[columns=2, title=Index, intoc]
Brauer–Manin obstruction for integral points on Markoff-type cubic surfaces
Abstract
Following [GS22], [LM20] and [CWX20], we study the Brauer–Manin obstruction for integral points on similar Markoff-type cubic surfaces. In particular, we construct a family of counterexamples to strong approximation which can be explained by the Brauer–Manin obstruction with some counting results of similar nature to those in [LM20] and [CWX20]. We also give some counterexamples to the integral Hasse principle which cannot be explained by the (algebraic) Brauer–Manin obstruction.
1 Introduction
Let be an affine variety over , and an integral model of over , i.e. an affine scheme of finite type over whose generic fiber is isomorphic to . Define the set of adelic points , where is a prime number or (with ). Similarly, define (with ). We say that fails the Hasse principle if
We say that fails the integral Hasse principle if
We say that satisfies weak approximation if the image of in is dense, where the product is taken over all places of . Finally, we say that satisfies strong approximation if is dense in , where denotes the set of connected components of . Note that we work with since is never dense in for topological reasons (see [Con12, Example 2.2]).
In general, the Hasse principle for varieties does not hold. In his 1970 ICM address [Man71], Manin introduced a natural cohomological obstruction to the Hasse principle, namely the Brauer–Manin obstruction (which has been extended to its integral version in [CX09]). If denotes the cohomological Brauer group of , i.e. , we have a natural pairing from class field theory:
If we define to be the left kernel of this pairing, then the exact sequence of Albert–Brauer–Hasse–Noether gives us the relation:
Similarly, by defining the Brauer–Manin set , we also have that
This gives the so-called integral Brauer–Manin obstruction. We say that the Brauer–Manin obstruction to the (resp. integral) Hasse principle is the only one if
If there is no confusion, we can omit the symbol for the set of local integral points and the corresponding Brauer–Manin set. \̃\ We are particularly interested in the case where is a hypersurface, defined by a polynomial equation of degree in an affine space. The case is easy. The case considers the arithmetic of quadratic forms: for rational points, the Hasse principle is always satisfied by the Hasse–Minkowski theorem, and for integral points, the Brauer–Manin obstruction to the integral Hasse principle is the only one (up to an isotropy assumption) due to work of Colliot-Thélène, Xu [CX09] and Harari [Har08]. However, the case (of cubic hypersurfaces) is still largely open, especially for integral points. Overall, the arithmetic of integral points on the affine cubic surfaces over number fields is still little understood. For example, the question to determine which integers can be written as sums of three cubes of integers is still open. In this first problem, for the affine variety defined by the equation
where is a fixed integer, Colliot-Thélène and Wittenberg in [CW12] proved that there is no Brauer–Manin obstruction to the integral Hasse principle (if is not of the form ). However, the existence of such an integer remains unknown in general, with the first example now being the case when . On the other hand, in a related problem, there is no Brauer–Manin obstruction to the existence of an integral point on the cubic surface defined by
for any , also proven in [CW12].
Another interesting example of affine cubic surfaces that we consider is given by Markoff surfaces which are defined by
where is an integer parameter. The very first (original) class of Markoff surfaces which was studied is the one given by this equation with in a series of papers [BGS16], [BGS16a], and recently [Che21]. They study strong approximation mod for for any prime and present a Strong Approximation Conjecture ([BGS16a, Conjecture 1]):
Conjecture 1.1.
For any prime , consists of two orbits, namely and . Here is a group of affine integral morphisms of generated by the permutations of the coordinates and the Vieta involutions.
Combining the above three papers by Bourgain, Gamburd, Sarnak, and Chen, the Conjecture is established for all but finitely many primes (see [Che21, Theorem 5.5.5]), which also implies that satisfies strong approximation mod for all by finitely many primes.
On the other hand, in [GS22], Ghosh and Sarnak study the integral points on those affine Markoff surfaces with general , both from a theoretical point of view and by numerical evidence. They prove that for almost all , the integral Hasse principle holds, and that there are infinitely many ’s for which it fails (Hasse failures). Furthermore, their numerical experiments suggest particularly a proportion of integers satisfying of the power for which the integral Hasse principle is not satisfied.
Subsequently, Loughran and Mitankin [LM20] proved that asymptotically only a proportion of of integers such that presents an integral Brauer–Manin obstruction to the Hasse principle. They also obtained a lower bound, asymptotically , for the number of Hasse failures which cannot be explained by the Brauer–Manin obstruction. After Colliot-Thélène, Wei, and Xu [CWX20] obtained a slightly stronger lower bound than the one given in [LM20], no better result than their number has been known until now. In other words, with all the current results, one does not have a satisfying comparison between the numbers of Hasse failures which can be explained by the Brauer–Manin obstruction and which cannot be explained by this obstruction. Meanwhile, for strong approximation, it has been proven to almost never hold for Markoff surfaces in [LM20] and then absolutely never be the case in [CWX20]. Here we recall an important conjecture given by Ghosh and Sarnak.
Conjecture 1.2.
The above conjecture also means that almost all counterexamples to the integral Hasse principle for Markoff surfaces cannot be explained by the Brauer–Manin obstruction, thanks to the result obtained by [LM20]. While the question of counting all counterexamples to the integral Hasse principle for Markoff surfaces remains largely open, we will focus on another family in this paper. More precisely, we are going to study the set of integral points of a different Markoff-type cubic surfaces whose origin is similar to that of the original Markoff surfaces , namely the relative character varieties which will be introduced in Section 2, using the Brauer–Manin obstruction as well. The surfaces are given by the cubic equation:
where are parameters that satisfy some specific relations to be discussed later. Due to the similar appearance as that of the original Markoff surfaces, one may expect to find some similarities in their arithmetic as well. One of the main results in our paper is the following, saying that a positive proportion of these relative character varieties have no (algebraic) Brauer–Manin obstruction to the integral Hasse principle as well as fail strong approximation, and those failures can be explained by the Brauer–Manin obstruction.
