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

Rationality of Brauer-Severi surface bundles over rational 3-folds

Shitan Xu Department of Mathematics, Michigan State University, East Lansing, MI 48824-1027, USA [email protected]
Abstract.

We give a sufficient condition for a Brauer-Severi surface bundle over a rational 3-fold to not be stably rational. Additionally, we present an example that satisfies this condition and demonstrate the existence of families of Brauer-Severi surface bundles whose general members are smooth and not stably rational.

Key words and phrases:
Rationality, Brauer-Severi surface bundle, algebraic geometry
2020 Mathematics Subject Classification:
Primary 14E08,14M20

1. Introduction

This paper is motivated by the study of the stable rationality of conic bundles and Brauer-Severi surface bundles. Let XX be a projective variety over an algebraically closed field kk. We say that XX is rational if XX is birational to a projective space kn{\mathbb{P}}^{n}_{k} for some natural number nn. We call XX stably rational if there exists a natural number mm such that X×kkmX\times_{k}{\mathbb{P}}^{m}_{k} is rational. In [HPT18], an example of a quadratic surface bundle over 2{\mathbb{P}}^{2} that is not stably rational and has a nontrivial unramified Brauer group is constructed. This example is realized as a divisor in 2×3{\mathbb{P}}^{2}\times{\mathbb{P}}^{3} of bidegree (2,2), with the quadratic surface bundle structure induced by the first projection.

In [ABvBP20], this elegant example was examined from a different perspective: it naturally becomes a conic bundle over 3{\mathbb{P}}^{3} via the second projection. This structure allowed the authors to establish a sufficient condition [ABvBP20, Thm. 2.6] for a conic bundle over 3{\mathbb{P}}^{3} to not be stably rational. Furthermore, they introduced a new example of a flat family of conic bundles over 3{\mathbb{P}}^{3}, where a very general member is not stably rational, using a theorem of Voisin [Voi15, Thm. 2.1], in the form of [CTP16, Thm. 2.3]. Voisin’s theorem serves as a powerful tool in the study of the stable rationality of families of varieties. This approach is commonly referred to in the literature as the specialization method. In a recent paper [Pir23], Pirutka introduced the notion of the relative unramified cohomology group, which combines the approach of constructing nontrivial unramified Brauer class via fibrations ([AM72],[CTO89]) and the method given in [Sch19a] and [Sch19b] to avoid geometric construction of universally CH0CH_{0}-trivial desingularization. Our proof in section 6 applied this idea in a different manner.

A natural next step following these advancements is to investigate the stable rationality of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}. In 2017, Kresch and Tschinkel introduced good models of Brauer-Severi surface bundles using the concept of a root stack [KT19]. With this definition, they constructed a flat family of Brauer-Severi surface bundles over 2{\mathbb{P}}^{2} [KT20, Thm. 1], in which the general member is smooth and not stably rational.

This paper begins by generalizing Theorem 2.6 of [ABvBP20] to Brauer-Severi surface bundles over 3-folds. After constructing a new (singular) example with a nontrivial unramified Brauer group, we obtain the following result:

Theorem 1.1.

There exists a flat projective family of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, where a general fiber in this family is smooth and not stably rational.

The structure of this paper is as follows: In Section 2, we review some basic facts about Brauer groups and introduce the unramified Brauer group, a stably birational invariant. By definition, this is the subgroup of the Brauer group of the function field, whose elements arise from the Brauer classes of a smooth model. Recently, significant progress has been made using the unramified Brauer group and the specialization method to show that a very general member of certain classes of varieties is not stably rational ([Pir18, Section 2.1]).

In Section 3, we generalize a criterion for the stable rationality of conic bundles over 3-folds ([ABvBP20, Thm. 2.6]) to the case of Brauer-Severi surface bundles. This provides a tool to construct an explicit (singular) Brauer-Severi surface bundle over 3{\mathbb{P}}^{3}_{{\mathbb{C}}} with a nontrivial unramified Brauer group, which we explore in Section 4. In Section 5, we verify that this example is indeed a Brauer-Severi surface bundle.

