This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

[columns=2, title=Index, intoc]

Brauer–Manin obstruction for integral points on Markoff-type cubic surfaces

Quang-Duc DAO
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 XX be an affine variety over \mathbb{Q}, and 𝒳\mathcal{X} an integral model of XX over \mathbb{Z}, i.e. an affine scheme of finite type over \mathbb{Z} whose generic fiber is isomorphic to UU. Define the set of adelic points X(A):=pX(p)X(\textbf{{A}}_{\mathbb{Q}}):=\sideset{}{{}^{\prime}}{\prod}_{p}X(\mathbb{Q}_{p}), where pp is a prime number or p=p=\infty (with =\mathbb{Q}_{\infty}=\mathbb{R}). Similarly, define 𝒳(A):=p𝒳(p)\mathcal{X}(\textbf{{A}}_{\mathbb{Z}}):=\prod_{p}\mathcal{X}(\mathbb{Z}_{p}) (with =\mathbb{Z}_{\infty}=\mathbb{R}). We say that XX fails the Hasse principle if

X(A)butX()=X(\textbf{{A}}_{\mathbb{Q}})\not=\emptyset\hskip 14.22636pt\textup{but}\hskip 14.22636ptX(\mathbb{Q})=\emptyset

We say that 𝒳\mathcal{X} fails the integral Hasse principle if

𝒳(A)but𝒳()=.\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})\not=\emptyset\hskip 14.22636pt\textup{but}\hskip 14.22636pt\mathcal{X}(\mathbb{Z})=\emptyset.

We say that XX satisfies weak approximation if the image of X()X(\mathbb{Q}) in vX(v)\prod_{v}X(\mathbb{Q}_{v}) is dense, where the product is taken over all places of \mathbb{Q}. Finally, we say that 𝒳\mathcal{X} satisfies strong approximation if 𝒳()\mathcal{X}(\mathbb{Z}) is dense in 𝒳(A):=p𝒳(p)×π0(X())\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}:=\prod_{p}\mathcal{X}(\mathbb{Z}_{p})\times\pi_{0}(X(\mathbb{R})), where π0(X())\pi_{0}(X(\mathbb{R})) denotes the set of connected components of X()X(\mathbb{R}). Note that we work with π0(X())\pi_{0}(X(\mathbb{R})) since 𝒳()\mathcal{X}(\mathbb{Z}) is never dense in X()X(\mathbb{R}) 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 BrX\textup{Br}\,X denotes the cohomological Brauer group of XX, i.e. BrX:=Hét2(X,𝔾m)\textup{Br}\,X:=\textup{H}^{2}_{\textup{ét}}(X,\mathbb{G}_{m}), we have a natural pairing from class field theory:

X(A)×BrX/.X(\textbf{{A}}_{\mathbb{Q}})\times\textup{Br}\,X\rightarrow\mathbb{Q}/\mathbb{Z}.

If we define X(A)BrX(\textbf{{A}}_{\mathbb{Q}})^{\textup{Br}} to be the left kernel of this pairing, then the exact sequence of Albert–Brauer–Hasse–Noether gives us the relation:

X()X(A)BrX(A).X(\mathbb{Q})\subseteq X(\textbf{{A}}_{\mathbb{Q}})^{\textup{Br}}\subseteq X(\textbf{{A}}_{\mathbb{Q}}).

Similarly, by defining the Brauer–Manin set 𝒳(A)Br\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{\textup{Br}}, we also have that

𝒳()𝒳(A)Br𝒳(A).\mathcal{X}(\mathbb{Z})\subseteq\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{\textup{Br}}\subseteq\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}.

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

X(A)BrX().X(\textbf{{A}}_{\mathbb{Q}})^{\textup{Br}}\not=\emptyset\iff X(\mathbb{Q})\not=\emptyset.
(𝒳(A)Br𝒳().)(\mathcal{X}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{\textup{Br}}\not=\emptyset\iff\mathcal{X}(\mathbb{Z})\not=\emptyset.)

If there is no confusion, we can omit the symbol \bullet for the set of local integral points and the corresponding Brauer–Manin set. \̃\ We are particularly interested in the case where XX is a hypersurface, defined by a polynomial equation of degree dd in an affine space. The case d=1d=1 is easy. The case d=2d=2 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 d=3d=3 (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

x3+y3+z3=a,x^{3}+y^{3}+z^{3}=a,

where aa is a fixed integer, Colliot-Thélène and Wittenberg in [CW12] proved that there is no Brauer–Manin obstruction to the integral Hasse principle (if aa is not of the form 9n±49n\pm 4). However, the existence of such an integer aa remains unknown in general, with the first example now being the case when a=114a=114. 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

x3+y3+2z3=a,x^{3}+y^{3}+2z^{3}=a,

for any aa\in\mathbb{Z}, also proven in [CW12].

Another interesting example of affine cubic surfaces that we consider is given by Markoff surfaces UmU_{m} which are defined by

x2+y2+z2xyz=m,x^{2}+y^{2}+z^{2}-xyz=m,

where mm is an integer parameter. The very first (original) class of Markoff surfaces which was studied is the one given by this equation with m=0m=0 in a series of papers [BGS16], [BGS16a], and recently [Che21]. They study strong approximation mod pp for U0U_{0} for any prime pp and present a Strong Approximation Conjecture ([BGS16a, Conjecture 1]):

Conjecture 1.1.

For any prime pp, U0(/p)U_{0}(\mathbb{Z}/p\mathbb{Z}) consists of two Γ\Gamma orbits, namely {(0,0,0)}\{(0,0,0)\} and U0(/p)=U0(/p)\{(0,0,0)}U_{0}^{*}(\mathbb{Z}/p\mathbb{Z})=U_{0}(\mathbb{Z}/p\mathbb{Z})\backslash\{(0,0,0)\}. Here Γ\Gamma is a group of affine integral morphisms of 𝔸3\mathbb{A}^{3} 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 U0U_{0} satisfies strong approximation mod pp for all by finitely many primes.

On the other hand, in [GS22], Ghosh and Sarnak study the integral points on those affine Markoff surfaces UmU_{m} with general mm, both from a theoretical point of view and by numerical evidence. They prove that for almost all mm, the integral Hasse principle holds, and that there are infinitely many mm’s for which it fails (Hasse failures). Furthermore, their numerical experiments suggest particularly a proportion of integers mm satisfying |m|M|m|\leq M of the power M0,8875+o(1)M^{0,8875\dots+o(1)} for which the integral Hasse principle is not satisfied.

Subsequently, Loughran and Mitankin [LM20] proved that asymptotically only a proportion of M1/2/(logM)1/2M^{1/2}/(\log M)^{1/2} of integers mm such that MmM-M\leq m\leq M presents an integral Brauer–Manin obstruction to the Hasse principle. They also obtained a lower bound, asymptotically M1/2/logMM^{1/2}/\log M, 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 M1/2/(logM)1/2M^{1/2}/(\log M)^{1/2} 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.

[GS22, Conjecture 10.2] The number of Hasse failures satisfies that

#{m:0mM,𝒰m(𝔸)but𝒰m()=}C0Mθ,\#\{m\in\mathbb{Z}:0\leq m\leq M,\ \mathcal{U}_{m}(\mathbb{A}_{\mathbb{Z}})\not=\emptyset\ \text{but}\ \mathcal{U}_{m}(\mathbb{Z})=\emptyset\}\approx C_{0}M^{\theta},

for some C0>0C_{0}>0 and some 12<θ<1\frac{1}{2}<\theta<1.

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 UmU_{m}, 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:

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where a,b,c,da,b,c,d\in\mathbb{Z} 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 𝒰\mathcal{U} be the affine scheme over \mathbb{Z} defined by

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where

{a=k1k2+k3k4b=k1k4+k2k3c=k1k3+k2k4andd=4i=14ki2i=14ki,\begin{cases*}a=k_{1}k_{2}+k_{3}k_{4}\\ b=k_{1}k_{4}+k_{2}k_{3}\\ c=k_{1}k_{3}+k_{2}k_{4}\end{cases*}\hskip 14.22636pt\textup{and}\hskip 14.22636ptd=4-\sum_{i=1}^{4}k_{i}^{2}-\prod_{i=1}^{4}k_{i},

such that the projective closure X3X\subset\mathbb{P}^{3}_{\mathbb{Q}} of U=𝒰×U=\mathcal{U}\times_{\mathbb{Z}}\mathbb{Q} is smooth. Then we have

#{k=(k1,k2,k3,k4)4,|ki|M 1i4:𝒰(A)Br1𝒰(A)}M4\#\{k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{Z}^{4},|k_{i}|\leq M\;\forall\;1\leq i\leq 4:\emptyset\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})\}\asymp M^{4}

as M+M\rightarrow+\infty.

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 kk be a field and k¯\overline{k} a separable closure of kk. We let Gk:=Gal(k¯/k)G_{k}:=\textup{Gal}(\overline{k}/k) be the absolute Galois group. A kk-variety is a separated kk-scheme of finite type. If XX is a kk-variety, we write X¯=X×kk¯\overline{X}=X\times_{k}\overline{k}. Let k[X]=H0(X,𝒪X)k[X]=\textup{H}^{0}(X,\mathcal{O}_{X}) and k¯[X]=H0(X¯,𝒪X¯)\overline{k}[X]=\textup{H}^{0}(\overline{X},\mathcal{O}_{\overline{X}}). If XX is an integral kk-variety, let k(X)k(X) denote the function field of XX. If XX is a geometrically integral kk-variety, let k¯(X)\overline{k}(X) denote the function field of X¯\overline{X}.

Let PicX=HZar1(X,𝔾m)=Hét1(X,𝔾m)\textup{Pic}\,X=\textup{H}^{1}_{\textup{Zar}}(X,\mathbb{G}_{m})=\textup{H}^{1}_{\textup{ét}}(X,\mathbb{G}_{m}) denote the Picard group of a scheme XX. Let BrX=Hét2(X,𝔾m)\textup{Br}\,X=\textup{H}^{2}_{\textup{ét}}(X,\mathbb{G}_{m}) denote the Brauer group of XX. Let

Br1X:=Ker[BrXBrX¯]\textup{Br}_{1}\,X:=\textup{Ker}[\textup{Br}\,X\rightarrow\textup{Br}\,\overline{X}]

denote the algebraic Brauer group of a kk-variety XX and let Br0XBr1X\textup{Br}_{0}\,X\subset\textup{Br}_{1}\,X denote the image of BrkBrX\textup{Br}\,k\rightarrow\textup{Br}\,X. The image of BrXBrX¯\textup{Br}\,X\rightarrow\textup{Br}\,\overline{X} is called the transcendental Brauer group of XX.

Given a field FF of characteristic zero containing a primitive nn-th root of unity ζ=ζn\zeta=\zeta_{n}, we have H2(F,μn2)=H2(F,μn)μn\textup{H}^{2}(F,\mu_{n}^{\otimes 2})=\textup{H}^{2}(F,\mu_{n})\otimes\mu_{n}. The choice of ζn\zeta_{n} then defines an isomorphism Br(F)[n]=H2(F,μn)H2(F,μn2)\textup{Br}(F)[n]=\textup{H}^{2}(F,\mu_{n})\cong\textup{H}^{2}(F,\mu_{n}^{\otimes 2}). Given two elements f,gF×f,g\in F^{\times}, we have their classes (f)(f) and (g)(g) in F×/F×n=H1(F,μn)F^{\times}/F^{\times n}=\textup{H}^{1}(F,\mu_{n}). We denote by (f,g)ζBr(F)[n]=H2(F,μn)(f,g)_{\zeta}\in\textup{Br}(F)[n]=\textup{H}^{2}(F,\mu_{n}) the class corresponding to the cup-product (f)(g)H2(F,μn2)(f)\cup(g)\in\textup{H}^{2}(F,\mu_{n}^{\otimes 2}). Suppose F/EF/E is a finite Galois extension with Galois group GG. Given σG\sigma\in G and f,gF×f,g\in F^{\times}, we have σ((f,g)ζn)=(σ(f),σ(g))σ(ζn)Br(F)\sigma((f,g)_{\zeta_{n}})=(\sigma(f),\sigma(g))_{\sigma(\zeta_{n})}\in\textup{Br}(F). In particular, if ζnE\zeta_{n}\in E, then σ((f,g)ζn)=(σ(f),σ(g))ζn\sigma((f,g)_{\zeta_{n}})=(\sigma(f),\sigma(g))_{\zeta_{n}}. For all the details, see [GS17, Sections 4.6, 4.7].

Let RR be a discrete valuation ring with fraction field FF and residue field κ\kappa. Let vv denote the valuation F×F^{\times}\rightarrow\mathbb{Z}. Let n>1n>1 be an integer invertible in RR. Assume that FF contains a primitive nn-th root of unity ζ\zeta. For f,gF×f,g\in F^{\times}, we have the residue map

R:H2(F,μn)H1(κ,/n)H1(κ,μn)=κ×/κ×n,\partial_{R}:\textup{H}^{2}(F,\mu_{n})\rightarrow\textup{H}^{1}(\kappa,\mathbb{Z}/n\mathbb{Z})\cong\textup{H}^{1}(\kappa,\mu_{n})=\kappa^{\times}/\kappa^{\times n},

where H1(κ,/n)H1(κ,μn)\textup{H}^{1}(\kappa,\mathbb{Z}/n\mathbb{Z})\cong\textup{H}^{1}(\kappa,\mu_{n}) is induced by the isomorphism /nμn\mathbb{Z}/n\mathbb{Z}\simeq\mu_{n} sending 11 to ζ\zeta. This map sends the class of (f,g)ζBr(F)[n]=H2(F,μn)(f,g)_{\zeta}\in\textup{Br}(F)[n]=\textup{H}^{2}(F,\mu_{n}) to

(1)v(f)v(g)class(gv(f)/fv(g))κ/κ×n.(-1)^{v(f)v(g)}\textup{class}(g^{v(f)}/f^{v(g)})\in\kappa/\kappa^{\times n}.

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 XX over a field kk, we denote by k[X]k[X] its coordinate ring over kk. If moreover XX is integral, then k(X)k(X) denotes its function field over kk. Given a commutative ring AA with unity, the elements of AA will be referred to as regular functions on the affine scheme SpecA\textup{Spec}\,A.

Definition 2.1.

Let π\pi be a finitely generated group. Its (SL2\textup{SL}_{2}) representation variety Rep(π)\textup{Rep}(\pi) is the affine scheme defined by the functor

AHom(π,SL2(A))A\mapsto\textup{Hom}(\pi,\textup{SL}_{2}(A))

for every commutative ring AA. Assume that π\pi has a sequence of generators of mm elements, then we have a presentation of Rep(π)\textup{Rep}(\pi) as a closed subscheme of SL2m\textup{SL}_{2}^{m} defined by equations coming from relations among the generators. For each aπa\in\pi, let tra\textup{tr}_{a} be the regular function on Rep(π)\textup{Rep}(\pi) given by ρtrρ(a)\rho\mapsto\textup{tr}\,\rho(a).

The (SL2\textup{SL}_{2}) character variety of π\pi over \mathbb{C} is then defined to be the affine invariant theoretic quotient

X(π):=Rep(π)SL2=Spec([Rep(π)]SL2())X(\pi):=\textup{Rep}(\pi)\sslash\textup{SL}_{2}=\textup{Spec}\left(\mathbb{C}[\textup{Rep}(\pi)]^{\textup{SL}_{2}(\mathbb{C})}\right)

under the conjugation action of SL2\textup{SL}_{2}.

The regular function tra\textup{tr}_{a} for each aπa\in\pi clearly descends to a regular function on X(π)X(\pi). Furthermore, from the fact that tr(I2)=2\textup{tr}(I_{2})=2 and tr(A)tr(B)=tr(AB)+tr(AB1)\textup{tr}(A)\textup{tr}(B)=\textup{tr}(AB)+\textup{tr}(AB^{-1}), for I2SL2()I_{2}\in\textup{SL}_{2}(\mathbb{C}) being the identity matrix and for any A,BSL2()A,B\in\textup{SL}_{2}(\mathbb{C}), we can deduce a natural model of X(π)X(\pi) over \mathbb{Z} as the spectrum of

R(π):=[tra:aπ]/(tr12,tratrbtrabtrab1).R(\pi):=\mathbb{Z}[\textup{tr}_{a}:a\in\pi]/(\textup{tr}_{1}-2,\textup{tr}_{a}\textup{tr}_{b}-\textup{tr}_{ab}-\textup{tr}_{ab^{-1}}).

Given any integral domain AA with fraction field FF of characteristic zero, the AA-points of X(π)X(\pi) parametrize the Jordan equivalence classes of SL2(F)\textup{SL}_{2}(F)-representations of π\pi having character valued in AA.

Example 2.2.

Denote by FmF_{m} the free group on m1m\geq 1 generators a1,,ama_{1},\dots,a_{m}. By Goldman’s results used in [Wha20], we have the following important examples:

  1. (1)

    tra1:X(F1)𝔸1\textup{tr}_{a_{1}}:X(F_{1})\simeq\mathbb{A}^{1}.

  2. (2)

    (tra1,tra2,tra3):X(F2)𝔸3(\textup{tr}_{a_{1}},\textup{tr}_{a_{2}},\textup{tr}_{a_{3}}):X(F_{2})\simeq\mathbb{A}^{3}.