Theorem 1.3.
Let be the affine scheme over defined by
where
such that the projective closure of is smooth. Then we have
as .
The structure of the paper is as follows. In Section 2, we provide some background on character varieties and a natural origin of the Markoff-type cubic surfaces. In Section 3, we first review some important results on affine cubic surfaces given by the complement of three coplanar lines and their Brauer groups. After the general setting, we turn our attention to the natural smooth projective compactifications of the Markoff-type cubic surfaces, where we explicitly calculate the (algebraic) Brauer group of the compactification, and then we complete the analysis of the Brauer group by calculating the Brauer group of the affine surfaces. In Section 4, we use the Brauer group to give explicit examples of Brauer–Manin obstructions to the integral Hasse principle, and give some counting results for the frequency of the obstructions. Finally, in Section 5, we make some important remarks to compare some results in this paper to those of Markoff surfaces in recent work that we follow, and We also give some counterexamples to the integral Hasse principle which cannot be explained by the Brauer–Manin obstruction.
Notation. Let be a field and a separable closure of . We let be the absolute Galois group. A -variety is a separated -scheme of finite type. If is a -variety, we write . Let and . If is an integral -variety, let denote the function field of . If is a geometrically integral -variety, let denote the function field of .
Let denote the Picard group of a scheme . Let denote the Brauer group of . Let
denote the algebraic Brauer group of a -variety and let denote the image of . The image of is called the transcendental Brauer group of .
Given a field of characteristic zero containing a primitive -th root of unity , we have . The choice of then defines an isomorphism . Given two elements , we have their classes and in . We denote by the class corresponding to the cup-product . Suppose is a finite Galois extension with Galois group . Given and , we have . In particular, if , then . For all the details, see [GS17, Sections 4.6, 4.7].
Let be a discrete valuation ring with fraction field and residue field . Let denote the valuation . Let be an integer invertible in . Assume that contains a primitive -th root of unity . For , we have the residue map
where is induced by the isomorphism sending to . This map sends the class of to
For a proof of these facts, see [GS17]. Here we recall some precise references. Residues in Galois cohomology with finite coefficients are defined in [GS17, Construction 6.8.5]. Comparison of residues in Milnor K-Theory and Galois cohomology is given in [GS17, Proposition 7.5.1]. The explicit formula for the residue in Milnor’s group K2 of a discretely valued field is given in [GS17, Example 7.1.5].
Acknowledgements. I thank Cyril Demarche for his help and supervision during my PhD study at the Institute of Mathematics of Jussieu. I thank Kevin Destagnol for his help with the computations in Section 4.3 using analytic number theory. I thank Vladimir Mitankin for his useful remarks and suggestions, especially regarding Section 5. I thank Jean-Louis Colliot-Thélène, Fei Xu, and Daniel Loughran for their helpful comments and encouragement. This project has received funding from the European Union’s Horizon 2020 Research and Innovation Programme under the Marie Skłodowska-Curie grant agreement No. 754362 and from the Vietnam Academy of Science and Technology’s 2022 Support Programme for Junior Researchers. I thank the reviewers for their careful reading of my paper and their many insightful comments and suggestions which helped me improve considerably the manuscript.
2 Background
The main reference to look up notations that we use here is [Wha20], mostly Chapter 2.
2.1 Character varieties
First, we introduce an important origin of the Markoff-type cubic surfaces which comes from character varieties, as studied in [Wha20]. Throughout this section, an algebraic variety is a scheme of finite type over a field. Given an affine variety over a field , we denote by its coordinate ring over . If moreover is integral, then denotes its function field over . Given a commutative ring with unity, the elements of will be referred to as regular functions on the affine scheme .
Definition 2.1.
Let be a finitely generated group. Its () representation variety is the affine scheme defined by the functor
for every commutative ring . Assume that has a sequence of generators of elements, then we have a presentation of as a closed subscheme of defined by equations coming from relations among the generators. For each , let be the regular function on given by .
The () character variety of over is then defined to be the affine invariant theoretic quotient
under the conjugation action of .
The regular function for each clearly descends to a regular function on . Furthermore, from the fact that and , for being the identity matrix and for any , we can deduce a natural model of over as the spectrum of
Given any integral domain with fraction field of characteristic zero, the -points of parametrize the Jordan equivalence classes of -representations of having character valued in .
Example 2.2.
Denote by the free group on generators . By Goldman’s results used in [Wha20], we have the following important examples:
-
(1)
.
-
(2)
.
-
(3)
The coordinate ring is the quotient of the polynomial ring
by the ideal generated by two elements
and
Now given a connected smooth compact manifold , we consider the moduli of local systems on which is the character variety of its fundamental group. More generally, given a smooth manifold with finitely many connected components for , define
The construction of the moduli space is functorial in the manifold . More precisely, any smooth map of manifolds induces a morphism , depending only on the homotopy class of , given by pullback of local systems.
Let be a surface. For each curve , there is a well-defined regular function , which agrees with for any freely homotopic to a parametrization of . The boundary curves of induce a natural morphism
Now since we can write , we have an identification
given by taking a local system on the disjoint union of circles to its sequence of traces along the curves. The morphism above may be viewed as an assignment to each its sequence of traces . The fibers of for will be denoted . Each is often called a relative character variety in the literature. If is a surface of type satisfying , then the relative character variety is an irreducible algebraic variety of dimension .
Given a fixed surface , a subset , and a subset , we shall denote by
the set of all such that and for every essential curve . By [Wha20, Lemma 2.5], there is no risk of ambiguity with this notation, i.e., has a model over and recovers the set of -valued points of in the sense of algebraic geometry.
2.2 Markoff-type cubic surfaces
Now we give a description of the moduli spaces for and . These cases are special since each is an affine cubic algebraic surface with an explicit equation.