In Section 6, we show that our example 4.1 satisfies the hypotheses required by the specialization method introduced by Voisin in 2014 [Voi15], and further developed by Colliot-Thélène and Pirutka in 2016 [CTP16]. Following Schreieder’s approach [Sch19a, Proposition 26], we verify this using a purely cohomological criterion. Finally, in Section 7, we prove Theorem 1.1 by constructing a flat family of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, where the general member is smooth and includes Example 4.1 as a member. Detailed calculations are provided separately in Appendix A.

1.1. Acknowledgements

I thank my advisor, Professor Rajesh Kulkarni, for his valuable discussions and suggestions on this work. I also extend my gratitude to Prof. Pirutka, Prof. Auel, Prof.Kresch and Prof. Schreieder for their comments on an earlier draft. During this project, I was partially supported by NSF grant DMS-2101761.

2. Background

2.1. Brauer groups and purity

For the definition and a detailed treatment of Brauer groups of fields and schemes, see [GS06] or [CTS21]. We use standard conventions in Galois cohomology below. In particular,

Hi(L,M)=Hi(Gal(L¯/L),M).H^{i}(L,M)=H^{i}(\textup{Gal}(\bar{L}/L),M).

Let LL be the function field of an integral scheme ZZ over the field of complex numbers, {\mathbb{C}}. The Kummer sequence provides a Galois cohomology sequence that identifies the 3-torsion subgroup of Br(L)\operatorname{Br}(L) with the second Galois cohomology group of LL with constant coefficients /3{\mathbb{Z}}/3:

H2(L,/3)Br(L)[3].H^{2}(L,{\mathbb{Z}}/3)\cong\operatorname{Br}(L)[3].

By the Merkurjev-Suslin theorem [GS06, Thm. 2.5.7], Br(L)[3]\operatorname{Br}(L)[3] is generated by the equivalence classes of cyclic algebras of order 3. These are denoted by (a,b)ω(a,b)_{\omega} and are defined by

(a,b)ω=x,y|x3=a,y3=b,xy=ωyx,(a,b)_{\omega}=\left<x,y|x^{3}=a,y^{3}=b,xy=\omega yx\right>,

where a,bL×a,b\in L^{\times} and ω\omega is a primitive third root of unity.

Now let ν\nu be a discrete valuation of LL with residue field k(ν)k(\nu), we have the following residue maps [GS06, Section 6.8]:

ν1:H1(L,/3)H0(k(ν),/3)\partial_{\nu}^{1}:H^{1}(L,{\mathbb{Z}}/3)\to H^{0}(k(\nu),{\mathbb{Z}}/3)
ν2:H2(L,/3)H1(k(ν),/3)\partial_{\nu}^{2}:H^{2}(L,{\mathbb{Z}}/3)\to H^{1}(k(\nu),{\mathbb{Z}}/3)

By Kummer theory, we can identify these residue maps as:

ν1:L×/L×3/3\partial_{\nu}^{1}:L^{\times}/L^{\times 3}\to{\mathbb{Z}}/3
ν2:Br(L)[3]k(ν)×/k(ν)×3\partial_{\nu}^{2}:\operatorname{Br}(L)[3]\to k(\nu)^{\times}/k(\nu)^{\times 3}
Lemma 2.1.

With notations as above, the two residue maps are defined by:

ν1([a])=ν(a)(mod 3)\partial_{\nu}^{1}([a])=\nu(a)(\text{mod}\ 3)
ν2([(a,b)ω])=(1)ν(a)ν(b)aν(b)bν(a)modk(ν)×3\partial_{\nu}^{2}([(a,b)_{\omega}])=(-1)^{\nu(a)\nu(b)}{\frac{a^{\nu(b)}}{b^{\nu(a)}}}\mod k(\nu)^{\times 3}
Proof.

See [IOOV17, Thm. 2.18]

Definition 2.2.

The 3-torsion of unramified Brauer group of a field LL over another field kk, denoted by Hnr2(L/k,/3)H^{2}_{nr}(L/k,{\mathbb{Z}}/3{\mathbb{Z}}) or Brnr(L/k)[3]\operatorname{Br}_{nr}(L/k)[3], is the intersection of kernels of all residue maps ν2\partial_{\nu}^{2}, where ν\nu take values in all divisorial valuations of LL which are trivial on kk.

From the discussions in [CT95] and [CTS21, Corollary 6.2.10], the unramified Brauer group of L is a stably birational invariant for any model X of L, whether the model is nonsingular or singular. Here, a model X of L refers to an integral projective variety with function field L. If the model is nonsingular, we have the following:

Lemma 2.3.