  3. (3)

    The coordinate ring [X(F3)]\mathbb{Q}[X(F_{3})] is the quotient of the polynomial ring

    [tra1,tra2,tra3,tra1a2,tra2a3,tra1a3,tra1a2a3,tra1a3a2]\mathbb{Q}[\textup{tr}_{a_{1}},\textup{tr}_{a_{2}},\textup{tr}_{a_{3}},\textup{tr}_{a_{1}a_{2}},\textup{tr}_{a_{2}a_{3}},\textup{tr}_{a_{1}a_{3}},\textup{tr}_{a_{1}a_{2}a_{3}},\textup{tr}_{a_{1}a_{3}a_{2}}]

    by the ideal generated by two elements

    tra1a2a3+tra1a3a2(tra1a2tra3+tra1a3tra2+tra2a3tra1tra1tra2tra3)\textup{tr}_{a_{1}a_{2}a_{3}}+\textup{tr}_{a_{1}a_{3}a_{2}}-(\textup{tr}_{a_{1}a_{2}}\textup{tr}_{a_{3}}+\textup{tr}_{a_{1}a_{3}}\textup{tr}_{a_{2}}+\textup{tr}_{a_{2}a_{3}}\textup{tr}_{a_{1}}-\textup{tr}_{a_{1}}\textup{tr}_{a_{2}}\textup{tr}_{a_{3}})

    and

    tra1a2a3tra1a3a2{\displaystyle\textup{tr}_{a_{1}a_{2}a_{3}}\textup{tr}_{a_{1}a_{3}a_{2}}-\{ (tra12+tra22+tra32)+(tra1a22+tra2a32+tra1a32)\displaystyle(\textup{tr}_{a_{1}}^{2}+\textup{tr}_{a_{2}}^{2}+\textup{tr}_{a_{3}}^{2})+(\textup{tr}_{a_{1}a_{2}}^{2}+\textup{tr}_{a_{2}a_{3}}^{2}+\textup{tr}_{a_{1}a_{3}}^{2})
    (tra1tra2tra1a2+tra2tra3tra2a3+tra1tra3tra1a3)\displaystyle-(\textup{tr}_{a_{1}}\textup{tr}_{a_{2}}\textup{tr}_{a_{1}a_{2}}+\textup{tr}_{a_{2}}\textup{tr}_{a_{3}}\textup{tr}_{a_{2}a_{3}}+\textup{tr}_{a_{1}}\textup{tr}_{a_{3}}\textup{tr}_{a_{1}a_{3}})
    +tra1a2tra2a3tra1a34}.\displaystyle+\textup{tr}_{a_{1}a_{2}}\textup{tr}_{a_{2}a_{3}}\textup{tr}_{a_{1}a_{3}}-4\}.

Now given a connected smooth compact manifold MM, we consider the moduli of local systems on MM which is the character variety X(M):=X(π1(M))X(M):=X(\pi_{1}(M)) of its fundamental group. More generally, given a smooth manifold M=M1MmM=M_{1}\sqcup\cdots\sqcup M_{m} with finitely many connected components MiM_{i} for 1im1\leq i\leq m, define

X(M):=X(M1)××X(Mm).X(M):=X(M_{1})\times\cdots\times X(M_{m}).

The construction of the moduli space X(M)X(M) is functorial in the manifold MM. More precisely, any smooth map f:MNf:M\rightarrow N of manifolds induces a morphism f:X(N)X(M)f^{*}:X(N)\rightarrow X(M), depending only on the homotopy class of ff, given by pullback of local systems.

Let Σ\Sigma be a surface. For each curve aΣa\in\Sigma, there is a well-defined regular function tra:X(Σ)X(a)𝔸1\textup{tr}_{a}:X(\Sigma)\rightarrow X(a)\simeq\mathbb{A}^{1}, which agrees with trα\textup{tr}_{\alpha} for any απ1(Σ)\alpha\in\pi_{1}(\Sigma) freely homotopic to a parametrization of aa. The boundary curves Σ\partial\Sigma of Σ\Sigma induce a natural morphism

trΣ=()|Σ:X(Σ)X(Σ).\textup{tr}_{\partial\Sigma}=(-)|_{\partial\Sigma}:X(\Sigma)\rightarrow X(\partial\Sigma).

Now since we can write Σ=c1cn\partial\Sigma=c_{1}\sqcup\cdots\sqcup c_{n}, we have an identification

X(Σ)=X(c1)××X(cn)𝔸nX(\partial\Sigma)=X(c_{1})\times\dots\times X(c_{n})\simeq\mathbb{A}^{n}

given by taking a local system on the disjoint union Σ\partial\Sigma of nn circles to its sequence of traces along the curves. The morphism trΣ\textup{tr}_{\partial\Sigma} above may be viewed as an assignment to each ρX(Σ)\rho\in X(\Sigma) its sequence of traces trρ(c1),,trρ(cn)\textup{tr}\,\rho(c_{1}),\dots,\textup{tr}\,\rho(c_{n}). The fibers of trΣ\textup{tr}_{\partial\Sigma} for k𝔸nk\in\mathbb{A}^{n} will be denoted Xk=Xk(Σ)X_{k}=X_{k}(\Sigma). Each XkX_{k} is often called a relative character variety in the literature. If Σ\Sigma is a surface of type (g,n)(g,n) satisfying 3g+n3>03g+n-3>0, then the relative character variety Xk(Σ)X_{k}(\Sigma) is an irreducible algebraic variety of dimension 6g+2n66g+2n-6.

Given a fixed surface Σ\Sigma, a subset KX(Σ,)K\subseteq X(\partial\Sigma,\mathbb{C}), and a subset AA\subseteq\mathbb{C}, we shall denote by

XK(A)=XK(Σ,A):=XK(Σ)(A)X_{K}(A)=X_{K}(\Sigma,A):=X_{K}(\Sigma)(A)

the set of all ρX(Σ,)\rho\in X(\Sigma,\mathbb{C}) such that trΣ(ρ)K\textup{tr}_{\partial\Sigma}(\rho)\in K and tra(ρ)A\textup{tr}_{a}(\rho)\in A for every essential curve aΣa\subset\Sigma. By [Wha20, Lemma 2.5], there is no risk of ambiguity with this notation, i.e., XkX_{k} has a model over AA and Xk(A)X_{k}(A) recovers the set of AA-valued points of XkX_{k} in the sense of algebraic geometry.

2.2 Markoff-type cubic surfaces

Now we give a description of the moduli spaces Xk(Σ)X_{k}(\Sigma) for (g,n)=(1,1)(g,n)=(1,1) and (0,4)(0,4). These cases are special since each XkX_{k} is an affine cubic algebraic surface with an explicit equation.

  1. (1)

    Let Σ\Sigma be a surface of type (g,n)=(1,1)(g,n)=(1,1), i.e. a one holed torus. Let (α,β,γ)(\alpha,\beta,\gamma) be an optimal sequence of generators for π1(Σ)\pi_{1}(\Sigma), as given in [Wha20, Definition 2.1]. By Example 2.2, we have an identification X(Σ)𝔸3X(\Sigma)\simeq\mathbb{A}^{3}. From the trace relations in SL2\textup{SL}_{2}, we obtain that

    trγ\displaystyle\textup{tr}_{\gamma} =trαβα1β1=trαβα1trβ1trαβα1β\displaystyle=\textup{tr}_{\alpha\beta\alpha^{-1}\beta^{-1}}=\textup{tr}_{\alpha\beta\alpha^{-1}}\textup{tr}_{\beta^{-1}}-\textup{tr}_{\alpha\beta_{\alpha}^{-1}\beta}
    =trβ2trαβtrα1β+trαα=trβ2trαβ(trα1trβtrαβ)+trα2tr1\displaystyle=\textup{tr}_{\beta}^{2}-\textup{tr}_{\alpha\beta}\textup{tr}_{\alpha^{-1}\beta}+\textup{tr}_{\alpha\alpha}=\textup{tr}_{\beta}^{2}-\textup{tr}_{\alpha\beta}(\textup{tr}_{\alpha^{-1}}\textup{tr}_{\beta}-\textup{tr}_{\alpha\beta})+\textup{tr}_{\alpha}^{2}-\textup{tr}_{1}
    =trα2+trβ2+trαβ2trαtrβtrαβ2.\displaystyle=\textup{tr}_{\alpha}^{2}+\textup{tr}_{\beta}^{2}+\textup{tr}_{\alpha\beta}^{2}-\textup{tr}_{\alpha}\textup{tr}_{\beta}\textup{tr}_{\alpha\beta}-2.

    Writing (x,y,z)=(trα,trβ,trαβ)(x,y,z)=(\textup{tr}_{\alpha},\textup{tr}_{\beta},\textup{tr}_{\alpha\beta}) so that each of the variables xx, yy, and zz corresponds to an essential curve on Σ\Sigma as depicted in [Wha20, Figure 22], the moduli space XkXX_{k}\subset X has an explicit presentation as an affine cubic algebraic surface in 𝔸x,y,z3\mathbb{A}^{3}_{x,y,z} with the equation

    x2+y2+z2xyz2=k.x^{2}+y^{2}+z^{2}-xyz-2=k.

    These are exactly the Markoff surfaces as studied in the series of papers [GS22], [LM20], and [CWX20] with m=k+2m=k+2.

  2. (2)

    Let Σ\Sigma be a surface of type (g,n)=(0,4)(g,n)=(0,4), i.e. a four holed sphere. Let (γ1,γ2,γ3,γ4)(\gamma_{1},\gamma_{2},\gamma_{3},\gamma_{4}) be an optimal sequence of generators for π1(Σ)\pi_{1}(\Sigma). Set

    (x,y,z)=(trγ1γ2,trγ2γ3,trγ1γ3)(x,y,z)=(\textup{tr}_{\gamma_{1}\gamma_{2}},\textup{tr}_{\gamma_{2}\gamma_{3}},\textup{tr}_{\gamma_{1}\gamma_{3}})

    so that each of the variables corresponds to an essential curve on Σ\Sigma. By Example 2.2, for k=(k1,k2,k3,k4)𝔸4()k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(\mathbb{C}) the relative character variety Xk=Xk(Σ)X_{k}=X_{k}(\Sigma) is an affine cubic algebraic surface in 𝔸x,y,z3\mathbb{A}^{3}_{x,y,z} given by the equation

    x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d, (1)

    where

    {a=k1k2+k3k4b=k1k4+k2k3c=k1k3+k2k4andd=4i=14ki2i=14ki.\begin{cases*}a=k_{1}k_{2}+k_{3}k_{4}\\ b=k_{1}k_{4}+k_{2}k_{3}\\ c=k_{1}k_{3}+k_{2}k_{4}\end{cases*}\hskip 14.22636pt\textup{and}\hskip 14.22636ptd=4-\sum_{i=1}^{4}k_{i}^{2}-\prod_{i=1}^{4}k_{i}. (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 (1)(1). 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 kk be an arbitrary field. Recall that for a variety XX over kk there is a natural filtration on the Brauer group

Br0XBr1XBrX\textup{Br}_{0}\,X\subset\textup{Br}_{1}\,X\subset\textup{Br}\,X

which is defined as follows.

Definition 3.1.

Let

Br0X=Im[BrkBrX],Br1X=Ker[BrXBrX¯].\textup{Br}_{0}\,X=\textup{Im}[\textup{Br}\,k\rightarrow\textup{Br}\,X],\hskip 14.22636pt\textup{Br}_{1}\,X=\textup{Ker}[\textup{Br}\,X\rightarrow\textup{Br}\,\overline{X}].

The subgroup Br1XBrX\textup{Br}_{1}\,X\subset\textup{Br}\,X is the algebraic Brauer group of XX and the quotient BrX/Br1X\textup{Br}\,X/\textup{Br}_{1}\,X is the transcendental Brauer group of XX.

From the Hochschild–Serre spectral sequence, we have the following spectral sequence:

E2pq=Hétp(k,Hétq(X¯,𝔾m))Hétp+q(X,𝔾m),E_{2}^{pq}=\textup{H}^{p}_{\textup{\'{e}t}}(k,\textup{H}_{\textup{\'{e}t}}^{q}(\overline{X},\mathbb{G}_{m}))\Longrightarrow\textup{H}^{p+q}_{\textup{\'{e}t}}(X,\mathbb{G}_{m}),

which is contravariantly functorial in the kk-variety XX. It gives rise to the functorial exact sequence of terms of low degree:

0\displaystyle 0 H1(k,k¯[X]×)PicXPicX¯GkH2(k,k¯[X]×)Br1X\displaystyle\longrightarrow\textup{H}^{1}(k,\overline{k}[X]^{\times})\longrightarrow\textup{Pic}\,X\longrightarrow\textup{Pic}\,\overline{X}^{G_{k}}\longrightarrow\textup{H}^{2}(k,\overline{k}[X]^{\times})\longrightarrow\textup{Br}_{1}\,X
H1(k,PicX¯)Ker[H3(k,k¯[X]×)Hét3(X,𝔾m)].\displaystyle\longrightarrow\textup{H}^{1}(k,\textup{Pic}\,\overline{X})\longrightarrow\textup{Ker}[\textup{H}^{3}(k,\overline{k}[X]^{\times})\rightarrow\textup{H}^{3}_{\textup{\'{e}t}}(X,\mathbb{G}_{m})].

Let XX be a variety over a field kk such that k¯[X]×=k¯×\overline{k}[X]^{\times}=\overline{k}^{\times}. By Hilbert’s Theorem 90 we have H1(k,k¯×)=0\textup{H}^{1}(k,\overline{k}^{\times})=0, then by the above sequence there is an exact sequence

0\displaystyle 0 PicXPicX¯GkBrkBr1X\displaystyle\longrightarrow\textup{Pic}\,X\longrightarrow\textup{Pic}\,\overline{X}^{G_{k}}\longrightarrow\textup{Br}\,k\longrightarrow\textup{Br}_{1}\,X
H1(k,PicX¯)Ker[H3(k,k¯×)Hét3(X,𝔾m)].\displaystyle\longrightarrow\textup{H}^{1}(k,\textup{Pic}\,\overline{X})\longrightarrow\textup{Ker}[\textup{H}^{3}(k,\overline{k}^{\times})\rightarrow\textup{H}^{3}_{\textup{\'{e}t}}(X,\mathbb{G}_{m})].

This sequence is also contravariantly functorial in XX.

Remark 3.2.

Let XX be a variety over a field kk such that k¯[X]×=k¯×\overline{k}[X]^{\times}=\overline{k}^{\times}. This assumption k¯[X]×=k¯×\overline{k}[X]^{\times}=\overline{k}^{\times} holds for any proper, geometrically connected and geometrically reduced kk-variety XX.

  1. (1)

    If XX has a kk-point, which defined a section of the structure morphism XSpeckX\rightarrow\textup{Spec}\,k, then each of the maps BrkBr1X\textup{Br}\,k\longrightarrow\textup{Br}_{1}\,X and H3(k,k¯×)Hét3(X,𝔾m)\textup{H}^{3}(k,\overline{k}^{\times})\rightarrow\textup{H}^{3}_{\textup{\'{e}t}}(X,\mathbb{G}_{m}) has a retraction, hence is injective. (Then PicXPicX¯Gk\textup{Pic}\,X\longrightarrow\textup{Pic}\,\overline{X}^{G_{k}} is an isomorphism.) Therefore, we have an isomorphism

    Br1X/BrkH1(k,PicX¯).\textup{Br}_{1}\,X/\textup{Br}\,k\cong\textup{H}^{1}(k,\textup{Pic}\,\overline{X}).
  2. (2)

    If kk is a number field, then H3(k,k¯×)=0\textup{H}^{3}(k,\overline{k}^{\times})=0 (see [CF67], Chapter VII, Section 11.4, p. 199). Thus for a variety XX over a number field kk such that k¯[X]×=k¯×\overline{k}[X]^{\times}=\overline{k}^{\times}, we have an isomorphism

    Br1X/Br0XH1(k,PicX¯).\textup{Br}_{1}\,X/\textup{Br}_{0}\,X\cong\textup{H}^{1}(k,\textup{Pic}\,\overline{X}).

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

U:f(u1,u2,u3)=0U:f(u_{1},u_{2},u_{3})=0

where ff is a polynomial of degree of 3. The closure of UU in 3\mathbb{P}^{3} is a cubic surface XX. The complement H=X\UH=X\ \backslash\ U is a hyperplane section on SS. Much of the geometry of UU can be understood in terms of the geometry of XX and HH, 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 HH is smooth, for example see [CW12]. Here we shall be interested in the case where the hyperplane section HH is singular; in particular, we focus on the case where HH 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 Xk3X\subset\mathbb{P}^{3}_{k} be a smooth projective cubic surface over a field kk of characteristic zero. Suppose a plane k2k3\mathbb{P}^{2}_{k}\subset\mathbb{P}^{3}_{k} cuts out on X¯\overline{X} three distinct lines L1,L2,L3L_{1},L_{2},L_{3} over k¯\overline{k}. Let UXU\subset X be the complement of this plane. Then the natural map k¯×k¯[U]×\overline{k}^{\times}\rightarrow\overline{k}[U]^{\times} is an isomorphism of Galois modules and the natural sequence

0i=13LiPicX¯PicU¯00\longrightarrow\bigoplus_{i=1}^{3}\mathbb{Z}L_{i}\longrightarrow\textup{Pic}\,\overline{X}\longrightarrow\textup{Pic}\,\overline{U}\longrightarrow 0

is an exact sequence of Galois lattices.

As PicU¯\text{Pic}\,\overline{U} 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 XX be a smooth projective cubic surface over a field kk of characteristic 0. Let HSH\subset S be a hyperplane section which is the union of 3 distinct lines L1,L2,L3L_{1},L_{2},L_{3} and let U=X\HU=X\ \backslash\ H. Then PicU¯\textup{Pic}\,\overline{U} is torsion free and Br1U/Br0UH1(k,PicU¯)\textup{Br}_{1}\,U/\textup{Br}_{0}\,U\cong\textup{H}^{1}(k,\textup{Pic}\,\overline{U}) is isomorphic to one of the following groups:

0,/4,/2×/4,(/2)r(r=1,2,3,4).0,\mathbb{Z}/4\mathbb{Z},\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/4\mathbb{Z},(\mathbb{Z}/2\mathbb{Z})^{r}\hskip 14.22636pt(r=1,2,3,4).

For the transcendental Brauer group, from the discussion on page 140 of [CS21], note that Br1U=Ker(BrUBrU¯Gk)\text{Br}_{1}\,U=\text{Ker}(\text{Br}\,U\rightarrow\text{Br}\,\overline{U}^{G_{k}}) so we have BrU/Br1UBrU¯Gk\text{Br}\,U/\text{Br}_{1}\,U\subset\text{Br}\,\overline{U}^{G_{k}}.

Proposition 3.3.