-
(1)
Let be a surface of type , i.e. a one holed torus. Let be an optimal sequence of generators for , as given in [Wha20, Definition 2.1]. By Example 2.2, we have an identification . From the trace relations in , we obtain that
Writing so that each of the variables , , and corresponds to an essential curve on as depicted in [Wha20, Figure ], the moduli space has an explicit presentation as an affine cubic algebraic surface in with the equation
These are exactly the Markoff surfaces as studied in the series of papers [GS22], [LM20], and [CWX20] with .
-
(2)
Let be a surface of type , i.e. a four holed sphere. Let be an optimal sequence of generators for . Set
so that each of the variables corresponds to an essential curve on . By Example 2.2, for the relative character variety is an affine cubic algebraic surface in given by the equation
(1) where
(2) These are the Markoff-type cubic surfaces that we are going to study in this paper.
3 The Brauer group of Markoff-type cubic surfaces
Our main interest is in the second Markoff-type cubic surfaces defined by . We are now going to give some explicit computations on the Brauer group of these surfaces. First of all, let us recall some basic definitions and results on the Brauer group of varieties over a field.
Let be an arbitrary field. Recall that for a variety over there is a natural filtration on the Brauer group
which is defined as follows.
Definition 3.1.
Let
The subgroup is the algebraic Brauer group of and the quotient is the transcendental Brauer group of .
From the Hochschild–Serre spectral sequence, we have the following spectral sequence:
which is contravariantly functorial in the -variety . It gives rise to the functorial exact sequence of terms of low degree:
Let be a variety over a field such that . By Hilbert’s Theorem 90 we have , then by the above sequence there is an exact sequence
This sequence is also contravariantly functorial in .
Remark 3.2.
Let be a variety over a field such that . This assumption holds for any proper, geometrically connected and geometrically reduced -variety .
-
(1)
If has a -point, which defined a section of the structure morphism , then each of the maps and has a retraction, hence is injective. (Then is an isomorphism.) Therefore, we have an isomorphism
-
(2)
If is a number field, then (see [CF67], Chapter VII, Section 11.4, p. 199). Thus for a variety over a number field such that , we have an isomorphism
3.1 Geometry of affine cubic surfaces
In this section, we study the geometry of affine cubic surfaces with special regards to the Brauer group. By an affine cubic surface, we mean an affine surface of the form
where is a polynomial of degree of 3. The closure of in is a cubic surface . The complement is a hyperplane section on . Much of the geometry of can be understood in terms of the geometry of and , especially in the case of Markoff-type cubic surfaces. There has been already much work on the Brauer groups of affine cubic surfaces when the hyperplane section is smooth, for example see [CW12]. Here we shall be interested in the case where the hyperplane section is singular; in particular, we focus on the case where is given by 3 coplanar lines. All results here are proven in either [CWX20] or [LM20].
We begin with an important result for cubic surfaces over an algebraically closed field.
Proposition 3.1.
[CWX20, Proposition 2.2] Let be a smooth projective cubic surface over a field of characteristic zero. Suppose a plane cuts out on three distinct lines over . Let be the complement of this plane. Then the natural map is an isomorphism of Galois modules and the natural sequence
is an exact sequence of Galois lattices.
As is torsion free, we have the following result for the algebraic Brauer group, using the computation by Magma.
Proposition 3.2.
[LM20, Proposition 2.5] Let be a smooth projective cubic surface over a field of characteristic 0. Let be a hyperplane section which is the union of 3 distinct lines and let . Then is torsion free and is isomorphic to one of the following groups:
For the transcendental Brauer group, from the discussion on page 140 of [CS21], note that so we have .
Proposition 3.3.
[CWX20, Proposition 2.1] [LM20, Proposition 2.4] Let be a smooth projective cubic surface over a field of characteristic 0. Suppose that is an open subset of such that is the union of three distinct -lines, by which we mean a smooth projective curve isomorphic to . Suppose any two lines intersect each another transversely in one point, and that the three intersection points are distinct. Let be one of the three lines and be the complement of the 2 intersection points of with the other two lines. Then the residue map
induces a -isomorphism
In particular, if contains no non-trivial roots of unity then
Lemma 3.4.
[CWX20, Lemma 2.4] Let be a field of characteristic 0. Let . Then is (noncanonically) isomorphic to , the group of roots of unity in .
We end this section by the following result which applies to number fields and more generally to function fields of varieties over number fields.
Corollary 3.5.
[CWX20, Corollary 2.3] Let be a field of characteristic 0 such that in any finite field extension there are only finitely many roots of unity. Let be a smooth projective cubic surface over . Suppose that a plane cuts out on three nonconcurrent lines. Let be the complement of the plane section. Then the quotient is finite.
3.2 The geometric Picard group and algebraic Brauer group
Using the equations, we can compute explicitly the algebraic Brauer group of the Markoff-type cubic surfaces in question. First, we have the following important result.
Lemma 3.6.
Let be a number field and let be a cubic surface defined by the equation
where are defined by for some . Then is singular if and only if we are in one of the following cases:
-
where and
-
at least one of the parameters equals .
If satisfies none of those two conditions and where
then the 27 lines on the smooth cubic surface are defined over by the following equations
and
-
1.
-
2.
-
3.
-
4.
-
5.
-
6.
with and . Furthermore, we have the intersection numbers
for any pair , for all .
Proof.
The necessary and sufficient condition for the affine open surface to be singular is proven in [CL09, Theorem 3.7]. It is easy to verify that there is no singular point at infinity on the projective surface .
Now without loss of generality, we consider the system of equations
and put them in the original equation of the cubic surfaces to solve for . We can work similarly for and to find all the given equations of the 27 lines. ∎
Now given the data of the lines, we can compute directly the algebraic Brauer group of the Markoff-type cubic surfaces in question.
Proposition 3.7.
Let be a number field. Let be a cubic surface defined by the equation
(3) |
where are defined by for some . Assume that is smooth over and , then
Proof.
