19, Remark X.4.7]. The groups appearing in this sequence are all
$\mathbb{F}_{3}$
-vector spaces and it follows that

(Since the two groups on the left-hand side sit inside the relevant Selmer groups, this automatically implies the inequality recorded in [Reference Bhargava, Elkies and Shnidman3, Proposition 42(i)].)
3.2 Binary cubic forms
For any Dedekind domain
$D$
, we write
$V(D)$
for the set of binary cubic forms
$f(u,v)=[a,3b,3c,d]$
, with
$a,b,c,d\in D$
. We put

for the (reduced) discriminant of
$f$
. We put
$V(D)_{A}$
for the set of
$f\in V(D)$
such that
$\operatorname{disc}(f)=A$
.
Taking
$D=\mathbb{Z}$
, we shall adopt the notation in [Reference Bhargava, Elkies and Shnidman3] and introduce the sets

where
$C_{f}$
is given by (1.2). Put
$V(\mathbb{Z})_{A}^{\text{sol}}$
and
$V(\mathbb{Z})_{A}^{\text{loc}\,\text{sol}}$
for the corresponding subsets restricted to have discriminant
$A$
. The group
$\text{GL}_{2}(\mathbb{Z})$
acts naturally on
$V(\mathbb{R})$
by linear change of variables and the discriminant is
$\text{SL}_{2}(\mathbb{Z})$
-invariant. We let
${\mathcal{F}}$
be a fundamental domain for the action of
$\text{SL}_{2}(\mathbb{Z})$
on
$V(\mathbb{R}).$
We set

This set has finite volume. The following result (cf. [Reference Bhargava, Elkies and Shnidman3, Theorem 37]) is due to Davenport [Reference Davenport8] and Davenport and Heilbronn [Reference Davenport and Heilbronn9].
Lemma 3.1 (Davenport–Heilbronn).
We have

as
$X\rightarrow \infty$
, where
$c_{1}=\operatorname{meas}({\mathcal{F}}_{1})=\unicode[STIX]{x1D70B}^{2}/3$
.
The following result allows us to associate a
$\unicode[STIX]{x1D711}$
-Selmer element of an elliptic curve
$E_{k}$
over
$\mathbb{Q}$
, with
$k\in \mathbb{Z}$
, to an element of
$V(\mathbb{Z})$
.
Lemma 3.2. Let
$k\in 3\mathbb{Z}$
. There exists an injective map from the set of non-identity elements of
$\,\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
to the set of irreducible
$f\in {\mathcal{F}}\cap V(\mathbb{Z})_{4k}^{\text{loc}\,\text{sol}}$
.
Proof. Assume throughout the proof that
$k\in 3\mathbb{Z}$
. According to [Reference Bhargava, Elkies and Shnidman3, Theorem 33] there is a bijection between the
$\text{SL}_{2}(\mathbb{Q})$
-equivalence classes of
$V(\mathbb{Q})_{4k}^{\text{loc}\,\text{sol}}$
and elements of
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
. Under this bijection, the identity element of
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
corresponds to the unique
$\text{SL}_{2}(\mathbb{Q})$
-equivalence class of reducible forms in the space
$V(\mathbb{Q})_{4k}$
(i.e. the orbit of
$f(u,v)=ku^{3}+3uv^{2}$
under
$\text{SL}_{2}(\mathbb{Q})$
). Next, it follows from [Reference Bhargava, Elkies and Shnidman3, Theorem 34] that any
$\text{SL}_{2}(\mathbb{Q})$
-equivalence class of
$V(\mathbb{Q})_{4k}^{\text{loc}\,\text{sol}}$
contains an element of
$V(\mathbb{Z})_{4k}$
. This therefore produces the required map from non-identity elements of
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{-27k})$
to
$\text{SL}_{2}(\mathbb{Z})$
-equivalence classes of irreducible elements of
$V(\mathbb{Z})_{4k}^{\text{loc}\,\text{sol}}$
. Injectivity follows from injectivity of the map in [Reference Bhargava, Elkies and Shnidman3, Theorem 33].◻
3.3 A positive proportion of curves with positive rank
We now turn to the work of Kriz and Li [Reference Kriz and Li15]. Let
${\mathcal{D}}\subset \mathbb{Z}$
be the subset of fundamental discriminants. Then
$d\in {\mathcal{D}}$
if and only if either
$d\equiv 1\hspace{0.2em}{\rm mod}\hspace{0.2em}4$
is square-free, or
$d=4d^{\prime }$
for some square-free integer
$d^{\prime }\equiv \{2,3\}\hspace{0.2em}{\rm mod}\hspace{0.2em}4$
. We have the following result.
Lemma 3.3. We have