Let kk be a field with characteristic not 3. Let XX be a regular, proper, integral variety over kk with function field LL, then we have

Br(X)[3]Brnr(L/k)[3]\operatorname{Br}(X)[3]\cong\operatorname{Br}_{nr}(L/k)[3]
Proof.

This is a direct result of [CTS21, Prop. 3.7.8]

Given a 3-torsion Brauer class in the Brauer group of the function field L, it is not easy to determine whether it belongs to the unramified subgroup using the definition alone, as there are usually too many divisorial valuations to consider. In [CT95], several theorems are established that reduce the number of valuations needed to check whether a Brauer class is unramified. Specifically, it suffices to check valuations corresponding to prime divisors in a smooth model. This result follows from the purity property of unramified Brauer groups. For more details on these theorems, see [CT95] and [CTS21, Section3.7]. A similar discussion can be found in [ABvBP20, Section2].

2.2. Brauer-Severi surface bundles

We use the definition of Brauer-Severi surface bundles from [KT19, Def. 4.1]:

Definition 2.4.

Let BB be a locally Noetherian scheme, in which 3 is invertible in the local rings. A Brauer-Severi surface bundle over BB is a flat projective morphism π:YB\pi:Y\to B such that the fiber over every geometric point of BB is isomorphic to one of the following:

  • 2{\mathbb{P}}^{2}

  • The union of three standard Hirzebruch surfaces 𝔽1\mathbb{F}_{1}, meeting transversally, such that any pair of them meets along a fiber of one and the (1)(-1)-curve of the other.

  • An irreducible scheme whose underlying reduced subscheme is isomorphic to the cone over a twisted cubic curve.

Let π:YB\pi:Y\to B be a Brauer-Severi surface bundle over a smooth projective rational threefold BB over a field kk whose generic fiber is smooth, and let S={bB|π1(b)is singular}S=\{b\in B|\pi^{-1}(b)\ \text{is singular}\} denote its discriminant locus. We consider SS with its reduced closed subscheme structure in BB, and since BB is Noetherian, SS consists of finitely many irreducible components, say S1,,SnS_{1},\cdots,S_{n}.

In the case of conic bundles considered in [ABvBP20, Def. 2.4], the authors focused on a special kind of conic bundles with a good discriminant locus. We will generalize the definition of a good discriminant locus to Brauer-Severi surface bundles in this context:

Definition 2.5.

We say the discriminant locus SS is good if the following conditions are satisfied:

  1. (1)

    Each irreducible component of SS is reduced. (This is assumed above.)

  2. (2)

    The fiber YsY_{s} over a general sSis\in S_{i} for each irreducible component SiS_{i} is geometrically the union of three standard Hirzebruch surfaces 𝔽1\mathbb{F}_{1} described in Definition 2.4.

  3. (3)

    The natural triple cover of SiS_{i} induced by π:YB\pi:Y\to B is irreducible.

  4. (4)

    By (3)(3), the fiber YFSiY_{F_{S_{i}}} over the generic point of SiS_{i} is irreducible. Thus, there is a natural map τ\tau from the cubic classes of function field of SiS_{i} to the cubic classes of function field of YFSiY_{F_{S_{i}}}. We assume the cubic extension over the function field of SiS_{i} induced by π:YB\pi:Y\to B generates the kernel of τ\tau.

Remark 2.6.

The last requirement is automatically satisfied in the case of conic bundles with the suitable definition given in [ABvBP20, Thm. 2.6]. However, in our case, this is not generally true. Note that the example we provided meets all these requirements (See Example 4.1).

Remark 2.7.

By Lemma 2.1, those SiS_{i} are precisely those irreducible surfaces such that the Brauer class of the generic fiber have a nontrivial residue along SiS_{i}. In particular, the discriminant locus is a divisor of the base with pure-dimensional irreducible components.

We end this section with a lemma that generalizes [ABvBP20, Lemma 2.3] to the 3-torsion case:

Lemma 2.8.