Since is geometrically rational, one has . By taking for instance, one clearly has , so . Since is a number field, by the Hochschild–Serre spectral sequence, we have an isomorphism
By the above lemma, we can easily verify that the six lines , , , , , and on the cubic surface are skew to each other, hence they may be simultaneously blown down to by [Har77], Chapter V, Proposition 4.10. For the sake of simplicity, here we shall write these six lines respectively as for . The class of the canonical divisor on is equal to , where is the inverse image of the class of lines in . By [Har77, Chapter V, Proposition 4.8], the classes form a basis of , and we have the following intersection properties: for .
Now we consider the action of the Galois group on . One clearly has , where
We have the following intersection numbers, noting that , , , and :
(5) |
(6) |
(7) |
and
(8) |
(9) |
(10) |
Hence, we obtain
(11) |
in by [Har77, Chapter V, Proposition 4.9]. As a result, we have
(12) |
and clearly . Then
(13) |
(14) |
and
(15) |
Given a finite cyclic group and a -module , by [NSW15, Proposition 1.7.1], recall that we have isomorphisms , where the latter group is the quotient of , the set of elements of of norm , by its subgroup .
By [NSW15, Proposition 1.6.7], we have
where . Then one has the following (inflation-restriction) exact sequence
hence . Continuing as above, we have
hence . Now we are left with
hence . Indeed, the last group can be computed as follows. We have
Considering the action of on this invariant group, we have
On the other hand,
Given these results, we conclude that
∎
Theorem 3.8.
Let be a number field. With the same notations as before, let such that where and is smooth over . Let be the affine cubic surface defined by the equation
where are defined by for some . Then we have
with a generator
where for . Furthermore, we also have
with a generator
over .
Proof.
We keep the notation as in the previous proposition. Then is given by the following quotient group
by Proposition 3.1 and formula . Here for any divisor , denote by its image in . By Proposition 3.1, we also have . By the Hochschild–Serre spectral sequence, we have the following injective homomorphism
and in fact it is an isomorphism because over a number field , we have from class field theory. Furthermore, the smooth compactification of has rational points, hence so does , which comes from the fact that any smooth cubic surface over an infinite field is unirational over as soon as it has a -point (see [Kol02]), so we also have .
Since is free and acts on trivially, we obtain that by [NSW15, Proposition 1.6.7]. Now in , as , and , we have the following equalities
(16) |
Using the inflation-restriction sequences similarly as in the previous proposition, with replaced by , we can compute that
Now we produce concrete generators in for . Then is defined by the equation
We will show that the following quaternion algebras in are non-constant elements of , and hence they are equal in .
(17) |
Indeed, it suffices to prove the claim for and we only need to show that , since its formula implies that becomes zero under the field extension , i.e., it is algebraic. By Grothendieck’s purity theorem ([Poo17, Theorem 6.8.3]), for any smooth integral variety over a field of characteristic , we have the exact sequence
where the last map is given by the residue along the codimension-one point . Therefore, to prove that our algebras come from a class in , it suffices to show that all their residues are trivial. We will show that
so that its corestriction is a well-defined element over . From the data of the 27 lines in Lemma 3.6 and the formula of , any non-trivial residue of must occur along an irreducible component of the following divisor(s)
However, clearly in the function field of any such irreducible component, is a square; standard formulae for residues in terms of the tame symbol [GS17, Example 7.1.5, Proposition 7.5.1] therefore show that is unramified, and hence . The residues of at the lines which form the complement of in are easily seen to be trivial. One thus also has . This element is non-constant by the Faddeev exact sequence ([CS21, Theorem 1.5.2]), since the residue of regarded as an element of is nontrivial at the closed point of . Alternatively, by [CS21, Corollary 11.3.5] on the Brauer group of conic bundles, the element is indeed non-constant. Furthermore, it will contribute to the Brauer–Manin obstruction to strong approximation in the next section.
Finally, the fact that viewed as an element of (by abuse of notation) and that is the open subset of defined by give the desired generator of . ∎
3.3 The transcendental Brauer group
We begin with a specific assumption.
Assumption 3.3.
Let and such that . For all , assume that such that is not a square in .
Now we compute the transcendental Brauer group in our particular case.
Proposition 3.9.
Let satisfy Assumption 3.3. Let be the affine cubic surface over defined by
where are defined by . Assume that its natural compactification is smooth over . Set and . Then the natural map is an isomorphism. Moreover, has trivial transcendental Brauer group over .
Proof.
The proof is inspired by that of [LM20, Proposition 4.1]. Let be a non-constant Brauer element. Multiplying by a constant algebra if necessary, by Proposition 3.3 and Theorem 3.8, we may assume that has order dividing (note that under our assumption of , the field extension is totally real and thus contains no nontrivial roots of unity). In order to show that , we only need to show that is unramified along the three lines on by Grothendieck’s purity theorem (see [CS21, Theorem 3.7.2]).
Let and . Let . Note that meets at two rational points, so is non-canonically isomorphic to . Let the point be the identity element of the group law. Then an isomorphism with is realized via the following homomorphism:
(18) |
The residue of along lies inside . Assume by contradiction that the residue is nontrivial. Since the order of is a power of dividing , then we can assume that the residue has order (up to replacing by ). This means that the residue corresponds to some irreducible degree finite étale cover .
Over the field , the conic fiber over the coordinate is split, i.e. a union of two lines over . These lines meet at the points
Let be the irreducible component of containing , i.e., . Consider the restriction of to . This is well-defined outside of , and since has constant Brauer group, actually extends to all of . As meets transversely, by the functoriality of residues ([CS21, Section 3.7]) we deduce that the residue of at is also trivial, so the fiber consists of exactly two rational points. This implies that is geometrically irreducible, hence non-canonically.
Now by choosing a rational point over and using the above group homomorphism, we may therefore identify the degree cover with the map
(19) |
However, our assumptions on imply that is not a square in , which gives a contradiction. Thus the residue of along is trivial, and the same holds for the other lines. We conclude that is everywhere unramified, hence .
Now let be a non-constant element. Then over the field extension , the corresponding image of comes from by the above argument. As , it is clear that is algebraic. The result follows.