[CWX20, Proposition 2.1] [LM20, Proposition 2.4] Let XX be a smooth projective cubic surface over a field kk of characteristic 0. Suppose that UU is an open subset of XX such that X\UX\ \backslash\ U is the union of three distinct kk-lines, by which we mean a smooth projective curve isomorphic to k1\mathbb{P}^{1}_{k}. Suppose any two lines intersect each another transversely in one point, and that the three intersection points are distinct. Let LL be one of the three lines and VLV\subset L be the complement of the 2 intersection points of LL with the other two lines. Then the residue map

L:Brk¯(X)H1(k¯(L),/)\partial_{L}:\textup{Br}\,\overline{k}(X)\rightarrow\textup{H}^{1}(\overline{k}(L),\mathbb{Q}/\mathbb{Z})

induces a GkG_{k}-isomorphism

BrU¯H1(V¯,/)H1(𝔾¯m,/)/(1).\textup{Br}\,\overline{U}\simeq\textup{H}^{1}(\overline{V},\mathbb{Q}/\mathbb{Z})\simeq\textup{H}^{1}(\overline{\mathbb{G}}_{m},\mathbb{Q}/\mathbb{Z})\simeq\mathbb{Q}/\mathbb{Z}(-1).

In particular, if kk contains no non-trivial roots of unity then

BrU¯Gk=/2.\textup{Br}\,\overline{U}^{G_{k}}=\mathbb{Z}/2\mathbb{Z}.
Lemma 3.4.

[CWX20, Lemma 2.4] Let kk be a field of characteristic 0. Let Gk=Gal(k¯/k)G_{k}=\textup{Gal}(\overline{k}/k). Then /(1)Gk\mathbb{Q}/\mathbb{Z}(-1)^{G_{k}} is (noncanonically) isomorphic to μ(k)\mu_{\infty}(k), the group of roots of unity in kk.

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 kk be a field of characteristic 0 such that in any finite field extension there are only finitely many roots of unity. Let Xk3X\subset\mathbb{P}^{3}_{k} be a smooth projective cubic surface over kk. Suppose that a plane cuts out on XX three nonconcurrent lines. Let UXU\subset X be the complement of the plane section. Then the quotient BrU/Br0U\textup{Br}\,U/\textup{Br}_{0}\,U 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 KK be a number field and let XK3X\subset\mathbb{P}^{3}_{K} be a cubic surface defined by the equation

t(x2+y2+z2)+xyz=t2(ax+by+cz)+dt3,t(x^{2}+y^{2}+z^{2})+xyz=t^{2}(ax+by+cz)+dt^{3},

where a,b,c,da,b,c,d are defined by (2)(2) for some k=(k1,k2,k3,k4)𝔸4(K)k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(K). Then XX is singular if and only if we are in one of the following cases:

  1. \bullet

    Δ(k)=0\Delta(k)=0 where k=(k1,k2,k3,k4)𝔸4(K)k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(K) and

    Δ(k)=(2(k12+k22+k32+k42)k1k2k3k416)2(4k12)(4k22)(4k32)(4k42),\Delta(k)=(2(k_{1}^{2}+k_{2}^{2}+k_{3}^{2}+k_{4}^{2})-k_{1}k_{2}k_{3}k_{4}-16)^{2}-(4-k_{1}^{2})(4-k_{2}^{2})(4-k_{3}^{2})(4-k_{4}^{2}),
  2. \bullet

    at least one of the parameters k1,k2,k3,k4k_{1},k_{2},k_{3},k_{4} equals ±2\pm 2.

If kk satisfies none of those two conditions and [E:K]=16[E:K]=16 where

E:=K(k124,k224,k324,k424),E:=K(\sqrt{k_{1}^{2}-4},\sqrt{k_{2}^{2}-4},\sqrt{k_{3}^{2}-4},\sqrt{k_{4}^{2}-4}),

then the 27 lines on the smooth cubic surface X¯\overline{X} are defined over EE by the following equations

L1:x=t=0;L2:y=t=0;L3:z=t=0L_{1}:x=t=0;\hskip 28.45274ptL_{2}:y=t=0;\hskip 28.45274ptL_{3}:z=t=0

and

  1. 1.

    1(ϵ,δ):x=(k1k2+ϵδ(k124)(k224))2t,y=(k1+ϵk124)(k2+δk224)4zcb(k1+ϵk124)(k2+δk224)4δk1k224+ϵk2k1242t;\ell_{1}(\epsilon,\delta):x=\displaystyle\frac{(k_{1}k_{2}+\epsilon\delta\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)})}{2}t,\\ y=-\displaystyle\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{2}+\delta\sqrt{k_{2}^{2}-4})}{4}z-\displaystyle\frac{c-b\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{2}+\delta\sqrt{k_{2}^{2}-4})}{4}}{\frac{\delta k_{1}\sqrt{k_{2}^{2}-4}+\epsilon k_{2}\sqrt{k_{1}^{2}-4}}{2}}t;

  2. 2.

    2(ϵ,δ):y=(k1k4+ϵδ(k124)(k424))2t,z=(k1+ϵk124)(k4+δk424)4xac(k1+ϵk124)(k4+δk424)4δk1k424+ϵk4k1242t;\ell_{2}(\epsilon,\delta):y=\displaystyle\frac{(k_{1}k_{4}+\epsilon\delta\sqrt{(k_{1}^{2}-4)(k_{4}^{2}-4)})}{2}t,\\ z=-\displaystyle\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}x-\displaystyle\frac{a-c\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}}{\frac{\delta k_{1}\sqrt{k_{4}^{2}-4}+\epsilon k_{4}\sqrt{k_{1}^{2}-4}}{2}}t;

  3. 3.

    3(ϵ,δ):z=(k1k3+ϵδ(k124)(k324))2t,y=(k1+ϵk124)(k3+δk324)4xab(k1+ϵk124)(k3+δk324)4δk1k324+ϵk3k1242t;\ell_{3}(\epsilon,\delta):z=\displaystyle\frac{(k_{1}k_{3}+\epsilon\delta\sqrt{(k_{1}^{2}-4)(k_{3}^{2}-4)})}{2}t,\\ y=-\displaystyle\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{3}+\delta\sqrt{k_{3}^{2}-4})}{4}x-\displaystyle\frac{a-b\frac{(k_{1}+\epsilon\sqrt{k_{1}^{2}-4})(k_{3}+\delta\sqrt{k_{3}^{2}-4})}{4}}{\frac{\delta k_{1}\sqrt{k_{3}^{2}-4}+\epsilon k_{3}\sqrt{k_{1}^{2}-4}}{2}}t;

  4. 4.

    4(ϵ,δ):x=(k3k4+ϵδ(k324)(k424))2t,y=(k3+ϵk324)(k4+δk424)4zcb(k3+ϵk324)(k4+δk424)4δk3k424+ϵk4k3242t;\ell_{4}(\epsilon,\delta):x=\displaystyle\frac{(k_{3}k_{4}+\epsilon\delta\sqrt{(k_{3}^{2}-4)(k_{4}^{2}-4)})}{2}t,\\ y=-\displaystyle\frac{(k_{3}+\epsilon\sqrt{k_{3}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}z-\displaystyle\frac{c-b\frac{(k_{3}+\epsilon\sqrt{k_{3}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}}{\frac{\delta k_{3}\sqrt{k_{4}^{2}-4}+\epsilon k_{4}\sqrt{k_{3}^{2}-4}}{2}}t;

  5. 5.

    5(ϵ,δ):y=(k2k3+ϵδ(k224)(k324))2t,z=(k2+ϵk224)(k3+δk324)4xac(k2+ϵk224)(k3+δk324)4δk2k324+ϵk3k2242t;\ell_{5}(\epsilon,\delta):y=\displaystyle\frac{(k_{2}k_{3}+\epsilon\delta\sqrt{(k_{2}^{2}-4)(k_{3}^{2}-4)})}{2}t,\\ z=-\displaystyle\frac{(k_{2}+\epsilon\sqrt{k_{2}^{2}-4})(k_{3}+\delta\sqrt{k_{3}^{2}-4})}{4}x-\displaystyle\frac{a-c\frac{(k_{2}+\epsilon\sqrt{k_{2}^{2}-4})(k_{3}+\delta\sqrt{k_{3}^{2}-4})}{4}}{\frac{\delta k_{2}\sqrt{k_{3}^{2}-4}+\epsilon k_{3}\sqrt{k_{2}^{2}-4}}{2}}t;

  6. 6.

    6(ϵ,δ):z=(k2k4+ϵδ(k224)(k424))2t,y=(k2+ϵk224)(k4+δk424)4xab(k2+ϵk224)(k4+δk424)4δk2k424+ϵk4k2242t\ell_{6}(\epsilon,\delta):z=\displaystyle\frac{(k_{2}k_{4}+\epsilon\delta\sqrt{(k_{2}^{2}-4)(k_{4}^{2}-4)})}{2}t,\\ y=-\displaystyle\frac{(k_{2}+\epsilon\sqrt{k_{2}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}x-\displaystyle\frac{a-b\frac{(k_{2}+\epsilon\sqrt{k_{2}^{2}-4})(k_{4}+\delta\sqrt{k_{4}^{2}-4})}{4}}{\frac{\delta k_{2}\sqrt{k_{4}^{2}-4}+\epsilon k_{4}\sqrt{k_{2}^{2}-4}}{2}}t

with ϵ=±1\epsilon=\pm 1 and δ=±1\delta=\pm 1. Furthermore, we have the intersection numbers

i(ϵ,δ).j(ϵ,δ)=0\ell_{i}(\epsilon,\delta).\ell_{j}(\epsilon,\delta)=0

for any pair (ϵ,δ)(\epsilon,\delta), for all 1ij61\leq i\not=j\leq 6.

Proof.

The necessary and sufficient condition for the affine open surface U=X{t=0}U=X\setminus\{t=0\} 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 XX.

Now without loss of generality, we consider the system of equations

{y=α1x+α2tz=β1x+β2t\begin{cases*}y=\alpha_{1}x+\alpha_{2}t\\ z=\beta_{1}x+\beta_{2}t\end{cases*}

and put them in the original equation of the cubic surfaces to solve αi,βi\alpha_{i},\beta_{i} for i=1,2i=1,2. We can work similarly for (z,x)(z,x) and (x,y)(x,y) 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 KK be a number field. Let XK3X\subset\mathbb{P}^{3}_{K} be a cubic surface defined by the equation

t(x2+y2+z2)+xyz=t2(ax+by+cz)+dt3,t(x^{2}+y^{2}+z^{2})+xyz=t^{2}(ax+by+cz)+dt^{3}, (3)

where a,b,c,da,b,c,d are defined by (2)(2) for some k=(k1,k2,k3,k4)𝔸4(K)k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(K). Assume that XX is smooth over KK and [E:K]=16[E:K]=16, then

BrX/Br0X=Br1X/Br0X/2.\textup{Br}\,X/\textup{Br}_{0}\,X=\textup{Br}_{1}\,X/\textup{Br}_{0}\,X\cong\mathbb{Z}/2\mathbb{Z}.
Proof.

Since XX is geometrically rational, one has BrX=Br1X\textup{Br}\,X=\textup{Br}_{1}\,X. By taking x=t=0x=t=0 for instance, one clearly has X(K)X(K)\not=\emptyset, so Br0X=BrK\textup{Br}_{0}\,X=\textup{Br}\,K. Since KK is a number field, by the Hochschild–Serre spectral sequence, we have an isomorphism

Br1X/Br0XH1(K,PicX¯).\textup{Br}_{1}\,X/\textup{Br}_{0}\,X\simeq\textup{H}^{1}(K,\textup{Pic}\,\overline{X}).

By the above lemma, we can easily verify that the six lines 1(1,1)\ell_{1}(1,1), 1(1,1)\ell_{1}(1,-1), 3(1,1)\ell_{3}(-1,1), 4(1,1)\ell_{4}(-1,-1), 4(1,1)\ell_{4}(-1,1), and L2L_{2} on the cubic surface X¯\overline{X} are skew to each other, hence they may be simultaneously blown down to 2\mathbb{P}^{2} by [Har77], Chapter V, Proposition 4.10. For the sake of simplicity, here we shall write these six lines respectively as i\ell_{i} for 1i61\leq i\leq 6. The class ω\omega of the canonical divisor on X¯\overline{X} is equal to 3+Σi=16i-3\ell+\Sigma_{i=1}^{6}\ell_{i}, where \ell is the inverse image of the class of lines in 2\mathbb{P}^{2}. By [Har77, Chapter V, Proposition 4.8], the classes ,i,i=1,6¯\ell,\ell_{i},i=\overline{1,6} form a basis of PicX¯\textup{Pic}\,\overline{X}, and we have the following intersection properties: (.)=1,(.i)=0(\ell.\ell)=1,(\ell.\ell_{i})=0 for 1i61\leq i\leq 6.

Since (L1.3)=0,(L1.i)=1,i3(L_{1}.\ell_{3})=0,(L_{1}.\ell_{i})=1,i\not=3; (L3.3)=(L3.6)=1,(L3.i)=0,i3,6(L_{3}.\ell_{3})=(L_{3}.\ell_{6})=1,(L_{3}.\ell_{i})=0,i\not=3,6; and L2=6L_{2}=\ell_{6}, one concludes that

L1=2Σi3i,L3=36L_{1}=2\ell-\Sigma_{i\not=3}\ell_{i},\hskip 28.45274ptL_{3}=\ell-\ell_{3}-\ell_{6} (4)

in PicX¯\textup{Pic}\,\overline{X} by [Har77, Chapter V, Proposition 4.8 (e)].

Now we consider the action of the Galois group G:=Gal(E/K)G:=\textup{Gal}(E/K) on PicX¯\textup{Pic}\,\overline{X}. One clearly has Gσ1×σ2×σ3×σ4G\cong\langle\sigma_{1}\rangle\times\langle\sigma_{2}\rangle\times\langle\sigma_{3}\rangle\times\langle\sigma_{4}\rangle, where

σi(ki24)=ki24andσi(kj24)=kj24,1ij4.\sigma_{i}(\sqrt{k_{i}^{2}-4})=-\sqrt{k_{i}^{2}-4}\hskip 14.22636pt\textup{and}\hskip 14.22636pt\sigma_{i}(\sqrt{k_{j}^{2}-4})=\sqrt{k_{j}^{2}-4},1\leq i\not=j\leq 4.

We have the following intersection numbers, noting that σ2(1)=2\sigma_{2}(\ell_{1})=\ell_{2}, σ2(2)=1\sigma_{2}(\ell_{2})=\ell_{1}, σ4(4)=5\sigma_{4}(\ell_{4})=\ell_{5}, and σ4(5)=4\sigma_{4}(\ell_{5})=\ell_{4}:

{(σ1(1).1)=(1(1,1).1(1,1))=0(σ1(1).2)=(1(1,1).1(1,1))=1(σ1(1).3)=(1(1,1).3(1,1))=1(σ1(1).4)=(1(1,1).4(1,1))=0(σ1(1).5)=(1(1,1).4(1,1))=0(σ1(1).6)=(1(1,1).L2)=0,\begin{cases}(\sigma_{1}(\ell_{1}).\ell_{1})=(\ell_{1}(-1,1).\ell_{1}(1,1))=0\\ (\sigma_{1}(\ell_{1}).\ell_{2})=(\ell_{1}(-1,1).\ell_{1}(1,-1))=1\\ (\sigma_{1}(\ell_{1}).\ell_{3})=(\ell_{1}(-1,1).\ell_{3}(-1,1))=1\\ (\sigma_{1}(\ell_{1}).\ell_{4})=(\ell_{1}(-1,1).\ell_{4}(-1,-1))=0\\ (\sigma_{1}(\ell_{1}).\ell_{5})=(\ell_{1}(-1,1).\ell_{4}(-1,1))=0\\ (\sigma_{1}(\ell_{1}).\ell_{6})=(\ell_{1}(-1,1).L_{2})=0,\end{cases} (5)
{(σ1(2).1)=(1(1,1).1(1,1))=1(σ1(2).2)=(1(1,1).1(1,1))=0(σ1(2).3)=(1(1,1).3(1,1))=1(σ1(2).4)=(1(1,1).4(1,1))=0(σ1(2).5)=(1(1,1).4(1,1))=0(σ1(2).6)=(1(1,1).L2)=0,\begin{cases}(\sigma_{1}(\ell_{2}).\ell_{1})=(\ell_{1}(-1,-1).\ell_{1}(1,1))=1\\ (\sigma_{1}(\ell_{2}).\ell_{2})=(\ell_{1}(-1,-1).\ell_{1}(1,-1))=0\\ (\sigma_{1}(\ell_{2}).\ell_{3})=(\ell_{1}(-1,-1).\ell_{3}(-1,1))=1\\ (\sigma_{1}(\ell_{2}).\ell_{4})=(\ell_{1}(-1,-1).\ell_{4}(-1,-1))=0\\ (\sigma_{1}(\ell_{2}).\ell_{5})=(\ell_{1}(-1,-1).\ell_{4}(-1,1))=0\\ (\sigma_{1}(\ell_{2}).\ell_{6})=(\ell_{1}(-1,-1).L_{2})=0,\end{cases} (6)
{(σ1(3).1)=(3(1,1).1(1,1))=1(σ1(3).2)=(3(1,1).1(1,1))=1(σ1(3).3)=(3(1,1).3(1,1))=0(σ1(3).4)=(3(1,1).4(1,1))=0(σ1(3).5)=(3(1,1).4(1,1))=0(σ1(3).6)=(3(1,1).L2)=0;\begin{cases}(\sigma_{1}(\ell_{3}).\ell_{1})=(\ell_{3}(1,1).\ell_{1}(1,1))=1\\ (\sigma_{1}(\ell_{3}).\ell_{2})=(\ell_{3}(1,1).\ell_{1}(1,-1))=1\\ (\sigma_{1}(\ell_{3}).\ell_{3})=(\ell_{3}(1,1).\ell_{3}(-1,1))=0\\ (\sigma_{1}(\ell_{3}).\ell_{4})=(\ell_{3}(1,1).\ell_{4}(-1,-1))=0\\ (\sigma_{1}(\ell_{3}).\ell_{5})=(\ell_{3}(1,1).\ell_{4}(-1,1))=0\\ (\sigma_{1}(\ell_{3}).\ell_{6})=(\ell_{3}(1,1).L_{2})=0;\end{cases} (7)