Let SS be a smooth nonsplit Brauer-Severi surface over an arbitrary field KK (in particular, SK¯K¯2S_{\bar{K}}\cong{\mathbb{P}}^{2}_{\bar{K}}). Then the pullback map Br(K)Br(S)\operatorname{Br}(K)\to\operatorname{Br}(S) induces an exact sequence:

0/3Br(K)Br(S)0,0\to{\mathbb{Z}}/3\to\operatorname{Br}(K)\to\operatorname{Br}(S)\to 0,

where the kernel is generated by the Brauer class αBr(K)[3]\alpha\in\operatorname{Br}(K)[3] determined by SS. Furthermore, if characteristic of KK 3\neq 3 and KK contains a primitive 99-th root of unit, then the above exact sequence restricts to :

0/3Br(K)[3]Br(S)[3]/300\to{\mathbb{Z}}/3\to\operatorname{Br}(K)[3]\to\operatorname{Br}(S)[3]\to{\mathbb{Z}}/3\to 0
Proof.

We first prove the exactness of the first sequence. Recall that the kernel ker(Br(K)Br(S))\textup{ker}(\operatorname{Br}(K)\to\operatorname{Br}(S)) is given by Amitsur’s theorem [GS06, Thm. 5.4.1]. To show surjectivity of Br(K)Br(S)\operatorname{Br}(K)\to\operatorname{Br}(S), consider the separable closure KsK^{s} of KK, and let Γ=Gal(Ks/K)\Gamma=\textup{Gal}(K^{s}/K). Then we have the Hochschild-Serre spectral sequence

Hp(Γ,Hq(SKs,𝔾m))Hp+q(S,𝔾m).H^{p}(\Gamma,H^{q}(S_{K^{s}},\mathbb{G}_{m}))\Rightarrow H^{p+q}(S,\mathbb{G}_{m}).

The low degree exact sequence reads

0Pic(S)Pic(SKs)ΓBr(K)ker(Br(S)Br(SKs)Γ)H1(Γ,)0\to\operatorname{Pic}(S)\to\operatorname{Pic}(S_{K^{s}})^{\Gamma}\to\operatorname{Br}(K)\to\ker\left(\operatorname{Br}(S)\to\operatorname{Br}(S_{K^{s}})^{\Gamma}\right)\to H^{1}(\Gamma,{\mathbb{Z}})

By the definition of a Brauer-Severi Variety, we know SKsKs2S_{K^{s}}\cong{\mathbb{P}}^{2}_{K^{s}}. Hence we have Br(SKs)Br(Ks)=0\operatorname{Br}(S_{K^{s}})\cong\operatorname{Br}(K^{s})=0. Since we clearly have H1(Γ,)=0H^{1}(\Gamma,{\mathbb{Z}})=0, it follows that Br(K)Br(S)\operatorname{Br}(K)\to\operatorname{Br}(S) is surjective.

For the second part, consider any aBr(S)[3]a\in\operatorname{Br}(S)[3]. There exists a lift aBr(K)a^{\prime}\in\operatorname{Br}(K) of aa such that 3a0Br(S)3a^{\prime}\mapsto 0\in\operatorname{Br}(S). Therefore, 3aker(Br(K)Br(S))/33a^{\prime}\in\ker\left(\operatorname{Br}(K)\to\operatorname{Br}(S)\right)\cong{\mathbb{Z}}/3. In other words, 9a=0Br(K)9a^{\prime}=0\in\operatorname{Br}(K). Hence, Br(K)[9]\operatorname{Br}(K)[9] surjects onto Br(S)[3]\operatorname{Br}(S)[3]. Notice that by the description of the kernel in Amitsur’s theorem, we clearly have

ker(Br(K)[3]Br(S)[3])/3.\ker\left(\operatorname{Br}(K)[3]\to\operatorname{Br}(S)[3]\right)\cong{\mathbb{Z}}/3.