∎
Remark 3.4.
Note that by Proposition 3.3. However, in the above proposition, the Galois invariant element of order does not descend to a Brauer group element over , which is also the case in [LM20, Proposition 4.1].
4 The Brauer–Manin obstruction
4.1 Review of the Brauer–Manin obstruction
Here we briefly recall how the Brauer–Manin obstruction works in our setting, following [Poo17, Section 8.2] and [CX09, Section 1]. For each place of there is a pairing
coming from the local invariant map
from local class field theory (this is an isomorphism if is a prime number). This pairing is locally constant on the left by [Poo17, Proposition 8.2.9]. For integral points, any element pairs trivially on for almost all primes , so we obtain a pairing . As the local pairings are locally constant, we obtain a well-defined pairing
For , let be the left kernel with respect to , and let . By abuse of notation, from now on we write the reduced Brauer–Manin set in the standard way as . Note that the set depends only on the image of in the quotient . By Theorem 3.8, the map is an isomorphism, hence . We have the inclusions , so that can obstruct the integral Hasse principle or strong approximation on .
Let be dense Zariski open in . As is smooth, the set is dense in for all places . Moreover, is open in , hence is dense in . As the local pairings are locally constant, we may restrict our attention to to calculate the local invariants of a given element in . In particular, here we take the open subset given by .
4.2 Brauer–Manin obstruction from a quaternion algebra
Now we consider the Markoff-type cubic surfaces over and their integral models over defined by the equation :
where are defined as in with . Set . First of all, we study the existence of local integral points on those affine cubic surfaces given by .
Proposition 4.1 (Assumption A).
If satisfies mod , mod and mod , mod for , such that , for , and , then .
Proof.
With our specific choice of in the assumption, we obtain:
-
(1)
Prime powers of : The only solution modulo is the singular (up to permutation). However, we find a solution modulo with mod , so twice the order valuation at of the partial derivative at is less than the order valuation at of . This solution then lifts to a -adic integer solution by Hensel’s lemma (fixing the variables ).
-
(2)
Prime powers of : We find the non-singular solution , which lifts to a -adic integer solution by Hensel’s lemma (fixing the variables ).
-
(3)
Prime powers of : We would like to find a non-singular solution modulo of the equations which does not satisfy simultaneously
For simplicity, we will find a sufficient condition for the existence of a non-singular solution whose at least one coordinate is zero. First, it is clear that the equation always has a solution whose one coordinate is : indeed, take , then is equivalent to , and every element in can be expressed as a sum of two squares. Now assume that all such points in with one coordinate equal to are singular. Then modulo , if the required equations become , plus all the permutations for and . For convenience, we drop the phrase “modulo ”. From these equations, we get
(20) plus all the permutations, respectively
(21) and
(22) We will choose which does not satisfy all of these equations simultaneously to get a contradiction to our assumption at the beginning.
Now if satisfies all the above equations, we first require that for any . Next, without loss of generality (WLOG), assume in that with , then putting the formulas of in gives us , hence either or . (We also have similar equations for all the other cases.)If , then , and so from we obtain for any , hence for all . Otherwise, if but , WLOG assume in that , then and so from we have the following possibilities: , or . Therefore, we immediately deduce a sufficient condition for the nonexistence of singular solutions modulo : , for any , and .
As a result, assuming the hypothesis of the proposition, it is clear that has a smooth point, which then lifts to a -adic integral point by Hensel’s lemma (with respect to the variable at which the partial derivative is nonzero modulo , fixing other variables).
∎
We keep Assumption A to ensure that and study the Brauer–Manin obstruction to the existence of integral points. Note that our specific choice of implies that . Indeed, it is clear that with our assumption, is not a square in for . Then if and only if there are some such that is a square in ; however, since and does not have any solution modulo with odd, that cannot happen.
Now let us calculate the local invariants of the following quaternion algebras as elements of the algebraic Brauer group :
where for . Now for each , we have the local invariant map
Lemma 4.2 (Assumption B).
Let and be a prime such that and mod . Then there exist such that is a quadratic nonresidue mod , mod , mod , and satisfy the cubic equation defining an affine Markoff surface :
Furthermore, for any satisfying all the above conditions (Assumption B) and Assumption A, the local invariant map of the quaternion algebra at is surjective.
Proof.
For convenience (and by abuse of notation), we will also write (resp. ) as (resp. ). We denote by and write it simply as . Set .
First of all, over the affine cubic equation of can be rewritten equivalently as
(23) | ||||
for all satisfying our hypothesis. Since , ramifies over , i.e., where is a nonzero prime ideal of . Therefore, we have and
Following the proof of [LM20, Proposition 5.5], for all , there exists an -point on the variety
which satisfy . As , we know from [LM20, Lemma 5.3] that this gives rise to a smooth -point of , hence a -point with the same residue modulo by Hensel’s lemma. Now to construct the given -point, we restrict our attention to the subvariety given by mod then assume that mod and is a quadratic nonresidue mod . The above equations then become
Factoring the left hand side, it suffices to solve the equations
This then gives the equation of an affine curve
By the argument in the proof of [LM20, Proposition 5.5], this affine curve has a unique singular point and has many -points. Of these points at most satisfy , and gives only at most points, hence providing , there exists an -point with the properties:
Since , we have , so in . Moreover, as is a quadratic nonresidue modulo , so is . Since is a nonzero square in , we deduce that is a quadratic nonresidue modulo , as required.
Now after lifting from the smooth -point to a -point (with the same residue modulo ), if mod then we can take , and (which still satisfy the cubic equation mod ) to have mod since mod and is odd. Therefore, there exist integers satisfying the required properties by the Chinese Remainder Theorem.