and

{(σ3(3).1)=(3(1,1).1(1,1))=0(σ3(3).2)=(3(1,1).1(1,1))=0(σ3(3).3)=(3(1,1).3(1,1))=0(σ3(3).4)=(3(1,1).4(1,1))=1(σ3(3).5)=(3(1,1).4(1,1))=1(σ3(3).6)=(3(1,1).L2)=0,\begin{cases}(\sigma_{3}(\ell_{3}).\ell_{1})=(\ell_{3}(-1,-1).\ell_{1}(1,1))=0\\ (\sigma_{3}(\ell_{3}).\ell_{2})=(\ell_{3}(-1,-1).\ell_{1}(1,-1))=0\\ (\sigma_{3}(\ell_{3}).\ell_{3})=(\ell_{3}(-1,-1).\ell_{3}(-1,1))=0\\ (\sigma_{3}(\ell_{3}).\ell_{4})=(\ell_{3}(-1,-1).\ell_{4}(-1,-1))=1\\ (\sigma_{3}(\ell_{3}).\ell_{5})=(\ell_{3}(-1,-1).\ell_{4}(-1,1))=1\\ (\sigma_{3}(\ell_{3}).\ell_{6})=(\ell_{3}(-1,-1).L_{2})=0,\end{cases} (8)
{(σ3(4).1)=(4(1,1).1(1,1))=0(σ3(4).2)=(4(1,1).1(1,1))=0(σ3(4).3)=(4(1,1).3(1,1))=1(σ3(4).4)=(4(1,1).4(1,1))=0(σ3(4).5)=(4(1,1).4(1,1))=1(σ3(4).6)=(4(1,1).L2)=0,\begin{cases}(\sigma_{3}(\ell_{4}).\ell_{1})=(\ell_{4}(1,-1).\ell_{1}(1,1))=0\\ (\sigma_{3}(\ell_{4}).\ell_{2})=(\ell_{4}(1,-1).\ell_{1}(1,-1))=0\\ (\sigma_{3}(\ell_{4}).\ell_{3})=(\ell_{4}(1,-1).\ell_{3}(-1,1))=1\\ (\sigma_{3}(\ell_{4}).\ell_{4})=(\ell_{4}(1,-1).\ell_{4}(-1,-1))=0\\ (\sigma_{3}(\ell_{4}).\ell_{5})=(\ell_{4}(1,-1).\ell_{4}(-1,1))=1\\ (\sigma_{3}(\ell_{4}).\ell_{6})=(\ell_{4}(1,-1).L_{2})=0,\end{cases} (9)
{(σ3(5).1)=(4(1,1).1(1,1))=0(σ3(5).2)=(4(1,1).1(1,1))=0(σ3(5).3)=(4(1,1).3(1,1))=1(σ3(5).4)=(4(1,1).4(1,1))=1(σ3(5).5)=(4(1,1).4(1,1))=0(σ3(5).6)=(4(1,1).L2)=0.\begin{cases}(\sigma_{3}(\ell_{5}).\ell_{1})=(\ell_{4}(1,1).\ell_{1}(1,1))=0\\ (\sigma_{3}(\ell_{5}).\ell_{2})=(\ell_{4}(1,1).\ell_{1}(1,-1))=0\\ (\sigma_{3}(\ell_{5}).\ell_{3})=(\ell_{4}(1,1).\ell_{3}(-1,1))=1\\ (\sigma_{3}(\ell_{5}).\ell_{4})=(\ell_{4}(1,1).\ell_{4}(-1,-1))=1\\ (\sigma_{3}(\ell_{5}).\ell_{5})=(\ell_{4}(1,1).\ell_{4}(-1,1))=0\\ (\sigma_{3}(\ell_{5}).\ell_{6})=(\ell_{4}(1,1).L_{2})=0.\end{cases} (10)

Hence, we obtain

{σ1(1)=23σ1(2)=13σ1(3)=12σ3(3)=45σ3(4)=35σ3(5)=34\begin{cases}\sigma_{1}(\ell_{1})=\ell-\ell_{2}-\ell_{3}\\ \sigma_{1}(\ell_{2})=\ell-\ell_{1}-\ell_{3}\\ \sigma_{1}(\ell_{3})=\ell-\ell_{1}-\ell_{2}\\ \sigma_{3}(\ell_{3})=\ell-\ell_{4}-\ell_{5}\\ \sigma_{3}(\ell_{4})=\ell-\ell_{3}-\ell_{5}\\ \sigma_{3}(\ell_{5})=\ell-\ell_{3}-\ell_{4}\end{cases} (11)

in PicX¯\textup{Pic}\,\overline{X} by [Har77, Chapter V, Proposition 4.9]. As a result, we have

{σ1()=2123σ3()=2345,\begin{cases}\sigma_{1}(\ell)=2\ell-\ell_{1}-\ell_{2}-\ell_{3}\\ \sigma_{3}(\ell)=2\ell-\ell_{3}-\ell_{4}-\ell_{5},\end{cases} (12)

and clearly σ2()=σ4()=\sigma_{2}(\ell)=\sigma_{4}(\ell)=\ell. Then

{Ker(1+σ1)=123Ker(1+σ2)=12Ker(1+σ3)=345Ker(1+σ4)=45,\begin{cases}\textup{Ker}(1+\sigma_{1})=\langle\ell-\ell_{1}-\ell_{2}-\ell_{3}\rangle\\ \textup{Ker}(1+\sigma_{2})=\langle\ell_{1}-\ell_{2}\rangle\\ \textup{Ker}(1+\sigma_{3})=\langle\ell-\ell_{3}-\ell_{4}-\ell_{5}\rangle\\ \textup{Ker}(1+\sigma_{4})=\langle\ell_{4}-\ell_{5}\rangle,\\ \end{cases} (13)
{Ker(1σ1)=1,2,3,4,5,6Ker(1σ2)=,1+2,3,4,5,6Ker(1σ3)=3,4,5,1,2,6Ker(1σ4)=,1,2,3,4+5,6,\begin{cases}\textup{Ker}(1-\sigma_{1})=\langle\ell-\ell_{1},\ell-\ell_{2},\ell-\ell_{3},\ell_{4},\ell_{5},\ell_{6}\rangle\\ \textup{Ker}(1-\sigma_{2})=\langle\ell,\ell_{1}+\ell_{2},\ell_{3},\ell_{4},\ell_{5},\ell_{6}\rangle\\ \textup{Ker}(1-\sigma_{3})=\langle\ell-\ell_{3},\ell-\ell_{4},\ell-\ell_{5},\ell_{1},\ell_{2},\ell_{6}\rangle\\ \textup{Ker}(1-\sigma_{4})=\langle\ell,\ell_{1},\ell_{2},\ell_{3},\ell_{4}+\ell_{5},\ell_{6}\rangle,\\ \end{cases} (14)

and

{(1σ1)PicX¯=123(1σ2)PicX¯=12(1σ3)PicX¯=345(1σ4)PicX¯=45.\begin{cases}(1-\sigma_{1})\textup{Pic}\,\overline{X}=\langle\ell-\ell_{1}-\ell_{2}-\ell_{3}\rangle\\ (1-\sigma_{2})\textup{Pic}\,\overline{X}=\langle\ell_{1}-\ell_{2}\rangle\\ (1-\sigma_{3})\textup{Pic}\,\overline{X}=\langle\ell-\ell_{3}-\ell_{4}-\ell_{5}\rangle\\ (1-\sigma_{4})\textup{Pic}\,\overline{X}=\langle\ell_{4}-\ell_{5}\rangle.\\ \end{cases} (15)

Given a finite cyclic group G=σG=\langle\sigma\rangle and a GG-module MM, by [NSW15, Proposition 1.7.1], recall that we have isomorphisms H1(G,M)H^1(G,M)\textup{H}^{1}(G,M)\cong\hat{\textup{H}}^{-1}(G,M), where the latter group is the quotient of MNG\prescript{}{N_{G}}{M}, the set of elements of MM of norm 0, by its subgroup (1σ)M(1-\sigma)M.

By [NSW15, Proposition 1.6.7], we have

H1(K,PicX¯)=H1(G,PicX¯),\textup{H}^{1}(K,\textup{Pic}\,\overline{X})=\textup{H}^{1}(G,\textup{Pic}\,\overline{X}),

where G=σi,1i4G=\langle\sigma_{i},1\leq i\leq 4\rangle. Then one has the following (inflation-restriction) exact sequence

0H1(σ1,σ2,σ3,PicX¯σ4)H1(G,PicX¯)H1(σ4,PicX¯)=0,0\rightarrow\textup{H}^{1}(\langle\sigma_{1},\sigma_{2},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{4}\rangle})\rightarrow\textup{H}^{1}(G,\textup{Pic}\,\overline{X})\rightarrow\textup{H}^{1}(\langle\sigma_{4}\rangle,\textup{Pic}\,\overline{X})=0,

hence H1(G,PicX¯)H1(σ1,σ2,σ3,PicX¯σ4)\textup{H}^{1}(G,\textup{Pic}\,\overline{X})\cong\textup{H}^{1}(\langle\sigma_{1},\sigma_{2},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{4}\rangle}). Continuing as above, we have

0H1(σ1,σ3,PicX¯σ2,σ4)H1(σ1,σ2,σ3,PicX¯σ4)H1(σ2,PicX¯σ4)=0,0\rightarrow\textup{H}^{1}(\langle\sigma_{1},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{4}\rangle})\rightarrow\textup{H}^{1}(\langle\sigma_{1},\sigma_{2},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{4}\rangle})\rightarrow\textup{H}^{1}(\langle\sigma_{2}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{4}\rangle})=0,

hence H1(σ1,σ2,σ3,PicX¯σ4)H1(σ1,σ3,PicX¯σ2,σ4)\textup{H}^{1}(\langle\sigma_{1},\sigma_{2},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{4}\rangle})\cong\textup{H}^{1}(\langle\sigma_{1},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{4}\rangle}). Now we are left with

0H1(σ1,PicX¯σ2,σ3,σ4)H1(σ1,σ3,PicX¯σ2,σ4)H1(σ3,PicX¯σ2,σ4)=0,0\rightarrow\textup{H}^{1}(\langle\sigma_{1}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle})\rightarrow\textup{H}^{1}(\langle\sigma_{1},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{4}\rangle})\rightarrow\textup{H}^{1}(\langle\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{4}\rangle})=0,

hence H1(σ1,σ3,PicX¯σ2,σ4)H1(σ1,PicX¯σ2,σ3,σ4)=/2\textup{H}^{1}(\langle\sigma_{1},\sigma_{3}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{4}\rangle})\cong\textup{H}^{1}(\langle\sigma_{1}\rangle,\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle})=\mathbb{Z}/2\mathbb{Z}. Indeed, the last group can be computed as follows. We have

PicX¯σ2,σ3,σ4=1+2,3,245,6.\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle}=\langle\ell_{1}+\ell_{2},\ell-\ell_{3},2\ell-\ell_{4}-\ell_{5},\ell_{6}\rangle.

Considering the action of σ1\sigma_{1} on this invariant group, we have

PicNσ1X¯σ2,σ3,σ4=Ker(1+σ1)PicX¯σ2,σ3,σ4=123.\prescript{}{N_{\sigma_{1}}}{\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle}}=\textup{Ker}(1+\sigma_{1})\cap\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle}=\langle\ell-\ell_{1}-\ell_{2}-\ell_{3}\rangle.

On the other hand,

(1σ1)PicX¯σ2,σ3,σ4=[(1σ1)PicX¯]PicX¯σ2,σ3,σ4=2(123).(1-\sigma_{1})\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle}=[(1-\sigma_{1})\textup{Pic}\,\overline{X}]\cap\textup{Pic}\,\overline{X}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle}=\langle 2(\ell-\ell_{1}-\ell_{2}-\ell_{3})\rangle.

Given these results, we conclude that

H1(K,PicX¯)=H1(G,PicX¯)/2.\textup{H}^{1}(K,\textup{Pic}\,\overline{X})=\textup{H}^{1}(G,\textup{Pic}\,\overline{X})\cong\mathbb{Z}/2\mathbb{Z}.

Theorem 3.8.

Let KK be a number field. With the same notations as before, let k𝔸4(K)k\in\mathbb{A}^{4}(K) such that [E:K]=16[E:K]=16 where E=K(ki24,1i4)E=K(\sqrt{k_{i}^{2}-4},1\leq i\leq 4) and XX is smooth over KK. Let UU be the affine cubic surface defined by the equation

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where a,b,c,da,b,c,d are defined by (2)(2) for some k=(k1,k2,k3,k4)𝔸4(K)k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(K). Then we have

Br1U/Br0U/2\textup{Br}_{1}\,U/\textup{Br}_{0}\,U\cong\mathbb{Z}/2\mathbb{Z}

with a generator

𝒜\displaystyle\mathcal{A} =CorKF1(xk1k2+(k124)(k224)2,(k1k224+k2k124)2)\displaystyle=\textup{Cor}^{F_{1}}_{K}\left(x-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2},(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2}\right)
=CorKF3(yk1k4+(k124)(k424)2,(k1k424+k4k124)2)\displaystyle=\textup{Cor}^{F_{3}}_{K}\left(y-\frac{k_{1}k_{4}+\sqrt{(k_{1}^{2}-4)(k_{4}^{2}-4)}}{2},(k_{1}\sqrt{k_{4}^{2}-4}+k_{4}\sqrt{k_{1}^{2}-4})^{2}\right)
=CorKF2(zk1k3+(k124)(k324)2,(k1k324+k3k124)2),\displaystyle=\textup{Cor}^{F_{2}}_{K}\left(z-\frac{k_{1}k_{3}+\sqrt{(k_{1}^{2}-4)(k_{3}^{2}-4)}}{2},(k_{1}\sqrt{k_{3}^{2}-4}+k_{3}\sqrt{k_{1}^{2}-4})^{2}\right),

where Fi=K((k124)(ki+124))F_{i}=K(\sqrt{(k_{1}^{2}-4)(k_{i+1}^{2}-4)}) for 1i31\leq i\leq 3. Furthermore, we also have

BrX/Br0X=Br1X/Br0XBr1U/Br0U\textup{Br}\,X/\textup{Br}_{0}\,X=\textup{Br}_{1}\,X/\textup{Br}_{0}\,X\cong\textup{Br}_{1}\,U/\textup{Br}_{0}\,U

with a generator

𝒜0\displaystyle\mathcal{A}_{0} =CorKF1(xtk1k2+(k124)(k224)2,(k1k224+k2k124)2)\displaystyle=\textup{Cor}^{F_{1}}_{K}\left(\frac{x}{t}-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2},(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2}\right)
=CorKF3(ytk1k4+(k124)(k424)2,(k1k424+k4k124)2)\displaystyle=\textup{Cor}^{F_{3}}_{K}\left(\frac{y}{t}-\frac{k_{1}k_{4}+\sqrt{(k_{1}^{2}-4)(k_{4}^{2}-4)}}{2},(k_{1}\sqrt{k_{4}^{2}-4}+k_{4}\sqrt{k_{1}^{2}-4})^{2}\right)
=CorKF2(ztk1k3+(k124)(k324)2,(k1k324+k3k124)2)\displaystyle=\textup{Cor}^{F_{2}}_{K}\left(\frac{z}{t}-\frac{k_{1}k_{3}+\sqrt{(k_{1}^{2}-4)(k_{3}^{2}-4)}}{2},(k_{1}\sqrt{k_{3}^{2}-4}+k_{3}\sqrt{k_{1}^{2}-4})^{2}\right)

over t0t\not=0.

Proof.

We keep the notation as in the previous proposition. Then PicU¯\textup{Pic}\,\overline{U} is given by the following quotient group

PicU¯PicX¯/(i=13Li)(i=16i)/2Σi3i,36,6)i=14[i]\textup{Pic}\,\overline{U}\cong\textup{Pic}\,\overline{X}/(\oplus_{i=1}^{3}\mathbb{Z}L_{i})\cong(\oplus_{i=1}^{6}\mathbb{Z}\ell_{i}\oplus\mathbb{Z}\ell)/\langle 2\ell-\Sigma_{i\not=3}\ell_{i},\ell-\ell_{3}-\ell_{6},\ell_{6})\rangle\cong\oplus_{i=1}^{4}\mathbb{Z}[\ell_{i}]

by Proposition 3.1 and formula (4)(4). Here for any divisor DPicX¯D\in\textup{Pic}\,\overline{X}, denote by [D][D] its image in PicU¯\textup{Pic}\,\overline{U}. By Proposition 3.1, we also have K¯×=K¯[U]×\overline{K}^{\times}=\overline{K}[U]^{\times}. By the Hochschild–Serre spectral sequence, we have the following injective homomorphism

Br1U/Br0UH1(K,PicU¯),\textup{Br}_{1}\,U/\textup{Br}_{0}\,U\hookrightarrow\textup{H}^{1}(K,\textup{Pic}\,\overline{U}),

and in fact it is an isomorphism because over a number field KK, we have H3(K,𝔾m)=0\textup{H}^{3}(K,\mathbb{G}_{m})=0 from class field theory. Furthermore, the smooth compactification XX of UU has rational points, hence so does UU, which comes from the fact that any smooth cubic surface over an infinite field kk is unirational over kk as soon as it has a kk-point (see [Kol02]), so we also have Br0U=BrK\textup{Br}_{0}\,U=\textup{Br}\,K.