To compute the cokernel, consider the short exact sequence of trivial Γ\Gamma-modules:

1μ3μ9μ31,1\to\mu_{3}\to\mu_{9}\to\mu_{3}\to 1,

and its associated long exact sequence:

H1(K,μ3)1H2(K,μ3)H2(K,μ9)H2(K,μ3)2H3(K,μ3)\cdots\to H^{1}(K,\mu_{3})\xrightarrow[]{\partial^{1}}H^{2}(K,\mu_{3})\to H^{2}(K,\mu_{9})\to H^{2}(K,\mu_{3})\xrightarrow[]{\partial^{2}}H^{3}(K,\mu_{3})\to\cdots

We claim the boundary maps 1,2\partial^{1},\partial^{2} in the above exact sequence are zero. Indeed, let ω\omega be a primitive third root of unit. By the proof of [TOP17, Lemma 4.1], 1\partial^{1} is given by the cup product with [ω]H1(K,/3)[\omega]\in H^{1}(K,{\mathbb{Z}}/3). By our assumption, ω\omega is a cube in KK, hence [ω]=0H1(K,/3)K/(K)3[\omega]=0\in H^{1}(K,{\mathbb{Z}}/3)\cong K^{*}/(K^{*})^{3}. This shows 1=0\partial^{1}=0.

To show 2=0\partial^{2}=0, consider the following commutative diagram given in [GS06, Lemma 7.5.10]:

μ3K2M(K){\mu_{3}\otimes K^{M}_{2}(K)}K3M(K)/3K3M(K){K^{M}_{3}(K)/3K^{M}_{3}(K)}H2(K,μ33){H^{2}(K,\mu_{3}^{\otimes 3})}H3(K,μ33){H^{3}(K,\mu_{3}^{\otimes 3})}ωhK,32\scriptstyle{\omega\cup h^{2}_{K,3}}{,}\scriptstyle{\{,\}}hK,33\scriptstyle{h^{3}_{K,3}}2~\scriptstyle{\tilde{\partial^{2}}}

Notations in the above diagram are explained in [GS06, Lemma 7.5.10]. Notice since μ3\mu_{3} are trivial Γ\Gamma-modules, we have the following isomorphism ([TOP17, Lemma 4.1]):

ϕji:Hi(K,/3)Hi(K,μ3j)\phi^{i}_{j}:H^{i}(K,{\mathbb{Z}}/3)\cong H^{i}(K,\mu_{3}^{\otimes j})
αα(ωω)j copies\alpha\mapsto\alpha\cup\underbrace{(\omega\cup\cdots\cup\omega)}_{j\text{ copies}}

Since the upper horizontal map in the commutative diagram is given by the symbol product with 0=[ω]H1(K,/3)0=[\omega]\in H^{1}(K,{\mathbb{Z}}/3), it follows that this map is 0. By the Merkurjev-Suslin theorem ([GS06, Theorem 8.6.5]), the left vertical map is surjective, hence 2~=0\tilde{\partial^{2}}=0. Given that cup products "commute" with boundary homomorphisms ([NSW08, Proposition 1.4.3]), we have the commutative diagram involving 2\partial^{2}:

H2(K,μ33){H^{2}(K,\mu_{3}^{\otimes 3})}H3(K,μ33){H^{3}(K,\mu_{3}^{\otimes 3})}H2(K,/3){H^{2}(K,{\mathbb{Z}}/3)}H3(K,/3){H^{3}(K,{\mathbb{Z}}/3)}(ϕ32)1\scriptstyle{(\phi^{2}_{3})^{-1}}2~\scriptstyle{\tilde{\partial^{2}}}(ϕ33)1\scriptstyle{(\phi^{3}_{3})^{-1}}2\scriptstyle{\partial^{2}}

From this, it is clear that 2=0\partial^{2}=0.

Next notice that KK has characteristic 3\neq 3, so we have Br(K)[n]H2(K,μn)\operatorname{Br}(K)[n]\cong H^{2}(K,\mu_{n}) when nn is a power of 3. Consider the following commutative diagram:

0{0}Br(K)[3]{\operatorname{Br}(K)[3]}Br(K)[9]{\operatorname{Br}(K)[9]}Br(K)[3]{\operatorname{Br}(K)[3]}0{0}0{0}Br(S)[3]{\operatorname{Br}(S)[3]}Br(S)[9]{\operatorname{Br}(S)[9]}Br(S)[3]{\operatorname{Br}(S)[3]}σ3\scriptstyle{\sigma_{3}}σ9\scriptstyle{\sigma_{9}}σ3\scriptstyle{\sigma_{3}}

Then the snake lemma gives us

ker(σ9)ϕ1ker(σ3)coker(σ3)ϕ2coker(σ9).\ker(\sigma_{9})\xrightarrow{\phi_{1}}\ker(\sigma_{3})\to\operatorname{coker}(\sigma_{3})\xrightarrow{\phi_{2}}\operatorname{coker}(\sigma_{9}).