Now assume that satisfies all the above congruence conditions and Assumption A. Since and , by Hensel’s lemma we deduce that when is a quadratic nonresidue modulo over , then is a quadratic nonresidue modulo over . Recalling that mod , we can choose a -point of such that
from which we have (in ):
and
Therefore , by Hensel’s lemma (fixing the variables ), this point lifts to a -point (abuse of notation) with the same residue modulo . Then since , the local invariant at of at this point is equal to
by the formulae in [Neu13, Proposition II.1.4 and Proposition III.3.3].
It is also clear that there exists a -point such that is not divisible by , which gives the local invariant of at the value . Indeed, if every -point satisfies that is divisible by , then as mod , the affine equation of over becomes
From the fact that is a quadratic nonresidue mod , one must obtain that (in ):
which gives . Therefore, in we have
which implies that . This cannot be true for every -point of , since by Assumption A we have proved in Proposition 4.1 that there always exists a non-singular solution modulo of the affine Markoff-type cubic equation (with at least one coordinate zero) which lifts to a -point by Hensel’s lemma.
In conclusion, the local invariant map of at the prime satisfying our hypothesis is indeed surjective. ∎
Theorem 4.3.
Let satisfy the hypotheses of Assumptions A and B. Then we have a Brauer–Manin obstruction to strong approximation for given by the class of in (i.e. ) and no algebraic Brauer–Manin obstruction to the integral Hasse principle for (i.e. ).
Proof.
By abuse of notation, we also denote the class of by the Brauer group element itself. For any point , from the above Lemma we can find a local point such that and such that . Then we obtain a point in by replacing the -part of u by , and a point in but not in by replacing the -part of u by .
Therefore, we have a Brauer–Manin obstruction to strong approximation and no algebraic Brauer–Manin obstruction to the integral Hasse principle for . ∎
Example 4.1.
For : Take .
Remark 4.2.
In addition, if satisfies Assumption 3.3 from the previous section, then we can drop the term “algebraic” from the statement of Theorem 4.3, and we will have no Brauer–Manin obstruction to the integral Hasse principle.
4.3 Some counting results
In this part, we compute the number of examples of existence for local integral points as well as the number of counterexamples to strong approximation for the Markoff-type cubic surfaces in question which can be explained by the Brauer–Manin obstruction. More precisely, we consider the natural density of satisfying mod , mod and mod , mod for , such that , for , and (then satisfying the additional hypothesis in Lemma 4.2). Note that the finite number of is negligible here.
In fact, for now we can only give an asymptotic lower bound. To get a lower bound, it is enough to count the number of examples satisfying stronger conditions, namely the congruence conditions for the and the common divisor condition
for . To do this, we make use of a natural generalization of Ekedahl–Poonen’s formula in [Poo03, Theorem 3.8] as follows.
Proposition 4.4.
Let , for some , be polynomials that are mutually relatively prime as elements of . For each , let
Then , where ranges over all primes of , and is the number of satisfying at least one of in for .
Proof.
We also have a generalization of [Poo03, Lemma 5.1] as follows.
Lemma 4.5.
Let , for some , be polynomials that are mutually relatively prime as elements of . For each , let
Then .
Proof.
The result immediately follows from the inequality and the original result of [Poo03, Lemma 5.1]. ∎
Next, we proceed similarly as in the proof of [Poo03, Theorem 3.1]. Let denote the set of prime numbers of such that . Approximate by
Define the ideal as the product of all for . Then is a union of cosets of the subgroup . Hence is the fraction of residue classes in in which for all , for all , at least one of and is nonzero modulo . Applying the Chinese Remainder Theorem, we obtain that . By the above lemma,
(24) |
Since and are relatively prime as elements of for any , there exists a nonzero such that every defines a subscheme of of codimension at least . Thus as , and the product converges. ∎
Theorem 4.6.
Let be the affine scheme over defined by
where
such that the projective closure of is smooth. Then we have
(25) |
and also
(26) |
as .
Proof.
Denote by the set in question and is the set of such that for . Hence, using the same notation as in Proposition 4.4, the density we are concerned with is given by
We apply Proposition 4.4 with the polynomials for in which we write , for (resp. where the is actually the and is the chosen prime in Lemma 4.2 and the residues satisfy the hypotheses of the lemma), and using the inclusion-exclusion principle we can compute that , ,
if and and
if and (resp. except for with mod , since , and (mod ), we have mod for all , and so Applying gives us a positive density in :
and also in :
Finally, we need to consider the number of surfaces which are singular (see necessary and sufficient conditions given in Lemma 3.6). By Lemma 3.6, the total number of with for such that the surfaces are singular is just as , hence it is negligible. Therefore, we obtain that . ∎
Remark 4.3.
Continuing from the previous remark, it would be interesting if one can find a way to include Assumption 3.3 into the counting result, which would help us consider the Brauer–Manin set with respect to the whole Brauer group instead of only its algebraic part.
5 Further remarks
In this section, we compare the results that we obtain in this paper with those in the previous papers studying Markoff surfaces, namely [GS22], [LM20] and [CWX20].
First of all, for Markoff surfaces, we see from [LM20] that given as , the number of counterexamples to the integral Hasse principle which can be explained by the Brauer–Manin obstruction is asymptotically. This implies that almost all Markoff surfaces with a nonempty set of local integral points have a nonempty Brauer–Manin set, and for the surfaces (relative character varieties) that we study in this chapter, a similar phenomenon is expected to occur (looking at the order of magnitude in the main counting result above). However, at present, we are not able to compute the number of these surfaces for which the Brauer–Manin obstruction can or cannot explain the integral Hasse principle in a similar way as in [LM20] and [CWX20]. If we can solve (partly) this problem, it will be really significant and then we may see a bigger picture of the arithmetic of Markoff-type cubic surfaces.