Since PicU¯\textup{Pic}\,\overline{U} is free and Gal(K¯/E)\textup{Gal}(\overline{K}/E) acts on PicU¯\textup{Pic}\,\overline{U} trivially, we obtain that H1(K,PicU¯)H1(G,PicU¯)\textup{H}^{1}(K,\textup{Pic}\,\overline{U})\cong\textup{H}^{1}(G,\textup{Pic}\,\overline{U}) by [NSW15, Proposition 1.6.7]. Now in PicU¯\textup{Pic}\,\overline{U}, as [6]=0[\ell_{6}]=0, []=[3][\ell]=[\ell_{3}] and 2[3]=[1]+[2]+[4]+[5]2[\ell_{3}]=[\ell_{1}]+[\ell_{2}]+[\ell_{4}]+[\ell_{5}], we have the following equalities

{σ1([1])=[2],σ1([2])=[1],σ1([3])=[3][1][2],σ1([4])=[4];σ2([1])=[2],σ2([2])=[1],σ2([3])=[3],σ2([4])=[4];σ3([1])=[1],σ3([2])=[2],σ3([3])=[1]+[2][3],σ3([4])=[1]+[2]+[4]2[3];σ4([1])=[1],σ4([2])=[2],σ4([3])=[3],σ4([4])=2[3][1][2][4].\begin{cases}\sigma_{1}([\ell_{1}])=-[\ell_{2}],\sigma_{1}([\ell_{2}])=-[\ell_{1}],\sigma_{1}([\ell_{3}])=[\ell_{3}]-[\ell_{1}]-[\ell_{2}],\sigma_{1}([\ell_{4}])=[\ell_{4}];\\ \sigma_{2}([\ell_{1}])=[\ell_{2}],\sigma_{2}([\ell_{2}])=[\ell_{1}],\sigma_{2}([\ell_{3}])=[\ell_{3}],\sigma_{2}([\ell_{4}])=[\ell_{4}];\\ \sigma_{3}([\ell_{1}])=[\ell_{1}],\sigma_{3}([\ell_{2}])=[\ell_{2}],\sigma_{3}([\ell_{3}])=[\ell_{1}]+[\ell_{2}]-[\ell_{3}],\sigma_{3}([\ell_{4}])=[\ell_{1}]+[\ell_{2}]+[\ell_{4}]-2[\ell_{3}];\\ \sigma_{4}([\ell_{1}])=[\ell_{1}],\sigma_{4}([\ell_{2}])=[\ell_{2}],\sigma_{4}([\ell_{3}])=[\ell_{3}],\sigma_{4}([\ell_{4}])=2[\ell_{3}]-[\ell_{1}]-[\ell_{2}]-[\ell_{4}].\end{cases} (16)

Using the inflation-restriction sequences similarly as in the previous proposition, with PicX¯\textup{Pic}\,\overline{X} replaced by PicU¯\textup{Pic}\,\overline{U}, we can compute that

H1(G,PicU¯)H1(σ1,PicU¯σ2,σ3,σ4)=H1(σ1,[1]+[2])=[1]+[2]2([1]+[2])/2.\textup{H}^{1}(G,\textup{Pic}\,\overline{U})\cong\textup{H}^{1}(\langle\sigma_{1}\rangle,\textup{Pic}\,\overline{U}^{\langle\sigma_{2},\sigma_{3},\sigma_{4}\rangle})=\textup{H}^{1}(\langle\sigma_{1}\rangle,\langle[\ell_{1}]+[\ell_{2}]\rangle)=\displaystyle\frac{\langle[\ell_{1}]+[\ell_{2}]\rangle}{\langle 2([\ell_{1}]+[\ell_{2}])\rangle}\cong\mathbb{Z}/2\mathbb{Z}.

Now we produce concrete generators in Br1U\textup{Br}_{1}\,U for Br1U/Br0U\textup{Br}_{1}\,U/\textup{Br}_{0}\,U. Then UU is defined by the equation

x2+y2+z2+xyz=ax+by+cz+d.x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d.

We will show that the following quaternion algebras in BrK(U)\textup{Br}\,K(U) are non-constant elements of Br1U\textup{Br}_{1}\,U, and hence they are equal in Br1U/Br0U\textup{Br}_{1}\,U/\textup{Br}_{0}\,U.

{𝒜1=CorKF1(xk1k2+(k124)(k224)2,(k1k224+k2k124)2),𝒜2=CorKF3(yk1k4+(k124)(k424)2,(k1k424+k4k124)2),𝒜3=CorKF2(zk1k3+(k124)(k324)2,(k1k324+k3k124)2).\begin{cases}\mathcal{A}_{1}=\textup{Cor}^{F_{1}}_{K}\left(x-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2},(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2}\right),\\ \mathcal{A}_{2}=\textup{Cor}^{F_{3}}_{K}\left(y-\frac{k_{1}k_{4}+\sqrt{(k_{1}^{2}-4)(k_{4}^{2}-4)}}{2},(k_{1}\sqrt{k_{4}^{2}-4}+k_{4}\sqrt{k_{1}^{2}-4})^{2}\right),\\ \mathcal{A}_{3}=\textup{Cor}^{F_{2}}_{K}\left(z-\frac{k_{1}k_{3}+\sqrt{(k_{1}^{2}-4)(k_{3}^{2}-4)}}{2},(k_{1}\sqrt{k_{3}^{2}-4}+k_{3}\sqrt{k_{1}^{2}-4})^{2}\right).\end{cases} (17)

Indeed, it suffices to prove the claim for 𝒜1\mathcal{A}_{1} and we only need to show that 𝒜1BrU\mathcal{A}_{1}\in\textup{Br}\,U, since its formula implies that 𝒜1\mathcal{A}_{1} becomes zero under the field extension KK(k124,k224)K\subset K(\sqrt{k_{1}^{2}-4},\sqrt{k_{2}^{2}-4}), i.e., it is algebraic. By Grothendieck’s purity theorem ([Poo17, Theorem 6.8.3]), for any smooth integral variety YY over a field LL of characteristic 0, we have the exact sequence

0BrYBrL(Y)DY(1)H1(L(D),/),0\rightarrow\textup{Br}\,Y\rightarrow\textup{Br}\,L(Y)\rightarrow\oplus_{D\in Y^{(1)}}\textup{H}^{1}(L(D),\mathbb{Q}/\mathbb{Z}),

where the last map is given by the residue along the codimension-one point DD. Therefore, to prove that our algebras come from a class in BrU\textup{Br}\,U, it suffices to show that all their residues are trivial. We will show that

𝒜1=(xk1k2+(k124)(k224)2,(k1k224+k2k124)2)BrUF1\mathcal{A}_{1}^{\prime}=\left(x-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2},(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2}\right)\in\textup{Br}\,U_{F_{1}}

so that its corestriction is a well-defined element over KK. From the data of the 27 lines in Lemma 3.6 and the formula of 𝒜1\mathcal{A}_{1}^{\prime}, any non-trivial residue of 𝒜1\mathcal{A}_{1}^{\prime} must occur along an irreducible component of the following divisor(s)

D1:x\displaystyle D_{1}:x =(k1k2+(k124)(k224))2,\displaystyle=\frac{(k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)})}{2},
y\displaystyle y =(k1±k124)(k2±k224)4zcb(k1±k124)(k2±k224)4±k1k224+k2k1242.\displaystyle=-\frac{(k_{1}\pm\sqrt{k_{1}^{2}-4})(k_{2}\pm\sqrt{k_{2}^{2}-4})}{4}z-\frac{c-b\frac{(k_{1}\pm\sqrt{k_{1}^{2}-4})(k_{2}\pm\sqrt{k_{2}^{2}-4})}{4}}{\pm\frac{k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4}}{2}}.

However, clearly in the function field of any such irreducible component, (k1k224+k2k124)2(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2} 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 𝒜1\mathcal{A}_{1}^{\prime} is unramified, and hence 𝒜1BrU\mathcal{A}_{1}\in\textup{Br}\,U. The residues of 𝒜1\mathcal{A}_{1} at the lines L1,L2,L3L_{1},L_{2},L_{3} which form the complement of UU in XX are easily seen to be trivial. One thus also has 𝒜1BrX\mathcal{A}_{1}\in\textup{Br}\,X. This element is non-constant by the Faddeev exact sequence ([CS21, Theorem 1.5.2]), since the residue of 𝒜1\mathcal{A}_{1}^{\prime} regarded as an element of BrF1(x)=BrF1(1)\textup{Br}\,F_{1}(x)=\textup{Br}\,F_{1}(\mathbb{P}^{1}) is nontrivial at the closed point (xk1k2+(k124)(k224)2)\left(x-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2}\right) of F11\mathbb{P}^{1}_{F_{1}}. Alternatively, by [CS21, Corollary 11.3.5] on the Brauer group of conic bundles, the element 𝒜1\mathcal{A}_{1} is indeed non-constant. Furthermore, it will contribute to the Brauer–Manin obstruction to strong approximation in the next section.

Finally, the fact that 𝒜1=𝒜2=𝒜3=𝒜\mathcal{A}_{1}=\mathcal{A}_{2}=\mathcal{A}_{3}=\mathcal{A} viewed as an element of Br1U/Br0U/2\textup{Br}_{1}\,U/\textup{Br}_{0}\,U\cong\mathbb{Z}/2\mathbb{Z} (by abuse of notation) and that UU is the open subset of XX defined by t0t\not=0 give the desired generator 𝒜0\mathcal{A}_{0} of BrX/Br0X/2\textup{Br}\,X/\textup{Br}_{0}\,X\cong\mathbb{Z}/2\mathbb{Z}. ∎

3.3 The transcendental Brauer group

We begin with a specific assumption.

Assumption 3.3.

Let k=(k1,k2,k3,k4)𝔸4()k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(\mathbb{Z}) and E:=(ki24,1i4)E:=\mathbb{Q}(\sqrt{k_{i}^{2}-4},1\leq i\leq 4) such that [E:]=16[E:\mathbb{Q}]=16. For all 1ij41\leq i\not=j\leq 4, assume that |ki|3|k_{i}|\geq 3 such that (ki+ki24)(kj+kj24)(k_{i}+\sqrt{k_{i}^{2}-4})(k_{j}+\sqrt{k_{j}^{2}-4}) is not a square in EE.

Now we compute the transcendental Brauer group in our particular case.

Proposition 3.9.

Let k=(k1,k2,k3,k4)𝔸4()k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{A}^{4}(\mathbb{Z}) satisfy Assumption 3.3. Let UU be the affine cubic surface over \mathbb{Q} defined by

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where a,b,c,da,b,c,d are defined by (2)(2). Assume that its natural compactification X3X\subset\mathbb{P}^{3} is smooth over \mathbb{Q}. Set XE:=X×EX_{E}:=X\times_{\mathbb{Q}}E and UE:=U×EU_{E}:=U\times_{\mathbb{Q}}E. Then the natural map BrXEBrUE\textup{Br}\,X_{E}\rightarrow\textup{Br}\,U_{E} is an isomorphism. Moreover, UU has trivial transcendental Brauer group over \mathbb{Q}.

Proof.

The proof is inspired by that of [LM20, Proposition 4.1]. Let BrUE\mathcal{B}\in\textup{Br}\,U_{E} be a non-constant Brauer element. Multiplying \mathcal{B} by a constant algebra if necessary, by Proposition 3.3 and Theorem 3.8, we may assume that \mathcal{B} has order dividing 44 (note that under our assumption of kk, the field extension EE is totally real and thus contains no nontrivial roots of unity). In order to show that BrXE\mathcal{B}\in\textup{Br}\,X_{E}, we only need to show that \mathcal{B} is unramified along the three lines LiL_{i} on XX by Grothendieck’s purity theorem (see [CS21, Theorem 3.7.2]).

Let L=L1L=L_{1} and C=L2L3C=L_{2}\cup L_{3}. Let L=L\CL^{\prime}=L\,\backslash\,C. Note that LL meets CC at two rational points, so LL^{\prime} is non-canonically isomorphic to 𝔾m\mathbb{G}_{m}. Let the point (x:y:z:t)=(0:1:1:0)L(x:y:z:t)=(0:1:1:0)\in L^{\prime} be the identity element of the group law. Then an isomorphism with 𝔾m\mathbb{G}_{m} is realized via the following homomorphism:

𝔾mX,u(0:u:1:0).\mathbb{G}_{m}\rightarrow X,\hskip 28.45274ptu\mapsto(0:u:1:0). (18)

The residue of \mathcal{B} along LL lies inside H1(L,/2)\textup{H}^{1}(L^{\prime},\mathbb{Z}/2\mathbb{Z}). Assume by contradiction that the residue is nontrivial. Since the order of \mathcal{B} is a power of 22 dividing 44, then we can assume that the residue has order 22 (up to replacing \mathcal{B} by 22\mathcal{B}). This means that the residue corresponds to some irreducible degree 22 finite étale cover f:L′′Lf:L^{\prime\prime}\rightarrow L^{\prime}.

Over the field EE, the conic fiber CC over the coordinate (x=k1k2+(k124)(k224)2t)\left(x=\displaystyle\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2}t\right) is split, i.e. a union of two lines over EE. These lines meet LL at the points

Q±=(0:(k1±k124)(k2±k224)4:1:0).Q_{\pm}=(0:\displaystyle\frac{(k_{1}\pm\sqrt{k_{1}^{2}-4})(k_{2}\pm\sqrt{k_{2}^{2}-4})}{4}:1:0).

Let C+C_{+} be the irreducible component of CC containing Q+Q_{+}, i.e., C+=1(1,1)C_{+}=\ell_{1}(1,1). Consider the restriction of \mathcal{B} to C+C_{+}. This is well-defined outside of Q+Q_{+}, and since C+\Q+𝔸1C_{+}\,\backslash\,Q_{+}\simeq\mathbb{A}^{1} has constant Brauer group, \mathcal{B} actually extends to all of C+C_{+}. As C+C_{+} meets LL transversely, by the functoriality of residues ([CS21, Section 3.7]) we deduce that the residue of \mathcal{B} at Q+Q_{+} is also trivial, so the fiber f1(Q+)f^{-1}(Q_{+}) consists of exactly two rational points. This implies that L′′L^{\prime\prime} is geometrically irreducible, hence L′′𝔾mL^{\prime\prime}\cong\mathbb{G}_{m} non-canonically.

Now by choosing a rational point over Q+Q_{+} and using the above group homomorphism, we may therefore identify the degree 22 cover L′′LL^{\prime\prime}\rightarrow L^{\prime} with the map

𝔾mX,u(0:u2:1:0).\mathbb{G}_{m}\rightarrow X,\hskip 28.45274ptu\mapsto(0:u^{2}:1:0). (19)

However, our assumptions on kk imply that (k1+k124)(k2+k224)4\displaystyle\frac{(k_{1}+\sqrt{k_{1}^{2}-4})(k_{2}+\sqrt{k_{2}^{2}-4})}{4} is not a square in E×E^{\times}, which gives a contradiction. Thus the residue of \mathcal{B} along LL is trivial, and the same holds for the other lines. We conclude that \mathcal{B} is everywhere unramified, hence BrXE\mathcal{B}\in\textup{Br}\,X_{E}.
 
Now let BBrUB\in\textup{Br}\,U be a non-constant element. Then over the field extension EE, the corresponding image of BB comes from BrXE\textup{Br}\,X_{E} by the above argument. As BrX¯=0\textup{Br}\,\overline{X}=0, it is clear that BB is algebraic. The result follows. ∎

Remark 3.4.

Note that (BrU¯)Gal(¯/)=/2(\textup{Br}\,U_{\overline{\mathbb{Q}}})^{\textup{Gal}(\overline{\mathbb{Q}}/\mathbb{Q})}=\mathbb{Z}/2\mathbb{Z} by Proposition 3.3. However, in the above proposition, the Galois invariant element of order 22 does not descend to a Brauer group element over \mathbb{Q}, 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 vv of \mathbb{Q} there is a pairing

U(v)×BrU/U(\mathbb{Q}_{v})\times\textup{Br}\,U\rightarrow\mathbb{Q}/\mathbb{Z}

coming from the local invariant map

invv:Brv/\textup{inv}_{v}:\textup{Br}\,\mathbb{Q}_{v}\rightarrow\mathbb{Q}/\mathbb{Z}

from local class field theory (this is an isomorphism if vv is a prime number). This pairing is locally constant on the left by [Poo17, Proposition 8.2.9]. For integral points, any element αBrU\alpha\in\textup{Br}\,U pairs trivially on 𝒰(p)\mathcal{U}(\mathbb{Z}_{p}) for almost all primes pp, so we obtain a pairing U(A)×BrU/U(\textbf{{A}}_{\mathbb{Q}})\times\textup{Br}\,U\rightarrow\mathbb{Q}/\mathbb{Z}. As the local pairings are locally constant, we obtain a well-defined pairing

𝒰(A)×BrU/.\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}\times\textup{Br}\,U\rightarrow\mathbb{Q}/\mathbb{Z}.

For BBrUB\subseteq\textup{Br}\,U, let 𝒰(A)B\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{B} be the left kernel with respect to BB, and let 𝒰(A)Br=𝒰(A)BrU\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{\textup{Br}}=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{\textup{Br}\,U}. By abuse of notation, from now on we write the reduced Brauer–Manin set 𝒰(A)B\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})_{\bullet}^{B} in the standard way as 𝒰(A)B\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{B}. Note that the set 𝒰(A)B\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{B} depends only on the image of BB in the quotient BrU/Br0U\textup{Br}\,U/\textup{Br}_{0}\,U. By Theorem 3.8, the map 𝒜Br1U/Br\langle\mathcal{A}\rangle\rightarrow\textup{Br}_{1}\,U/\textup{Br}\,\mathbb{Q} is an isomorphism, hence 𝒰(A)Br1=𝒰(A)Br1U=𝒰(A)𝒜\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}\,U}=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\mathcal{A}}. We have the inclusions 𝒰()𝒰(A)𝒜𝒰(A)\mathcal{U}(\mathbb{Z})\subseteq\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\mathcal{A}}\subseteq\mathcal{U}(\textbf{{A}}_{\mathbb{Z}}), so that 𝒜\mathcal{A} can obstruct the integral Hasse principle or strong approximation on 𝒰\mathcal{U}.

Let VV be dense Zariski open in UU. As UU is smooth, the set V(v)V(\mathbb{Q}_{v}) is dense in U(v)U(\mathbb{Q}_{v}) for all places vv. Moreover, 𝒰(p)\mathcal{U}(\mathbb{Z}_{p}) is open in U(p)U(\mathbb{Q}_{p}), hence V(p)𝒰(p)V(\mathbb{Q}_{p})\cap\mathcal{U}(\mathbb{Z}_{p}) is dense in 𝒰(p)\mathcal{U}(\mathbb{Z}_{p}). As the local pairings are locally constant, we may restrict our attention to VV to calculate the local invariants of a given element in BrU\textup{Br}\,U. In particular, here we take the open subset VV given by ((k124)(k224)(2xk1k2)2)((k124)(k424)(2yk1k4)2)((k124)(k324)(2zk1k3)2)0((k_{1}^{2}-4)(k_{2}^{2}-4)-(2x-k_{1}k_{2})^{2})((k_{1}^{2}-4)(k_{4}^{2}-4)-(2y-k_{1}k_{4})^{2})((k_{1}^{2}-4)(k_{3}^{2}-4)-(2z-k_{1}k_{3})^{2})\not=0.

4.2 Brauer–Manin obstruction from a quaternion algebra

Now we consider the Markoff-type cubic surfaces UU over \mathbb{Q} and their integral models 𝒰\mathcal{U} over \mathbb{Z} defined by the equation (1)(1):

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where a,b,c,da,b,c,d\in\mathbb{Z} are defined as in (2)(2) with k4k\in\mathbb{Z}^{4}. Set f:=x2+y2+z2+xyzaxbyczd[x,y,z]f:=x^{2}+y^{2}+z^{2}+xyz-ax-by-cz-d\in\mathbb{Z}[x,y,z]. First of all, we study the existence of local integral points on those affine cubic surfaces given by f=0f=0.