as
$X\rightarrow \infty$
.
Proof. It follows from [Reference Kriz and Li15, Theorem 1.8] that the analytic rank of
$E_{-432k}$
is one for at least
${\textstyle \frac{1}{6}}$
of fundamental discriminants
$k\in {\mathcal{D}}$
. By appealing to the celebrated work of Gross and Zagier [Reference Gross and Zagier10], and Kolyvagin [Reference Kolyvagin14], we deduce that the same is true for the arithmetic rank. Finally, the statement of the lemma follows on recalling that
$\#{\mathcal{D}}\cap (0,X]=(3/\unicode[STIX]{x1D70B}^{2})X+o(X)$
, as
$X\rightarrow \infty$
.◻
3.4 Proof of the main result
Since
$432k\in 3\mathbb{Z}$
, we begin by deducing from Lemma 3.2 that there is an injective map from soluble non-identity elements of
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{27\cdot 432k})$
to the set of irreducible elements in
${\mathcal{F}}\cap V(\mathbb{Z})_{-4\cdot 432k}^{\text{sol}}$
. Recalling that
$\widehat{\unicode[STIX]{x1D711}}_{-432k}=\unicode[STIX]{x1D711}_{27\cdot 432k}$
, we may make the change of variables
$27k\rightarrow -k$
, in order to deduce that there is an injective map from soluble non-identity elements of
$\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{-432k})$
to the set of irreducible elements in
${\mathcal{F}}\cap V(\mathbb{Z})_{64k}^{\text{sol}}$
. In view of (3.1), if
$k\in \mathbb{Z}$
is such that
$\operatorname{rank}E_{-432k}(\mathbb{Q})=1$
, then either
$E_{27\cdot 432k}(\mathbb{Q})/\unicode[STIX]{x1D711}(E_{-432k}(\mathbb{Q}))$
or
$E_{-432k}(\mathbb{Q})/\widehat{\unicode[STIX]{x1D711}}(E_{27\cdot 432k}(\mathbb{Q}))$
must have size at least 3. But then it follows that there are at least two soluble non-identity elements belonging to
$\operatorname{Sel}_{\unicode[STIX]{x1D711}}(E_{-432k})$
or
$\operatorname{Sel}_{\widehat{\unicode[STIX]{x1D711}}}(E_{27\cdot 432k})$
. This implies that

by Lemma 3.3. We henceforth set
$c_{2}=1/(2^{6}\cdot 3^{3}\cdot \unicode[STIX]{x1D70B}^{2})$
for the constant appearing in this bound.
We need a version of this which runs over irreducible elements of
$V(\mathbb{Z})^{\text{sol}}$
which are ordered by height, for which we ape an argument of Bhargava [Reference Bhargava1, §4]. Let
$B=[-1,1]^{4}$
. We require a lower bound for the counting function

as
$X\rightarrow \infty$
. Since
${\mathcal{F}}_{X}$
has finite volume, it follows that for any given
$\unicode[STIX]{x1D700}>0$
, there exists
$r_{\unicode[STIX]{x1D700}}>0$
such that

for all
$X>1$
. Thus there exist constants
$\unicode[STIX]{x1D6FF}>0$
and
$r>0$
such that

for any
$X>1$
, where we recall from Lemma 3.1 that
$c_{1}=\operatorname{meas}({\mathcal{F}}_{1})=\unicode[STIX]{x1D70B}^{2}/3$
. Now it easily follows from Lemma 3.1 and Hilbert’s irreducibility theorem that

The number of binary cubic forms contributing to the left-hand side which do not belong
$V(\mathbb{Z})^{\text{sol}}$
is

by Lemma 3.1 and (3.2). Putting
$Y=r^{-4}X$
, it therefore follows that

This completes the proof of Theorem 1.3.
Acknowledgements
The author is very grateful to Tim Dokchitser, Adam Morgan, Jack Thorne and Olivier Wittenberg for useful conversations, and to the anonymous referee for several helpful comments. Thanks are also due to Jörg Jahnel for help with Remark 2.1. While working on this paper the author was supported by ERC grant 306457.