Since ϕ1\phi_{1} is multiplication by 3 and ker(σ9)/3\ker(\sigma_{9})\cong{\mathbb{Z}}/3, ϕ1=0\phi_{1}=0. The map ϕ2\phi_{2} is also zero as Br(K)[9]\operatorname{Br}(K)[9] maps onto Br(S)[3]\operatorname{Br}(S)[3]. Hence we have coker(σ3)ker(σ3)/3.\operatorname{coker}(\sigma_{3})\cong\ker(\sigma_{3})\cong{\mathbb{Z}}/3.

3. Brauer groups of Brauer-Severi surface bundles

In this section, we present sufficient conditions under which a Brauer-Severi surface bundle 5-fold is not stably rational. These conditions are derived by generalizing [ABvBP20, Thm. 2.6] to the 3-torsion case. However, the details in these two cases are quite different.

Theorem 3.1.

Let kk be an algebraically closed field of characteristic 3\neq 3 and let π:YB\pi:Y\to B be a Brauer-Severi surface bundle over a smooth projective threefold BB over kk with a smooth generic fiber. Assume Br(B)[3]=0\operatorname{Br}(B)[3]=0 and Hét3(B,/3)=0H^{3}_{\text{\'{e}t}}(B,{\mathbb{Z}}/3)=0. (For example, take B=3B={\mathbb{P}}^{3}.) Let αBr(K)[3]\alpha\in\operatorname{Br}(K)[3] be the Brauer class over K=k(B)K=k(B) corresponding to the generic fiber of π\pi, and it can be represented by a cyclic algebra of index 3. Assume the discriminant locus is good [Def. 2.5] with irreducible components S1,,SnS_{1},\cdots,S_{n}. Further suppose the following conditions also hold:

  1. (1)

    Any irreducible curve in BB is contained in at most two surfaces from the set {S1,,Sn}\left\{S_{1},\cdots,S_{n}\right\}.

  2. (2)

    Through any point of BB, there pass at most three surfaces from the set {S1,,Sn}\left\{S_{1},\cdots,S_{n}\right\}.

  3. (3)

    For all iji\neq j, SiS_{i} and SjS_{j} are factorial at every point of SiSjS_{i}\cap S_{j}.

Let γi=Si2(α)H1(k(Si),/3)\gamma_{i}=\partial^{2}_{S_{i}}(\alpha)\in H^{1}(k(S_{i}),{\mathbb{Z}}/3). Let Γ\Gamma be the subgroup of i=1nH1(k(Si),/3)\bigoplus_{i=1}^{n}H^{1}(k(S_{i}),{\mathbb{Z}}/3) given by Γ=i=1nγi(/3)n\Gamma=\bigoplus_{i=1}^{n}\left<\gamma_{i}\right>\cong({\mathbb{Z}}/3)^{n}. We write elements of Γ\Gamma as (x1,x2,xn)(x_{1},x_{2}\cdots,x_{n}) with xi{0,1,2}x_{i}\in\{0,1,2\}.

Let HΓH\subset\Gamma consist of those elements (x1,,xn)(/3)n(x_{1},\cdots,x_{n})\in({\mathbb{Z}}/3)^{n} such that xi=xjx_{i}=x_{j} whenever there exists an irreducible component CC of SiSjS_{i}\cap S_{j}, such that

(C1(γi),C1(γj))=(1,2)or(2,1).(\partial^{1}_{C}(\gamma_{i}),\partial^{1}_{C}(\gamma_{j}))=(1,2)\ \textup{or}\ (2,1).

Let HHH^{\prime}\subset H be the subgroup consisting of elements (x1,,xn)(/3)n(x_{1},\cdots,x_{n})\in({\mathbb{Z}}/3)^{n} such that xi=xjx_{i}=x_{j} whenever there exists an irreducible components CC of SiSjS_{i}\cap S_{j}, such that

(C1(γi),C1(γj))=(0,0), and γi|C and γj|C are not both trivial in H1(k(C),/3).(\partial^{1}_{C}(\gamma_{i}),\partial^{1}_{C}(\gamma_{j}))=(0,0),\text{ and }\gamma_{i}|_{C}\text{ and }\gamma_{j}|_{C}\text{ are not both trivial in }H^{1}(k(C),{\mathbb{Z}}/3).