5.1 Markoff descent and reduction theory
Recall that, in order to show that the integral Hasse principle fails in [CWX20], the authors also make use of the fundamental set, or box, in [GS22] as a very useful tool to bound the set of integral points significantly and then use the Brauer group elements to finish the proofs. This mixed method has been the most effective way to prove counterexamples to the integral Hasse principle for Markoff surfaces which cannot be explained by the Brauer–Manin obstruction until now. However, in the case of Markoff-type cubic surfaces which we consider in this paper, the bounds do not prove themselves to be so effective when considered in a similar way. Indeed, let us recall the Markoff descent below for convenience; see [GS22] and [Wha20] for more details on the notation. Given , we write
Surfaces of type . Let be a surface of type . By Section 2, we have an identification of the moduli space with the affine cubic algebraic surface in given by the equation
The mapping class group acts on via polynomial transformations. Up to finite index, it coincides with the group of automorphisms of generated by the transpositions and even sign changes of coordinates as well as the Vieta involutions of the form .
The mapping class group dynamics on was analyzed in detail by Goldman as discussed in [Wha20], and the work of Ghosh–Sarnak [GS22] (Theorem 1.1) establishes a remarkable exact fundamental set for the action of on the integral points for admissible . More generally, the results below establish Markoff descent for complex points.
Lemma 5.1.
[Wha20, Lemma 4.2] Let be a surface of type . There is a constant independent of such that, given any , there exists some such that satisfies
Surfaces of type . Let be a surface of type . By Section 2, we have an identification of the moduli space with the affine cubic algebraic surface in given by the equation
with appropriately determined by . The mapping class group acts on via polynomial transformations. Let be the group of automorphisms of generated by the Vieta involutions
Two points are -equivalent if and only if they are -equivalent or is -equivalent to all of , and .
Lemma 5.2.
[Wha20, Lemma 4.4] Let be a surface of type . There is a constant independent of such that, given any , there exists some such that satisfies one of the following conditions:
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
,
-
(5)
.
Clearly, the restrictions in this lemma are weaker than those in the previous lemma, and so seems their effect. It would be interesting if one can find a way to apply these fundamental sets to produce some family of counterexamples to the integral Hasse principle which cannot be explained by the Brauer–Manin obstruction. A natural continuation from our work would be to find some sufficient hypothesis for such that the Brauer–Manin set of the general Markoff-type surface is nonempty, as inspired by [LM20, Corollary 5.11], which we will discuss in some particular cases in the next part. Ultimately, similar to the case of Markoff surfaces, it is still reasonable for us to expect that the number of counterexamples which cannot be explained by the Brauer–Manin obstruction for these relative character varieties is asymptotically greater than the number of those which can be explained by this obstruction.
Example 5.1.
Take , then satisfies Assumption A and we now have an explicit example of a Markoff-type cubic surface (where local integral points exist) given by the equation
By using the box as discussed above, we will prove that this example gives no integral solution, i.e., it is a counterexample to the integral Hasse principle.
Indeed, from the proof of [Wha20, Lemma 4.4], we can find that the constant independent of may take the value in the condition and in all the other conditions (thus we may choose 48 to be the desired constant in the statement of the Lemma). With this information, we can run a program on SageMath [SJ05] to find integral points satisfying one of the restrictions , along with the help of Dario Alpern’s website (Alpertron) [Alp] to find integral points whose one coordinate satisfies the restriction . More precisely, SageMath shows that there is no integral point in any of the boxes defined by , while Alpertron deals with conic equations after fixing one variable bounded in by transforming them into homogeneous quadratic equations and showing that some corresponding modular equations do not have solutions. As a result, we are able to prove that there is clearly no integral point in all those five cases, hence no integral point on the corresponding Markoff-type cubic surface.
5.2 Some special cases of Markoff-type cubic surfaces
Since in implies that either three of the are equal to or or any other permutation of (with ), the only cases that our Markoff-type cubic surfaces recover the original Markoff surfaces are given by equations of the form
or
for some . For the former equation, it is an interesting Markoff surface which can be studied similarly as in previous work, with a remark that for any odd the equation will not be everywhere locally solvable as the right hand side is congruent to mod (see [GS22]). For the latter equation, it always has the integral solutions . Although these special cases are only of magnitude compared to the total number of cases that we consider, it is clear that our family of examples do not recover these particular surfaces. We will discuss it here in a more general situation.
Naturally, one can find different hypotheses from that of Assumption A to work for more general cases, since Assumption A exists for technical reasons as well as specific counting results. More precisely, we will now consider the special case when are integers such that the total field extension and the set of local integral points on the Markoff-type cubic surface is nonempty. Interestingly, we are under the same framework as that of the general case in [CWX20, Section 3], to compute the algebraic Brauer group of the affine surface, which gives us
with three generators
Following [LM20, Proposition 5.7 and Lemma 5.8] or [CWX20, Lemmas 5.4 and 5.5], we have
and
as multisets. As are odd, so mod , hence there is no way to describe this number by the form for , which is favorable to give Hasse failures as in previous work. Due to the complexity of the integral values of the polynomial and their prime divisors, for now we can only deduce the vanishing of Brauer–Manin obstructions to the integral Hasse principle using a similar method as in [LM20].
Proposition 5.3.
Let are integers satisfying the congruence conditions mod (resp. mod ), and the divisibility conditions: , , and , such that . Moreover, assume that there is a prime such that divides to an odd order and mod . Then there is no algebraic Brauer–Manin obstruction to the integral Hasse principle, but there is a Brauer–Manin obstruction to strong approximation for .
Proof.
Following similar arguments as in the proof of Proposition 4.1, we can prove that the set of local integral points is indeed nonempty. Note that under our assumption, mod and mod , so mod in the defining equation of . Now for , we let and follow the same arguments in the proof of [LM20, Proposition 5.5] to prove that the map
induced by the Brauer–Manin pairing, is surjective. Now we need to show that . Let . Then by surjectivity, there exist such that for all and such that for some . Then we obtain a point in by replacing the -part of u by , and a point in but not in by replacing the -part of u by , as required.