Proposition 4.1 (Assumption A).

If k=(k1,k2,k3,k4)(\[2,2])4k=(k_{1},k_{2},k_{3},k_{4})\in(\mathbb{Z}\,\backslash\,[-2,2])^{4} satisfies k11k_{1}\equiv-1 mod 1616, ki5k_{i}\equiv 5 mod 1616 and k11k_{1}\equiv 1 mod 99, ki5k_{i}\equiv 5 mod 99 for 2i42\leq i\leq 4, such that (ki,kj)=1(k_{i},k_{j})=1, (ki24,kj24)=3(k_{i}^{2}-4,k_{j}^{2}-4)=3 for 1ij41\leq i\not=j\leq 4, and (k122,k222,k322,k422)=1(k_{1}^{2}-2,k_{2}^{2}-2,k_{3}^{2}-2,k_{4}^{2}-2)=1, then 𝒰(A)\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})\not=\emptyset.

Proof.

With our specific choice of kk in the assumption, we obtain:

  1. (1)

    Prime powers of p=2p=2: The only solution modulo 22 is the singular (1,0,0)(1,0,0) (up to permutation). However, we find a solution (1,0,0)(1,0,0) modulo 88 with 2x+yz22x+yz\equiv 2 mod 88, so twice the order valuation at 22 of the partial derivative at xx is less than the order valuation at 22 of f(x,y,z)f(x,y,z). This solution then lifts to a 22-adic integer solution by Hensel’s lemma (fixing the variables y,zy,z).

  2. (2)

    Prime powers of p=3p=3: We find the non-singular solution (1,0,0)(1,0,0), which lifts to a 33-adic integer solution by Hensel’s lemma (fixing the variables y,zy,z).

  3. (3)

    Prime powers of p5p\geq 5: We would like to find a non-singular solution modulo pp of the equations f=0f=0 which does not satisfy simultaneously

    df=0:2x+yz=a,2y+xz=b,2z+xy=c.df=0:2x+yz=a,2y+xz=b,2z+xy=c.

    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 f=0f=0 always has a solution whose one coordinate is 0: indeed, take z=0z=0, then f=0f=0 is equivalent to (2xa)2+(2yb)2=(k12+k324)(k22+k424)(2x-a)^{2}+(2y-b)^{2}=(k_{1}^{2}+k_{3}^{2}-4)(k_{2}^{2}+k_{4}^{2}-4), and every element in 𝔽p\mathbb{F}_{p} can be expressed as a sum of two squares. Now assume that all such points in U(𝔽p)U(\mathbb{F}_{p}) with one coordinate equal to 0 are singular. Then modulo pp, if z=0z=0 the required equations become f=0,(2x=a,2y=b,xy=c)f=0,(2x=a,2y=b,xy=c), plus all the permutations for x=0x=0 and y=0y=0. For convenience, we drop the phrase “modulo pp”. From these equations, we get

    (k12+k324)(k22+k424)=0,ab=4c(=4xy)(k_{1}^{2}+k_{3}^{2}-4)(k_{2}^{2}+k_{4}^{2}-4)=0,ab=4c\,(=4xy) (20)

    plus all the permutations, respectively

    (k12+k224)(k32+k424)=0,bc=4a(k_{1}^{2}+k_{2}^{2}-4)(k_{3}^{2}+k_{4}^{2}-4)=0,bc=4a (21)

    and

    (k12+k424)(k22+k324)=0,ac=4b.(k_{1}^{2}+k_{4}^{2}-4)(k_{2}^{2}+k_{3}^{2}-4)=0,ac=4b. (22)

    We will choose kk which does not satisfy all of these equations simultaneously to get a contradiction to our assumption at the beginning.
     
    Now if kk satisfies all the above equations, we first require that (ki24,kj24)=3(k_{i}^{2}-4,k_{j}^{2}-4)=3 for any 1ij41\leq i\not=j\leq 4. Next, without loss of generality (WLOG), assume in (22)(22) that k12+k424=0k_{1}^{2}+k_{4}^{2}-4=0 with ac=4b=(k12+k42)bac=4b=(k_{1}^{2}+k_{4}^{2})b, then putting the formulas of a,b,ca,b,c in ac(k12+k42)b=0ac-(k_{1}^{2}+k_{4}^{2})b=0 gives us k1k4(k22+k32k12k42)=0k_{1}k_{4}(k_{2}^{2}+k_{3}^{2}-k_{1}^{2}-k_{4}^{2})=0, hence either k22+k324=0k_{2}^{2}+k_{3}^{2}-4=0 or k1k4=0k_{1}k_{4}=0. (We also have similar equations for all the other cases.)

    If k22+k324=0k_{2}^{2}+k_{3}^{2}-4=0, then i=14ki28=0\sum_{i=1}^{4}k_{i}^{2}-8=0, and so from (20),(21)(20),(21) we obtain ki2+kj24=0k_{i}^{2}+k_{j}^{2}-4=0 for any iji\not=j, hence ki22=0k_{i}^{2}-2=0 for all 1i41\leq i\leq 4. Otherwise, if k22+k3240k_{2}^{2}+k_{3}^{2}-4\not=0 but k1k4=0k_{1}k_{4}=0, WLOG assume in (22)(22) that k1=0k_{1}=0, then k424=0k_{4}^{2}-4=0 and so from (20),(21)(20),(21) we have the following possibilities: k224=k324=0k_{2}^{2}-4=k_{3}^{2}-4=0, or k2=k3=0k_{2}=k_{3}=0. Therefore, we immediately deduce a sufficient condition for the nonexistence of singular solutions modulo p5p\geq 5: (ki,kj)=1(k_{i},k_{j})=1, (ki24,kj24)=3(k_{i}^{2}-4,k_{j}^{2}-4)=3 for any 1ij41\leq i\not=j\leq 4, and (k122,k222,k322,k422)=1(k_{1}^{2}-2,k_{2}^{2}-2,k_{3}^{2}-2,k_{4}^{2}-2)=1.

    As a result, assuming the hypothesis of the proposition, it is clear that U(𝔽p)U(\mathbb{F}_{p}) has a smooth point, which then lifts to a pp-adic integral point by Hensel’s lemma (with respect to the variable at which the partial derivative is nonzero modulo pp, fixing other variables).

We keep Assumption A to ensure that 𝒰(A)\mathcal{U}(\textbf{A}_{\mathbb{Z}})\not=\emptyset and study the Brauer–Manin obstruction to the existence of integral points. Note that our specific choice of kk implies that [E:]=16[E:\mathbb{Q}]=16. Indeed, it is clear that with our assumption, ki24>0k_{i}^{2}-4>0 is not a square in \mathbb{Q} for 1i41\leq i\leq 4. Then [E:]<16[E:\mathbb{Q}]<16 if and only if there are some iji\not=j such that (ki24)(kj24)(k_{i}^{2}-4)(k_{j}^{2}-4) is a square in \mathbb{Q}; however, since (ki24,kj24)=3(k_{i}^{2}-4,k_{j}^{2}-4)=3 and ki24=3l2k_{i}^{2}-4=3l^{2} does not have any solution modulo 88 with kik_{i} odd, that cannot happen.

Now let us calculate the local invariants of the following quaternion algebras as elements of the algebraic Brauer group Br1U\textup{Br}_{1}\,U:

{𝒜1=CorF1(xk1k2+(k124)(k224)2,(k1k224+k2k124)2),𝒜2=CorF3(yk1k4+(k124)(k444)2,(k1k424+k4k124)2),𝒜3=CorF2(zk1k3+(k124)(k324)2,(k1k324+k3k124)2),\begin{cases*}\mathcal{A}_{1}=\textup{Cor}^{F_{1}}_{\mathbb{Q}}\left(x-\frac{k_{1}k_{2}+\sqrt{(k_{1}^{2}-4)(k_{2}^{2}-4)}}{2},(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4})^{2}\right),\\ \mathcal{A}_{2}=\textup{Cor}^{F_{3}}_{\mathbb{Q}}\left(y-\frac{k_{1}k_{4}+\sqrt{(k_{1}^{2}-4)(k_{4}^{4}-4)}}{2},(k_{1}\sqrt{k_{4}^{2}-4}+k_{4}\sqrt{k_{1}^{2}-4})^{2}\right),\\ \mathcal{A}_{3}=\textup{Cor}^{F_{2}}_{\mathbb{Q}}\left(z-\frac{k_{1}k_{3}+\sqrt{(k_{1}^{2}-4)(k_{3}^{2}-4)}}{2},(k_{1}\sqrt{k_{3}^{2}-4}+k_{3}\sqrt{k_{1}^{2}-4})^{2}\right),\end{cases*}

where Fi=((k124)(ki+124))F_{i}=\mathbb{Q}(\sqrt{(k_{1}^{2}-4)(k_{i+1}^{2}-4)}) for 1i31\leq i\leq 3. Now for each ii, we have the local invariant map

invp𝒜i:U(p)/2,xinvp𝒜i(x).\textup{inv}_{p}\,\mathcal{A}_{i}:U(\mathbb{Q}_{p})\rightarrow\mathbb{Z}/2\mathbb{Z},\hskip 28.45274ptx\mapsto\textup{inv}_{p}\,\mathcal{A}_{i}(x).
Lemma 4.2 (Assumption B).

Let k=(k1,k2,k3,k4)4k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{Z}^{4} and p11p\geq 11 be a prime such that (2,p)p=1/2(2,p)_{p}=1/2 and k12pk_{1}-2\equiv p mod p3p^{3}. Then there exist k2,k3,k4k_{2},k_{3},k_{4} such that k224k_{2}^{2}-4 is a quadratic nonresidue mod pp, k3k40,2k_{3}\equiv k_{4}\not\equiv 0,2 mod pp, k222k3k_{2}-2\not\equiv 2k_{3} mod pp, and k2,k3,k4k_{2},k_{3},k_{4} satisfy the cubic equation defining an affine Markoff surface SS:

k22+k32+k42k2k3k4(p+2)2modp2.k_{2}^{2}+k_{3}^{2}+k_{4}^{2}-k_{2}k_{3}k_{4}\equiv(p+2)^{2}\hskip 7.11317pt\textup{mod}\,p^{2}.

Furthermore, for any kk satisfying all the above conditions (Assumption B) and Assumption A, the local invariant map of the quaternion algebra 𝒜1\mathcal{A}_{1} at pp is surjective.

Proof.

For convenience (and by abuse of notation), we will also write 𝒜1\mathcal{A}_{1} (resp. F1F_{1}) as 𝒜\mathcal{A} (resp. FF). We denote (k1k224+k2k124)(k_{1}\sqrt{k_{2}^{2}-4}+k_{2}\sqrt{k_{1}^{2}-4}) by α1\alpha_{1} and write it simply as α\alpha. Set D:=(k124)(k224)D:=(k_{1}^{2}-4)(k_{2}^{2}-4).

First of all, over F=(D)F=\mathbb{Q}(\sqrt{D}) the affine cubic equation of 𝒰\mathcal{U} can be rewritten equivalently as

f(x,y,z,k1,k2,k3,k4)=\displaystyle f(x,y,z,k_{1},k_{2},k_{3},k_{4})= (23)
(xk1k2+D2)(xk1k2D2k3k4+yz)\displaystyle\left(x-\frac{k_{1}k_{2}+\sqrt{D}}{2}\right)\left(x-\frac{k_{1}k_{2}-\sqrt{D}}{2}-k_{3}k_{4}+yz\right) +(y+k1k2+D4zb2)2\displaystyle+\left(y+\frac{k_{1}k_{2}+\sqrt{D}}{4}z-\frac{b}{2}\right)^{2}
α216(z2b(k1k2+D)8cα2)2=0,\displaystyle-\frac{\alpha^{2}}{16}\left(z-\frac{2b(k_{1}k_{2}+\sqrt{D})-8c}{\alpha^{2}}\right)^{2}=0,

for all (k1,k2,k3,k4)4(k_{1},k_{2},k_{3},k_{4})\in\mathbb{Z}^{4} satisfying our hypothesis. Since vp(D)=1\textup{v}_{p}(D)=1, pp ramifies over F=(D)F=\mathbb{Q}(\sqrt{D}), i.e., p𝒪F=𝔭2p\mathcal{O}_{F}=\mathfrak{p}^{2} where 𝔭\mathfrak{p} is a nonzero prime ideal of 𝒪F\mathcal{O}_{F}. Therefore, we have N𝔭=p\textup{N}\mathfrak{p}=p and

𝒪F/𝔭𝔽p=/p.\mathcal{O}_{F}/\mathfrak{p}\cong\mathbb{F}_{p}=\mathbb{Z}/p\mathbb{Z}.

Following the proof of [LM20, Proposition 5.5], for all u2,u3,u4𝔽pu_{2},u_{3},u_{4}\in\mathbb{F}_{p}^{*}, there exists an 𝔽p\mathbb{F}_{p}-point on the variety

(2k2k3k4)2=(k324)(k424),k22=u2v22,k32=u3v32,k42=u4v42(2k_{2}-k_{3}k_{4})^{2}=(k_{3}^{2}-4)(k_{4}^{2}-4),k_{2}-2=u_{2}v_{2}^{2},k_{3}-2=u_{3}v_{3}^{2},k_{4}-2=u_{4}v_{4}^{2}

which satisfy v2v3v40v_{2}v_{3}v_{4}\not=0. As (k22)(k32)(k42)0(k_{2}-2)(k_{3}-2)(k_{4}-2)\not=0, we know from [LM20, Lemma 5.3] that this gives rise to a smooth 𝔽p\mathbb{F}_{p}-point of SS, hence a /p2\mathbb{Z}/p^{2}\mathbb{Z}-point with the same residue modulo pp by Hensel’s lemma. Now to construct the given 𝔽p\mathbb{F}_{p}-point, we restrict our attention to the subvariety given by k3=k4k_{3}=k_{4} mod pp then assume that u3=u4u_{3}=u_{4} mod pp and u2u_{2} is a quadratic nonresidue mod pp. The above equations then become

(2k2k3k4)2=(k324)2,k22=u2v22,k32=u3v32.(2k_{2}-k_{3}k_{4})^{2}=(k_{3}^{2}-4)^{2},k_{2}-2=u_{2}v_{2}^{2},k_{3}-2=u_{3}v_{3}^{2}.

Factoring the left hand side, it suffices to solve the equations

k2=k322,k22=u2v22,k32=u3v32.k_{2}=k_{3}^{2}-2,k_{2}-2=u_{2}v_{2}^{2},k_{3}-2=u_{3}v_{3}^{2}.

This then gives the equation of an affine curve

u2v22=u32v34+4u3v32.u_{2}v_{2}^{2}=u_{3}^{2}v_{3}^{4}+4u_{3}v_{3}^{2}.

By the argument in the proof of [LM20, Proposition 5.5], this affine curve has a unique singular point (v2,v3)=(0,0)(v_{2},v_{3})=(0,0) and has p2p-2 many 𝔽p\mathbb{F}_{p}-points. Of these points at most 33 satisfy v2v3=0v_{2}v_{3}=0, and k3=0k_{3}=0 gives only at most 44 points, hence providing p234=p9>0p-2-3-4=p-9>0, there exists an 𝔽p\mathbb{F}_{p}-point (k2,k3,k4)(k_{2},k_{3},k_{4}) with the properties:

(k22)(k32)(k42)0,k3=k40.(k_{2}-2)(k_{3}-2)(k_{4}-2)\not=0,k_{3}=k_{4}\not=0.

Since k30k_{3}\not=0, we have k2=k3222k_{2}=k_{3}^{2}-2\not=-2, so k2240k_{2}^{2}-4\not=0 in 𝔽p\mathbb{F}_{p}. Moreover, as u2u_{2} is a quadratic nonresidue modulo pp, so is k22=k324k_{2}-2=k_{3}^{2}-4. Since k2+2=k32k_{2}+2=k_{3}^{2} is a nonzero square in 𝔽p\mathbb{F}_{p}, we deduce that k224k_{2}^{2}-4 is a quadratic nonresidue modulo pp, as required.

Now after lifting from the smooth 𝔽p\mathbb{F}_{p}-point to a /p2\mathbb{Z}/p^{2}\mathbb{Z}-point (with the same residue modulo pp), if k222k3k_{2}-2\equiv 2k_{3} mod pp then we can take k2=k2k_{2}^{\prime}=k_{2}, k3=k3k_{3}^{\prime}=-k_{3} and k4=k4k_{4}^{\prime}=-k_{4} (which still satisfy the cubic equation mod p2p^{2}) to have k222k3k^{\prime}_{2}-2\not\equiv 2k_{3}^{\prime} mod pp since k30k_{3}\not\equiv 0 mod pp and pp is odd. Therefore, there exist integers k2,k3,k4k_{2},k_{3},k_{4} satisfying the required properties by the Chinese Remainder Theorem.
 