Then Hnr2(k(Y)/k,/3)H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3) contains the subquotient H/1,,1H^{\prime}/\left<1,\cdots,1\right>.

Proof.

First, note that under these assumptions, YY is necessarily integral(see 5.7), hence we can talk about its function field k(Y)k(Y). We have the following commutative diagram:

0{0}/3{{\mathbb{Z}}/3}0{0}Hnr2(k(Y)/Y,/3){H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3)}Hnr2(k(Y)/K,/3){H^{2}_{nr}(k(Y)/K,{\mathbb{Z}}/3)}TYB(1)H1(k(T),/3){\displaystyle\bigoplus_{T\in Y^{(1)}_{B}}H^{1}(k(T),{\mathbb{Z}}/3)}Brnr(K)[3]=0{\operatorname{Br}_{nr}(K)[3]=0}H2(K,/3){H^{2}(K,{\mathbb{Z}}/3)}SB(1)H1(k(S),/3){\displaystyle\bigoplus_{S\in B^{(1)}}H^{1}(k(S),{\mathbb{Z}}/3)}CB(2)/3{\displaystyle\bigoplus_{C\in B^{(2)}}{\mathbb{Z}}/3}α{\left<\alpha\right>}Γ{\Gamma}0{0}0{0}T2\scriptstyle{\oplus\partial^{2}_{T}}σ\scriptstyle{\sigma}S2\scriptstyle{\oplus\partial^{2}_{S}}τ\scriptstyle{\tau}C1\scriptstyle{\oplus\partial^{1}_{C}}

We make some observations related to this diagram:

  1. (1)

    By definition, Hnr2(k(Y)/Y,/3)H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3) denotes all those classes in H2(k(Y),/3)H^{2}(k(Y),{\mathbb{Z}}/3) which are unramified with respect to divisorial valuations corresponding to prime divisors on YY, Since the singular locus of YY has codimension 2\geq 2, we can also characterize Hnr2(k(Y)/Y,/3)H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3) as all those classes in H2(k(Y),/3)H^{2}(k(Y),{\mathbb{Z}}/3) that are unramified with respect to divisorial valuations which have centers on Y which are not contained in YsingY_{sing} [ABvBP20, Cor. 2.2].

  2. (2)

    Hnr2(k(Y)/K,/3)H^{2}_{nr}(k(Y)/K,{\mathbb{Z}}/3) denotes those classes in H2(k(Y),/3)H^{2}(k(Y),{\mathbb{Z}}/3) which are killed by residue maps associated to divisorial valuations that are trivial on KK, hence correspond to prime divisors of YY dominating the base BB. We use YB(1)Y_{B}^{(1)} to denote all prime divisors in YY that do not dominate the base BB. Then the upper row is exact by definition.

  3. (3)

    The second row is obtained from Bloch-Ogus complex [BO74], which is exact under the assumptions

    Br(B)[3]=0,andHét3(B,/3)=0.\operatorname{Br}(B)[3]=0,\textup{and}\;H^{3}_{\text{\'{e}t}}(B,{\mathbb{Z}}/3)=0.
  4. (4)

    The left vertical row is exact by Lemma 2.8, because we have

    Hnr2(k(Y)/K,/3)Hnr2(K(S0)/K,/3)Br(S0)[3],H^{2}_{nr}(k(Y)/K,{\mathbb{Z}}/3)\cong H^{2}_{nr}(K(S_{0})/K,{\mathbb{Z}}/3)\cong Br(S_{0})[3],

    where S0S_{0} is the Brauer-Severi surface (over KK) corresponding to the generic fiber α\alpha. Hence S0S_{0} is smooth and we have the last isomorphism in the above statement.

  5. (5)

    In the right vertical row, the map τ\tau is induced by the field extensions k(S)k(T)k(S)\subset k(T), coincides with the induced map

    k(S)×/k(S)×3k(T)×/k(T)×3.k(S)^{\times}/k(S)^{\times 3}\to k(T)^{\times}/k(T)^{\times 3}.