We are left with the case . Using [LM20, Proposition 5.7] and [CWX20, Lemma 5.5], we know that the image of the above map induced by the Brauer–Manin pairing contains all the nontrivial elements. Let . In fact, there always exists such that for some . Next, if for some , there still exists such that for all . Now if for all , we can consider another local point whose -parts () are the same as those of u and -part is a permutation of (note that the Markoff-type cubic equation here is symmetric in ) such that their images under the local invariant map at are different permutations of , hence we will get for some . The proof is now complete. ∎
Again, the hypothesis of Proposition 5.3 can be modified or generalized to give the same results, which we hope to achieve in possible future work. For now, let us give some concrete examples from Proposition 5.3 using the help of SageMath [SJ05] and Alpertron [Alp] which have been mentioned in Example 5.1.
Example 5.2.
Consider and . Then there is no integral point on the corresponding Markoff-type cubic surface. By Proposition 5.3 (with ), we obtain a counterexample to the integral Hasse principle which cannot be explained by the algebraic Brauer–Manin obstruction.
Example 5.3.
Consider and . Then we find an integral point on the corresponding Markoff-type cubic surface. By Proposition 5.3 (with ), we get a counterexample to strong approximation which can be explained by the Brauer–Manin obstruction.
Finally, we end with a counterexample to the integral Hasse principle for which it is unclear whether the (algebraic) Brauer–Manin obstruction exists or not.
Example 5.4.
Consider and . Since , this example does not completely satisfy any of our previous assumptions. However, by using the programs on SageMath and Alpertron as discussed above, we find that there is indeed no integral point on the corresponding Markoff-type cubic surface.
References
- [Alp] Dario Alpern “Dario Alpern’s Web site : Generic two integer variable equation solver” URL: https://www.alpertron.com/QUAD.HTM
- [BGS16] J. Bourgain, A. Gamburd and P. Sarnak “Markoff surfaces and strong approximation: 1” In arxiv:1607.01530, 2016
- [BGS16a] J. Bourgain, A. Gamburd and P. Sarnak “Markoff triples and strong approximation” In Comptes Rendus Mathématique 354.2, 2016, pp. 131–135
- [CF67] J. Cassels and A. Fr$o$hlich “Algebraic Number Theory”, Proc. Instructional Conf. Academic Press, 1967
- [Che21] W.. Chen “Nonabelian level structures, Nielsen equivalence, and Markoff triples” In arxiv:2011.12940, 2021
- [CL09] S. Cantat and F. Loray “Dynamics on Character Varieties and Malgrange irreducibility of Painlevé VI equation” In Ann. Inst. Fourier 59.7, 2009, pp. 2927–2978
- [Con12] B. Conrad “Weil and Grothendieck approaches to adelic points” In Enseign. Math. 58.2, 2012, pp. 61–97
- [CS21] J.-L. Colliot-Thélène and A.. Skorobogatov “The Brauer–Grothendieck group”, Ergebnisse der Mathematik und ihrer Grenzgebiete, 3. Folge Springer Cham, 2021
- [CW12] J.-L. Colliot-Thélène and O. Wittenberg “Groupe de Brauer et points entiers de deux familles de surfaces cubiques affines” In Am. J. Math. 134.5, 2012, pp. 1303–1327
- [CWX20] J.-L. Colliot-Thélène, D. Wei and F. Xu “Brauer–Manin obstruction for Markoff surfaces” In Annali della Scuola Normale Superiore di Pisa XXI.5, 2020, pp. 1257–1313
- [CX09] J.-L. Colliot-Thélène and F. Xu “Brauer–Manin obstruction for integral points of homoheneous spaces and representation by integral quadratic forms” In Compos. Math. 145, 2009, pp. 309–363
- [GS17] P. Gille and T. Szamuely “Central simple algebras and Galois cohomology” Second Edition, Cambridge Studies in Advanced Mathematics 101 Cambridge University Press, 2017
- [GS22] A. Ghosh and P. Sarnak “Integral points on Markoff type cubic surfaces” In Invent. math. 229, 2022, pp. 689–749
- [Har08] D. Harari “Le défaut d’approximation forte pour les groupes algébriques commutatifs” In Algebra and Number Theory 2.5, 2008, pp. 595–611
- [Har77] R. Hartshorne “Algebraic Geometry”, Graduate Texts in Mathematics Springer-Verlag New York, 1977
- [Kol02] J. Kollár “Unirationality of cubic hypersurfaces” In J. Inst. Math. Jussieu 1.3, 2002, pp. 467–476
- [LM20] D. Loughran and V. Mitankin “Integral Hasse Principle and Strong Approximation for Markoff Surfaces” In International Mathematics Research Notices 2021.18, 2020, pp. 14086–14122
- [Man71] Yu.. Manin “Le groupe de Brauer–Grothendieck en géométrie diophantienne” (Nice, 1970), Actes Congr. Int. Math. Gauthier-Villars, 1971, pp. 14086–14122
- [Neu13] J. Neukirch “Class field theory”, The Bonn lectures Springer-Verlag Berlin Heidelberg New York, 2013
- [NSW15] J. Neukirch, A. Schmidt and K. Wingberg “Cohomology of Number Fields” Second Edition, Grundlehren der Math. 323 Springer-Verlag Berlin Heidelberg New York, 2015
- [Poo03] B. Poonen “Squarefree Values of Multivariate Polynomials” In Duke Math. J. 118.2, 2003, pp. 353–373
- [Poo17] B. Poonen “Rational points on varieties”, Graduate Studies in Mathematics 186 American Mathematical Society, 2017
- [SJ05] W. Stein and D. Joyner “SAGE: System for Algebra and Geometry Experimentation” In ACM SIGSAM Bulletin 39.2, 2005, pp. 61–64
- [Wha20] J.. Whang “Nonlinear descent on moduli of local systems” In Israel Journal of Mathematics 240, 2020, pp. 935–1004
CTXW20
Sorbonne Université and Université Paris Cité, CNRS, IMJ-PRG, F-75005 Paris, France
E-mail address: [email protected]
Institute of Mathematics, Vietnam Academy of Science and Technology, Cau Giay, 122300 Hanoi, Vietnam
E-mail address: [email protected]