Now assume that k=(k1,k2,k3,k4)k=(k_{1},k_{2},k_{3},k_{4}) satisfies all the above congruence conditions and Assumption A. Since p𝒪F=𝔭2p\mathcal{O}_{F}=\mathfrak{p}^{2} and 𝒪F/𝔭/p\mathcal{O}_{F}/\mathfrak{p}\cong\mathbb{Z}/p\mathbb{Z}, by Hensel’s lemma we deduce that when k224k_{2}^{2}-4 is a quadratic nonresidue modulo pp over \mathbb{Z}, then k224k_{2}^{2}-4 is a quadratic nonresidue modulo 𝔭\mathfrak{p} over 𝒪F\mathcal{O}_{F}. Recalling that k12pk_{1}-2\equiv p mod p3p^{3}, we can choose a /p3\mathbb{Z}/p^{3}\mathbb{Z}-point (x,y,z)(x,y,z) of 𝒰\mathcal{U} such that

xk2=p,yk4=p,zk3=p,x-k_{2}=p,y-k_{4}=p,z-k_{3}=p,

from which we have (in /p3\mathbb{Z}/p^{3}\mathbb{Z}):

f(x,y,z,k1,k2,k3,k4)=f(x,y,z,2,k2,k3,k4)i{2,3,4}pki(ki+p)+4p+p2+pk2k3k4\displaystyle f(x,y,z,k_{1},k_{2},k_{3},k_{4})=f(x,y,z,2,k_{2},k_{3},k_{4})-\sum_{i\in\{2,3,4\}}pk_{i}(k_{i}+p)+4p+p^{2}+pk_{2}k_{3}k_{4}
=(xk2)(xk2+yzk3k4)+(y+k2z2b2)2k2244(zk3)2\displaystyle=(x-k_{2})(x-k_{2}+yz-k_{3}k_{4})+\left(y+\frac{k_{2}z}{2}-\frac{b}{2}\right)^{2}-\frac{k_{2}^{2}-4}{4}(z-k_{3})^{2}
i{2,3,4}pki(ki+p)+4p+p2+pk2k3k4\displaystyle-\sum_{i\in\{2,3,4\}}pk_{i}(k_{i}+p)+4p+p^{2}+pk_{2}k_{3}k_{4}
=p(p+p2+p(k3+k4))+p2(1+k22)2p2(k2244)i{2,3,4}pki(ki+p)+4p+p2+pk2k3k4\displaystyle=p(p+p^{2}+p(k_{3}+k_{4}))+p^{2}\left(1+\frac{k_{2}}{2}\right)^{2}-p^{2}\left(\frac{k_{2}^{2}-4}{4}\right)-\sum_{i\in\{2,3,4\}}pk_{i}(k_{i}+p)+4p+p^{2}+pk_{2}k_{3}k_{4}
=p(p+p2+p(k3+k4))+p(2p+pk2)i{2,3,4}p(ki2+pki)+p(4+p+k2k3k4)\displaystyle=p(p+p^{2}+p(k_{3}+k_{4}))+p(2p+pk_{2})-\sum_{i\in\{2,3,4\}}p(k_{i}^{2}+pk_{i})+p(4+p+k_{2}k_{3}k_{4})
=p(4+4p+p2(k22+k32+k42)+k2k3k4)\displaystyle=p(4+4p+p^{2}-(k_{2}^{2}+k_{3}^{2}+k_{4}^{2})+k_{2}k_{3}k_{4})
=p((p+2)2(k22+k32+k42k2k3k4))=0,\displaystyle=p((p+2)^{2}-(k_{2}^{2}+k_{3}^{2}+k_{4}^{2}-k_{2}k_{3}k_{4}))=0,

and

fx(x,y,z)=2x+yzk1k2k3k4=2(xk2)pk2+p2+p(k3+k4)=p(p+2k2+k3+k4).f^{\prime}_{x}(x,y,z)=2x+yz-k_{1}k_{2}-k_{3}k_{4}=2(x-k_{2})-pk_{2}+p^{2}+p(k_{3}+k_{4})=p(p+2-k_{2}+k_{3}+k_{4}).

Therefore vp(f)3>2=2vp(fx)\textup{v}_{p}(f)\geq 3>2=2\textup{v}_{p}(f^{\prime}_{x}), by Hensel’s lemma (fixing the variables y,zy,z), this point lifts to a p\mathbb{Z}_{p}-point (x,y,z)(x,y,z) (abuse of notation) with the same residue modulo p2p^{2}. Then since v𝔭(D)=vp(D)=1\textup{v}_{\mathfrak{p}}(\sqrt{D})=\textup{v}_{p}(D)=1, the local invariant at pp of 𝒜\mathcal{A} at this point is equal to

(xk1k2+D2,α2)𝔭=(xk1k2+D2,k224)𝔭=1/2\left(x-\frac{k_{1}k_{2}+\sqrt{D}}{2},\alpha^{2}\right)_{\mathfrak{p}}=\left(x-\frac{k_{1}k_{2}+\sqrt{D}}{2},k_{2}^{2}-4\right)_{\mathfrak{p}}=1/2

by the formulae in [Neu13, Proposition II.1.4 and Proposition III.3.3].

It is also clear that there exists a p\mathbb{Z}_{p}-point such that xk2x-k_{2} is not divisible by pp, which gives the local invariant of 𝒜\mathcal{A} at pp the value 0. Indeed, if every p\mathbb{Z}_{p}-point satisfies that xk2x-k_{2} is divisible by pp, then as k120k_{1}-2\equiv 0 mod pp, the affine equation of 𝒰\mathcal{U} over 𝔽p\mathbb{F}_{p} becomes

(xk2)(xk2+yzk3k4)+(y+k2z2b2)2k2244(zk3)2=0.\displaystyle(x-k_{2})(x-k_{2}+yz-k_{3}k_{4})+\left(y+\frac{k_{2}z}{2}-\frac{b}{2}\right)^{2}-\frac{k_{2}^{2}-4}{4}(z-k_{3})^{2}=0.

From the fact that k224k_{2}^{2}-4 is a quadratic nonresidue mod pp, one must obtain that (in 𝔽p\mathbb{F}_{p}):

y+k2z2b2=0,zk3=0,y+\frac{k_{2}z}{2}-\frac{b}{2}=0,z-k_{3}=0,

which gives y=k4,z=k3y=k_{4},z=k_{3}. Therefore, in 𝔽p\mathbb{F}_{p} we have

2x+yz=k1k2+k3k4,2y+xz=k1k4+k2k3,2z+xy=k1k3+k2k4,2x+yz=k_{1}k_{2}+k_{3}k_{4},2y+xz=k_{1}k_{4}+k_{2}k_{3},2z+xy=k_{1}k_{3}+k_{2}k_{4},

which implies that fx=fy=fz=0f^{\prime}_{x}=f^{\prime}_{y}=f^{\prime}_{z}=0. This cannot be true for every p\mathbb{Z}_{p}-point (x,y,z)(x,y,z) of 𝒰\mathcal{U}, since by Assumption A we have proved in Proposition 4.1 that there always exists a non-singular solution modulo pp of the affine Markoff-type cubic equation (with at least one coordinate zero) which lifts to a p\mathbb{Z}_{p}-point by Hensel’s lemma.

In conclusion, the local invariant map of 𝒜\mathcal{A} at the prime pp satisfying our hypothesis is indeed surjective. ∎

Theorem 4.3.

Let k(\[2,2])4k\in(\mathbb{Z}\,\backslash\,[-2,2])^{4} satisfy the hypotheses of Assumptions A and B. Then we have a Brauer–Manin obstruction to strong approximation for 𝒰\mathcal{U} given by the class of 𝒜=𝒜1\mathcal{A}=\mathcal{A}_{1} in Br1U/Br\textup{Br}_{1}\,U/\textup{Br}\,\mathbb{Q} (i.e. 𝒰(A)𝒰(A)𝒜=𝒰(A)Br1\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\mathcal{A}}=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}) and no algebraic Brauer–Manin obstruction to the integral Hasse principle for 𝒰\mathcal{U} (i.e. 𝒰(A)Br1\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}\not=\emptyset).

Proof.

By abuse of notation, we also denote the class of 𝒜\mathcal{A} by the Brauer group element itself. For any point u=(x,y,z)𝒰(A)\textbf{u}=(x,y,z)\in\mathcal{U}(\textbf{A}_{\mathbb{Z}}), from the above Lemma we can find a local point up=(xp,yp,zp)𝒰(p)\textbf{u}_{p}=(x_{p},y_{p},z_{p})\in\mathcal{U}(\mathbb{Z}_{p}) such that invp𝒜(up)=qpinvq𝒜(u)\textup{inv}_{p}\,\mathcal{A}(\textbf{u}_{p})=-\sum_{q\not=p}\textup{inv}_{q}\mathcal{A}(\textbf{u}) and up=(xp,yp,zp)𝒰(p)\textbf{u}_{p}^{\prime}=(x_{p}^{\prime},y_{p}^{\prime},z_{p}^{\prime})\in\mathcal{U}(\mathbb{Z}_{p}) such that invp𝒜(up)=1/2qpinvq𝒜(u)\textup{inv}_{p}\,\mathcal{A}(\textbf{u}_{p}^{\prime})=1/2-\sum_{q\not=p}\textup{inv}_{q}\,\mathcal{A}(\textbf{u}). Then we obtain a point in 𝒰(A)𝒜\mathcal{U}(\textbf{A}_{\mathbb{Z}})^{\mathcal{A}} by replacing the pp-part of u by up\textbf{u}_{p}, and a point in 𝒰(A)\mathcal{U}(\textbf{A}_{\mathbb{Z}}) but not in 𝒰(A)𝒜\mathcal{U}(\textbf{A}_{\mathbb{Z}})^{\mathcal{A}} by replacing the pp-part of u by up\textbf{u}_{p}^{\prime}.

Therefore, we have a Brauer–Manin obstruction to strong approximation and no algebraic Brauer–Manin obstruction to the integral Hasse principle for 𝒰\mathcal{U}. ∎

Example 4.1.

For p=11p=11: Take k1=113.102+13=135775,k2=112.112+111=13663,k3=112.119+6=14405,k4=112.(119+144)+6=31829k_{1}=11^{3}.102+13=135775,k_{2}=11^{2}.112+111=13663,k_{3}=11^{2}.119+6=14405,k_{4}=11^{2}.(119+144)+6=31829.

Remark 4.2.

In addition, if kk 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 k(\[2,2])4k\in(\mathbb{Z}\,\backslash\,[-2,2])^{4} satisfying k11k_{1}\equiv-1 mod 1616, ki5k_{i}\equiv 5 mod 1616 and k11k_{1}\equiv 1 mod 99, ki5k_{i}\equiv 5 mod 99 for 2i42\leq i\leq 4, such that (ki,kj)=1(k_{i},k_{j})=1, (ki24,kj24)=3(k_{i}^{2}-4,k_{j}^{2}-4)=3 for 1ij41\leq i\not=j\leq 4, and (k122,k222,k322,k422)=1(k_{1}^{2}-2,k_{2}^{2}-2,k_{3}^{2}-2,k_{4}^{2}-2)=1 (then satisfying the additional hypothesis in Lemma 4.2). Note that the finite number of k[2,2]4k\in[-2,2]^{4} 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 kik_{i} and the common divisor condition

gcd(ki(ki22)(ki24),kj(kj22)(kj24))=3\textup{gcd}(k_{i}(k_{i}^{2}-2)(k_{i}^{2}-4),k_{j}(k_{j}^{2}-2)(k_{j}^{2}-4))=3

for 1ij41\leq i\not=j\leq 4. 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 fi[x1,,xn]f_{i}\in\mathbb{Z}[x_{1},\dots,x_{n}], 1is1\leq i\leq s for some s>1s\in\mathbb{Z}_{>1}, be ss polynomials that are mutually relatively prime as elements of [x1,,xn]\mathbb{Q}[x_{1},\dots,x_{n}]. For each 1ijs1\leq i\not=j\leq s, let

i,j:={an:gcd(fi(a),fj(a))=1}.\mathcal{R}_{i,j}:=\{a\in\mathbb{Z}^{n}:\textup{gcd}(f_{i}(a),f_{j}(a))=1\}.

Then μ(iji,j)=p(1cp/pn)\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j})=\prod_{p}(1-c_{p}/p^{n}), where pp ranges over all primes of \mathbb{Z}, and cpc_{p} is the number of x(/p)nx\in(\mathbb{Z}/p\mathbb{Z})^{n} satisfying at least one of fi(x)=fj(x)=0f_{i}(x)=f_{j}(x)=0 in /p\mathbb{Z}/p\mathbb{Z} for 1ijs1\leq i\not=j\leq s.

Proof.

We also have a generalization of [Poo03, Lemma 5.1] as follows.

Lemma 4.5.

Let fi[x1,,xn]f_{i}\in\mathbb{Z}[x_{1},\dots,x_{n}], 1is1\leq i\leq s for some s>1s\in\mathbb{Z}_{>1}, be ss polynomials that are mutually relatively prime as elements of [x1,,xn]\mathbb{Q}[x_{1},\dots,x_{n}]. For each 1ijs1\leq i\not=j\leq s, let

𝒬i,j,M:={an:psuch thatpMandp|fi(a),fj(a)}.\mathcal{Q}_{i,j,M}:=\{a\in\mathbb{Z}^{n}:\exists\,p\;\textup{such that}\;p\geq M\;\textup{and}\;p\,|\,f_{i}(a),f_{j}(a)\}.

Then limMμ¯(ij𝒬i,j,M)=0\lim_{M\rightarrow\infty}\overline{\mu}(\bigcup_{i\not=j}\mathcal{Q}_{i,j,M})=0.

Proof.

The result immediately follows from the inequality μ¯(ij𝒬i,j,M)ijμ¯(𝒬i,j,M)\overline{\mu}(\bigcup_{i\not=j}\mathcal{Q}_{i,j,M})\leq\sum_{i\not=j}\overline{\mu}(\mathcal{Q}_{i,j,M}) and the original result of [Poo03, Lemma 5.1]. ∎

Next, we proceed similarly as in the proof of [Poo03, Theorem 3.1]. Let PMP_{M} denote the set of prime numbers of \mathbb{Z} such that p<Mp<M. Approximate i,j\mathcal{R}_{i,j} by

i,j,M:={an:fi(a)andfj(a)are not both divisible by any primepPM}.\mathcal{R}_{i,j,M}:=\{a\in\mathbb{Z}^{n}:f_{i}(a)\;\textup{and}\;f_{j}(a)\;\textup{are not both divisible by any prime}\;p\in P_{M}\}.

Define the ideal II as the product of all (p)(p) for pPMp\in P_{M}. Then i,j,M\mathcal{R}_{i,j,M} is a union of cosets of the subgroup InnI^{n}\subset\mathbb{Z}^{n}. Hence μ(iji,j,M)\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j,M}) is the fraction of residue classes in (/I)n(\mathbb{Z}/I)^{n} in which for all pPMp\in P_{M}, for all 1ijs1\leq i\not=j\leq s, at least one of fi(a)f_{i}(a) and fj(a)f_{j}(a) is nonzero modulo pp. Applying the Chinese Remainder Theorem, we obtain that μ(iji,j,M)=pPM(1cp/pn)\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j,M})=\prod_{p\in P_{M}}(1-c_{p}/p^{n}). By the above lemma,

μ(iji,j)=limMμ(iji,j,M)=p(1cp/pn).\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j})=\lim_{M\rightarrow\infty}\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j,M})=\prod_{p}(1-c_{p}/p^{n}). (24)

Since fif_{i} and fjf_{j} are relatively prime as elements of [x1,,xn]\mathbb{Q}[x_{1},\dots,x_{n}] for any iji\not=j, there exists a nonzero uu\in\mathbb{Z} such that every fi=fj=0f_{i}=f_{j}=0 defines a subscheme of 𝔸[1/u]n\mathbb{A}^{n}_{\mathbb{Z}[1/u]} of codimension at least 22. Thus cp=O(pn2)c_{p}=O(p^{n-2}) as pp\rightarrow\infty, and the product converges. ∎

Theorem 4.6.

Let 𝒰\mathcal{U} be the affine scheme over \mathbb{Z} defined by

x2+y2+z2+xyz=ax+by+cz+d,x^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d,

where

{a=k1k2+k3k4b=k1k4+k2k3c=k1k3+k2k4andd=4i=14ki2i=14ki,\begin{cases*}a=k_{1}k_{2}+k_{3}k_{4}\\ b=k_{1}k_{4}+k_{2}k_{3}\\ c=k_{1}k_{3}+k_{2}k_{4}\end{cases*}\hskip 14.22636pt\textup{and}\hskip 14.22636ptd=4-\sum_{i=1}^{4}k_{i}^{2}-\prod_{i=1}^{4}k_{i},

such that the projective closure X3X\subset\mathbb{P}^{3}_{\mathbb{Q}} of U=𝒰×U=\mathcal{U}\times_{\mathbb{Z}}\mathbb{Q} is smooth. Then we have

#{k=(k1,k2,k3,k4)4,|ki|M 1i4:𝒰(A)}M4\#\{k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{Z}^{4},|k_{i}|\leq M\;\forall\;1\leq i\leq 4:\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})\not=\emptyset\}\asymp M^{4} (25)

and also

#{k=(k1,k2,k3,k4)4,|ki|M 1i4:𝒰(A)𝒰(A)Br1}M4\#\{k=(k_{1},k_{2},k_{3},k_{4})\in\mathbb{Z}^{4},|k_{i}|\leq M\;\forall\;1\leq i\leq 4:\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}\not=\emptyset\}\asymp M^{4} (26)

as M+M\rightarrow+\infty.

Proof.

Denote by 𝒮\mathcal{S} the set in question and \mathcal{M} is the set of kk such that |ki|M|k_{i}|\leq M for 1i41\leq i\leq 4. Hence, using the same notation as in Proposition 4.4, the density we are concerned with is given by

μ(𝒮)=limM#(𝒮)#=limM#(𝒮)M4.\mu(\mathcal{S})=\lim_{M\rightarrow\infty}\frac{\#(\mathcal{S}\cap\mathcal{M})}{\#\mathcal{M}}=\lim_{M\rightarrow\infty}\frac{\#(\mathcal{S}\cap\mathcal{M})}{M^{4}}.

We apply Proposition 4.4 with the polynomials fi=xi(xi22)(xi24)/3[x1,x2,x3,x4]f_{i}=x_{i}(x_{i}^{2}-2)(x_{i}^{2}-4)/3\in\mathbb{Z}[x^{\prime}_{1},x^{\prime}_{2},x^{\prime}_{3},x^{\prime}_{4}] for 1i41\leq i\leq 4 in which we write x1=144x1+127x_{1}=144x^{\prime}_{1}+127, xj=144xj+5x_{j}=144x^{\prime}_{j}+5 for 2j42\leq j\leq 4 (resp. xi=144.p03xi+rix_{i}=144.p_{0}^{3}x^{\prime}_{i}+r_{i} where the xix_{i} is actually the kik_{i} and p011p_{0}\geq 11 is the chosen prime in Lemma 4.2 and the residues rir_{i} satisfy the hypotheses of the lemma), and using the inclusion-exclusion principle we can compute that c2=0c_{2}=0, c3=0c_{3}=0,

cp,1=C42.32.p2(3.C43.33.p+3.34)+(4.33.p+16.34)C64.34+C65.3434=27(2p28p+9)c_{p,1}=C^{2}_{4}.3^{2}.p^{2}-(3.C^{3}_{4}.3^{3}.p+3.3^{4})+(4.3^{3}.p+16.3^{4})-C^{4}_{6}.3^{4}+C^{5}_{6}.3^{4}-3^{4}=27(2p^{2}-8p+9)

if p>3p>3 and p±3(mod 8)p\equiv\pm 3\,(\textup{mod}\,8) and

cp,2=C42.52.p2(3.C43.53.p+3.54)+(4.53.p+16.54)C64.54+C65.5454=25(6p240p+75)c_{p,2}=C^{2}_{4}.5^{2}.p^{2}-(3.C^{3}_{4}.5^{3}.p+3.5^{4})+(4.5^{3}.p+16.5^{4})-C^{4}_{6}.5^{4}+C^{5}_{6}.5^{4}-5^{4}=25(6p^{2}-40p+75)

if p>3p>3 and p±1(mod 8)p\equiv\pm 1\,(\textup{mod}\,8) (resp. except for p0p_{0} with p0±3p_{0}\equiv\pm 3 mod 88, since k2k322k422k_{2}\equiv k_{3}^{2}-2\equiv k_{4}^{2}-2, k3k40k_{3}\equiv k_{4}\not\equiv 0 and k2240k_{2}^{2}-4\not\equiv 0 (mod p0p_{0}), we have ki(ki22)(ki24)0k_{i}(k_{i}^{2}-2)(k_{i}^{2}-4)\not\equiv 0 mod p0p_{0} for all 2i42\leq i\leq 4, and so cp0=0).c_{p_{0}}=0\big{)}. Applying (24)(24) gives us a positive density in (25)(25):

μ(iji,j)=11444p>3p±3(mod 8)(1cp,1p4)p>3p±1(mod 8)(1cp,2p4),\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j})=\frac{1}{144^{4}}\prod_{\begin{subarray}{c}p>3\\ p\equiv\pm 3\,(\textup{mod}\,8)\end{subarray}}\left(1-\frac{c_{p,1}}{p^{4}}\right)\prod_{\begin{subarray}{c}p>3\\ p\equiv\pm 1\,(\textup{mod}\,8)\end{subarray}}\left(1-\frac{c_{p,2}}{p^{4}}\right),

and also in (26)(26):

μ(iji,j)=11444.p012p>3,pp0p±3(mod 8)(1cp,1p4)p>3p±1(mod 8)(1cp,2p4).\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j})=\frac{1}{144^{4}.p_{0}^{12}}\prod_{\begin{subarray}{c}p>3,p\not=p_{0}\\ p\equiv\pm 3\,(\textup{mod}\,8)\end{subarray}}\left(1-\frac{c_{p,1}}{p^{4}}\right)\prod_{\begin{subarray}{c}p>3\\ p\equiv\pm 1\,(\textup{mod}\,8)\end{subarray}}\left(1-\frac{c_{p,2}}{p^{4}}\right).

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 kk with |ki|M|k_{i}|\leq M for 1i41\leq i\leq 4 such that the surfaces are singular is just O(M3)O(M^{3}) as MM\rightarrow\infty, hence it is negligible. Therefore, we obtain that μ(𝒮)μ(iji,j)>0\mu(\mathcal{S})\geq\mu(\bigcap_{i\not=j}\mathcal{R}_{i,j})>0. ∎

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 |m|M|m|\leq M as M+M\rightarrow+\infty, the number of counterexamples to the integral Hasse principle which can be explained by the Brauer–Manin obstruction is M1/2/(logM)1/2M^{1/2}/(\log M)^{1/2} 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 k=(k1,,kn)𝔸n()k=(k_{1},\dots,k_{n})\in\mathbb{A}^{n}(\mathbb{C}), we write

H(k)=max{1,|k1|,,|kn|}.\textup{H}(k)=\max\{1,|k_{1}|,\dots,|k_{n}|\}.

Surfaces of type (1,1)(1,1). Let Σ\Sigma be a surface of type (1,1)(1,1). By Section 2, we have an identification of the moduli space Xk=Xk(Σ)X_{k}=X_{k}(\Sigma) with the affine cubic algebraic surface in 𝔸x,y,z3\mathbb{A}^{3}_{x,y,z} given by the equation

x2+y2+z2xyz2=k.x^{2}+y^{2}+z^{2}-xyz-2=k.

The mapping class group Γ=Γ(Σ)\Gamma=\Gamma(\Sigma) acts on XkX_{k} via polynomial transformations. Up to finite index, it coincides with the group Σ\Sigma^{\prime} of automorphisms of XkX_{k} generated by the transpositions and even sign changes of coordinates as well as the Vieta involutions of the form (x,y,z)(x,y,xyz)(x,y,z)\mapsto(x,y,xy-z).

The mapping class group dynamics on Xk()X_{k}(\mathbb{R}) 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 Σ\Sigma^{\prime} on the integral points Xk()X_{k}(\mathbb{Z}) for admissible kk. More generally, the results below establish Markoff descent for complex points.

Lemma 5.1.

[Wha20, Lemma 4.2] Let Σ\Sigma be a surface of type (1,1)(1,1). There is a constant C>0C>0 independent of kk\in\mathbb{C} such that, given any ρXk(Σ,)\rho\in X_{k}(\Sigma,\mathbb{C}), there exists some γΓ\gamma\in\Gamma such that γρ=(x,y,z)\gamma^{*}\rho=(x,y,z) satisfies

min{|x|,|y|,|z|}CH(k)1/3.\min\{|x|,|y|,|z|\}\leq C\cdot\textup{H}(k)^{1/3}.

Surfaces of type (0,4)(0,4). Let Σ\Sigma be a surface of type (0,4)(0,4). By Section 2, we have an identification of the moduli space Xk=Xk(Σ)X_{k}=X_{k}(\Sigma) with the affine cubic algebraic surface in 𝔸x,y,z3\mathbb{A}^{3}_{x,y,z} given by the equation

x2+y2+z2+xyz=ax+by+cz+dx^{2}+y^{2}+z^{2}+xyz=ax+by+cz+d

with a,b,c,da,b,c,d appropriately determined by kk. The mapping class group Γ=Γ(Σ)\Gamma=\Gamma(\Sigma) acts on XkX_{k} via polynomial transformations. Let Σ\Sigma^{\prime} be the group of automorphisms of XkX_{k} generated by the Vieta involutions

τx\displaystyle\tau^{*}_{x} :(x,y,z)(ayzx,y,z),\displaystyle:(x,y,z)\mapsto(a-yz-x,y,z),
τy\displaystyle\tau^{*}_{y} :(x,y,z)(x,bxzy,z),\displaystyle:(x,y,z)\mapsto(x,b-xz-y,z),
τz\displaystyle\tau^{*}_{z} :(x,y,z)(x,y,cxyz).\displaystyle:(x,y,z)\mapsto(x,y,c-xy-z).

Two points ρ,ρXk()\rho,\rho^{\prime}\in X_{k}(\mathbb{C}) are Σ\Sigma^{\prime}-equivalent if and only if they are Σ\Sigma-equivalent or ρ\rho is Σ\Sigma-equivalent to all of τxρ,τyρ\tau^{*}_{x}\rho^{\prime},\tau^{*}_{y}\rho^{\prime}, and τzρ\tau^{*}_{z}\rho^{\prime}.

Lemma 5.2.

[Wha20, Lemma 4.4] Let Σ\Sigma be a surface of type (0,4)(0,4). There is a constant C>0C>0 independent of k4k\in\mathbb{C}^{4} such that, given any ρXk()\rho\in X_{k}(\mathbb{C}), there exists some γk\gamma\in k such that γρ=(x,y,z)\gamma^{*}\rho=(x,y,z) satisfies one of the following conditions:

  1. (1)

    min{|x|,|y|,|z|}C\min\{|x|,|y|,|z|\}\leq C,

  2. (2)

    |yz|CH(a)|yz|\leq C\cdot\textup{H}(a),

  3. (3)

    |xz|CH(b)|xz|\leq C\cdot\textup{H}(b),

  4. (4)

    |xy|CH(c)|xy|\leq C\cdot\textup{H}(c),

  5. (5)

    |xyz|CH(d)|xyz|\leq C\cdot\textup{H}(d).

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 kk 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 k1=127,k2=5,k3=5+144.5=725,k4=5+144.10=1445k_{1}=127,k_{2}=5,k_{3}=5+144.5=725,k_{4}=5+144.10=1445, then k=(k1,k2,k3,k4)k=(k_{1},k_{2},k_{3},k_{4}) 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

x2+y2+z2+xyz=1048260x+187140y+99300z667871675.x^{2}+y^{2}+z^{2}+xyz=1048260x+187140y+99300z-667871675.

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 CC independent of k4k\in\mathbb{C}^{4} may take the value C=48C=\textbf{48} in the condition (1)(1) and C=24C=\textbf{24} in all the other conditions (2),(3),(4),(5)(2),(3),(4),(5) (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 (2),(3),(4),(5)(2),(3),(4),(5), along with the help of Dario Alpern’s website (Alpertron) [Alp] to find integral points whose one coordinate satisfies the restriction (1)(1). More precisely, SageMath shows that there is no integral point in any of the boxes defined by (2),(3),(4),(5)(2),(3),(4),(5), while Alpertron deals with conic equations after fixing one variable bounded in (1)(1) 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 a=b=c=0a=b=c=0 in (2)(2) implies that either three of the kik_{i} are equal to 0 or k1=k2=k3=k4k_{1}=k_{2}=k_{3}=-k_{4} or any other permutation of kik_{i} (with 1i41\leq i\leq 4), the only cases that our Markoff-type cubic surfaces recover the original Markoff surfaces are given by equations of the form

x2+y2+z2+xyz=4k02,x^{2}+y^{2}+z^{2}+xyz=4-k_{0}^{2},

or

x2+y2+z2+xyz=(2k02)2x^{2}+y^{2}+z^{2}+xyz=(2-k_{0}^{2})^{2}

for some k0k_{0}\in\mathbb{Z}. 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 k0k_{0} the equation will not be everywhere locally solvable as the right hand side is congruent to 33 mod 44 (see [GS22]). For the latter equation, it always has the integral solutions (±(2k02),0,0)(\pm(2-k_{0}^{2}),0,0). Although these special cases are only of magnitude M1/4M^{1/4} 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 k1k2=k3=k4k_{1}\not=k_{2}=k_{3}=k_{4} are integers such that the total field extension [E:]=4[E:\mathbb{Q}]=4 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

Br1U/Br0U(/2)3\textup{Br}_{1}\,U/\textup{Br}_{0}\,U\cong(\mathbb{Z}/2\mathbb{Z})^{3}

with three generators

{(x2,k224),(y2,k224),(z2,k224)}.\{(x-2,k_{2}^{2}-4),(y-2,k_{2}^{2}-4),(z-2,k_{2}^{2}-4)\}.

Following [LM20, Proposition 5.7 and Lemma 5.8] or [CWX20, Lemmas 5.4 and 5.5], we have

{(x2,k224)2,(y2,k224)2,(z2,k224)2}={0,1/2,1/2}\{(x-2,k_{2}^{2}-4)_{2},(y-2,k_{2}^{2}-4)_{2},(z-2,k_{2}^{2}-4)_{2}\}=\{0,1/2,1/2\}

and

{(x2,k224)3,(y2,k224)3,(z2,k224)3}={0,0,}\{(x-2,k_{2}^{2}-4)_{3},(y-2,k_{2}^{2}-4)_{3},(z-2,k_{2}^{2}-4)_{3}\}=\{0,0,-\}

as multisets. As k2k_{2} are odd, so k2245k_{2}^{2}-4\equiv 5 mod 88, hence there is no way to describe this number by the form 3v23v^{2} for vv\in\mathbb{Z}, which is favorable to give Hasse failures as in previous work. Due to the complexity of the integral values of the polynomial X24X^{2}-4 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 k1k2=k3=k4k_{1}\not=k_{2}=k_{3}=k_{4} are integers satisfying the congruence conditions k1k25k_{1}\equiv-k_{2}\equiv-5 mod 1616 (resp. mod 99), and the divisibility conditions: (k1,k2)=1(k_{1},k_{2})=1, (k1,k224)=(k124,k2)=1(k_{1},k_{2}^{2}-4)=(k_{1}^{2}-4,k_{2})=1, and (k122,k222)=1(k_{1}^{2}-2,k_{2}^{2}-2)=1, such that [E:]=4[E:\mathbb{Q}]=4. Moreover, assume that there is a prime p>3p>3 such that pp divides k224k_{2}^{2}-4 to an odd order and k1k2k_{1}\equiv-k_{2} mod pp. Then there is no algebraic Brauer–Manin obstruction to the integral Hasse principle, but there is a Brauer–Manin obstruction to strong approximation for 𝒰\mathcal{U}.

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, a=b=c0a=b=c\equiv 0 mod pp and d=(2k22)24d=(2-k_{2}^{2})^{2}\equiv 4 mod pp, so ax+by+cz+d4ax+by+cz+d\equiv 4 mod pp in the defining equation (1)(1) of 𝒰\mathcal{U}. Now for p>5p>5, we let =(x2,k224),(y2,k224),(z2,k224)\mathcal{B}=\langle(x-2,k_{2}^{2}-4),(y-2,k_{2}^{2}-4),(z-2,k_{2}^{2}-4)\rangle and follow the same arguments in the proof of [LM20, Proposition 5.5] to prove that the map

𝒰(p)Hom(,/),u(βinvpβ(u)),\mathcal{U}(\mathbb{Z}_{p})\rightarrow\textup{Hom}(\mathcal{B},\mathbb{Q}/\mathbb{Z}),\hskip 28.45274pt\textbf{u}\mapsto(\beta\mapsto\textup{inv}_{p}\,\beta(\textbf{u})),

induced by the Brauer–Manin pairing, is surjective. Now we need to show that 𝒰(A)Br1𝒰(A)\emptyset\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}})^{\textup{Br}_{1}}\not=\mathcal{U}(\textbf{{A}}_{\mathbb{Z}}). Let u𝒰(A)\textbf{u}\in\mathcal{U}(\textbf{{A}}_{\mathbb{Z}}). Then by surjectivity, there exist up𝒰(p)\textbf{u}_{p}\in\mathcal{U}(\mathbb{Z}_{p}) such that invpβ(up)=vpinvvβ(u)\textup{inv}_{p}\,\beta(\textbf{u}_{p})=-\sum_{v\not=p}\textup{inv}_{v}\beta(\textbf{u}) for all β\beta\in\mathcal{B} and up𝒰(p)\textbf{u}_{p}^{\prime}\in\mathcal{U}(\mathbb{Z}_{p}) such that invpβ(up)=1/2vpinvvβ(u)\textup{inv}_{p}\,\beta(\textbf{u}_{p}^{\prime})=1/2-\sum_{v\not=p}\textup{inv}_{v}\,\beta(\textbf{u}) for some β\beta\in\mathcal{B}. Then we obtain a point in 𝒰(A)Br1\mathcal{U}(\textbf{A}_{\mathbb{Z}})^{\textup{Br}_{1}} by replacing the pp-part of u by up\textbf{u}_{p}, and a point in 𝒰(A)\mathcal{U}(\textbf{A}_{\mathbb{Z}}) but not in 𝒰(A)Br1\mathcal{U}(\textbf{A}_{\mathbb{Z}})^{\textup{Br}_{1}} by replacing the pp-part of u by up\textbf{u}_{p}^{\prime}, as required.

We are left with the case p=5p=5. 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 u𝒰(A)\textbf{u}\in\mathcal{U}(\textbf{{A}}_{\mathbb{Z}}). In fact, there always exists up𝒰(p)\textbf{u}_{p}^{\prime}\in\mathcal{U}(\mathbb{Z}_{p}) such that invpβ(up)=1/2vpinvvβ(u)\textup{inv}_{p}\,\beta(\textbf{u}_{p}^{\prime})=1/2-\sum_{v\not=p}\textup{inv}_{v}\,\beta(\textbf{u}) for some β\beta\in\mathcal{B}. Next, if vpinvvβ(u)0\sum_{v\not=p}\textup{inv}_{v}\beta^{\prime}(\textbf{u})\not=0 for some β\beta^{\prime}\in\mathcal{B}, there still exists up𝒰(p)\textbf{u}_{p}\in\mathcal{U}(\mathbb{Z}_{p}) such that invpβ(up)=vpinvvβ(u)\textup{inv}_{p}\,\beta(\textbf{u}_{p})=-\sum_{v\not=p}\textup{inv}_{v}\beta(\textbf{u}) for all β\beta\in\mathcal{B}. Now if vpinvvβ(u)=0\sum_{v\not=p}\textup{inv}_{v}\beta(\textbf{u})=0 for all β\beta\in\mathcal{B}, we can consider another local point u′′\textbf{u}^{\prime\prime} whose pp-parts (p2p\not=2) are the same as those of u and 22-part (x2′′,y2′′,z2′′)(x_{2}^{\prime\prime},y_{2}^{\prime\prime},z_{2}^{\prime\prime}) is a permutation of u2=(x2,y2,z2)\textbf{u}_{2}=(x_{2},y_{2},z_{2}) (note that the Markoff-type cubic equation here is symmetric in x,y,zx,y,z) such that their images under the local invariant map at 22 are different permutations of (0,1/2,1/2)(0,1/2,1/2), hence we will get vpinvvβ(u)0\sum_{v\not=p}\textup{inv}_{v}\beta(\textbf{u})\not=0 for some β\beta\in\mathcal{B}. 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 k1=144.7.25=2011k_{1}=144.7.2-5=2011 and k2=k3=k4=5k_{2}=k_{3}=k_{4}=5. Then there is no integral point on the corresponding Markoff-type cubic surface. By Proposition 5.3 (with p=7p=7), we obtain a counterexample to the integral Hasse principle which cannot be explained by the algebraic Brauer–Manin obstruction.

Example 5.3.

Consider k1=144.7.35=3019k_{1}=144.7.3-5=3019 and k2=k3=k4=5k_{2}=k_{3}=k_{4}=5. Then we find an integral point (x,y,z)=(24,409,672)(x,y,z)=(24,409,672) on the corresponding Markoff-type cubic surface. By Proposition 5.3 (with p=7p=7), 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 k1=127k_{1}=127 and k2=k3=k4=5k_{2}=k_{3}=k_{4}=5. Since (12724,524)=3(127^{2}-4,5^{2}-4)=3, 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
\missing

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]