    If SS is not contained in the discriminant locus, the generic fiber of TST\to S is geometrically integral. Then k(S)k(S) is algebraically closed in k(T)k(T), and thus the induced map above is injective. If S=SiS=S_{i} is a component of the discriminant locus, then after taking the base change to the cubic extension F/k(Si)F/k(S_{i}) defined by the residue class γiH1(k(Si),/3)\gamma_{i}\in H^{1}(k(S_{i}),{\mathbb{Z}}/3), the generic fiber of TiSiT_{i}\to S_{i} is a union of three Hirzebruch surfaces 𝔽1\mathbb{F}_{1}, meeting transversally so that any pair of them meet along a fiber of one and a (1)(-1)-curve of the other. (This is correct because TiT_{i} is indeed the preimage of SiS_{i} under π\pi, with the third assumption in Definition 2.5). In this case, the low degree long exact sequence from the Hochschild-Serre spectral sequence

    Hp(Gal(F/k(Si)),Hq(Spec(F),/3))Hp+q(Spec(k(Si)),/3)H^{p}(\operatorname{Gal}(F/k(S_{i})),H^{q}(\operatorname{Spec}(F),{\mathbb{Z}}/3))\Rightarrow H^{p+q}(\operatorname{Spec}(k(S_{i})),{\mathbb{Z}}/3)

    implies the kernel of the natural map H1(k(Si),/3)H1(F,/3)H^{1}(k(S_{i}),{\mathbb{Z}}/3)\to H^{1}(F,{\mathbb{Z}}/3) is generated by γi\gamma_{i}. By the last assumption in Definition 2.5, we know ker(τ)=Γ\ker(\tau)=\Gamma.

Then we can prove that Hnr2(k(Y)/Y,/3)H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3) lies in the image of σ\sigma. In fact, let ξHnr2(k(Y)/Y,/3)\xi\in H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3) and denote by ξ\xi again its the image in Hnr2(k(Y)/Y,/3)H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3). Then ξ\xi is killed by T2\oplus\partial^{2}_{T}. If ξ\xi is not in the image of σ\sigma, it lifts to a class ξH2(K,/9)\xi^{{}^{\prime}}\in H^{2}(K,{\mathbb{Z}}/9) by Lemma 2.8. We have the following exact sequence which is similar to the second row in above diagram with coefficients /9{\mathbb{Z}}/9:

0H2(K,/9)S2SB(1)H1(k(S),/9)C1CB(2)/90\to H^{2}(K,{\mathbb{Z}}/9)\xrightarrow{\oplus\partial^{2}_{S}}\displaystyle\bigoplus_{S\in B^{(1)}}H^{1}(k(S),{\mathbb{Z}}/9)\xrightarrow{\oplus\partial^{1}_{C}}\displaystyle\bigoplus_{C\in B^{(2)}}{\mathbb{Z}}/9\\

Hence at least one residue S2(ξ)\partial^{2}_{S}(\xi^{{}^{\prime}}) must have order 9 (since S2\oplus\partial^{2}_{S} is injective both for 3-torsion and 9-torsion cases ). On the other hand,

ker(SB(1)H1(k(S),/9)TYB(1)H1(k(T),/9))Γ\ker(\bigoplus_{S\in B^{(1)}}H^{1}(k(S),{\mathbb{Z}}/9)\to\bigoplus_{T\in Y_{B}^{(1)}}H^{1}(k(T),{\mathbb{Z}}/9))\cong\Gamma

This is correct because (again we use FF to denote a separable closure of k(Si)k(S_{i})) in the long exact sequence associate to

Hp(Gal(F/k(Si)),Hq(Spec(F),/9))Hp+q(Spec(k(Si)),/9)H^{p}(\operatorname{Gal}(F/k(S_{i})),H^{q}(\operatorname{Spec}(F),{\mathbb{Z}}/9))\Rightarrow H^{p+q}(\operatorname{Spec}(k(S_{i})),{\mathbb{Z}}/9)

We have

0H1(Gal(F/k(Si)),/9)H1(Spec(k(Si)),/9)H1(Spec(F),/9)0\to H^{1}(Gal(F/k(S_{i})),{\mathbb{Z}}/9)\to H^{1}(\operatorname{Spec}(k(S_{i})),{\mathbb{Z}}/9)\to H^{1}(\operatorname{Spec}(F),{\mathbb{Z}}/9)

As we can calculate ètale cohomology of spectrum of a field using Galois cohomology, we have the kernel:

H1(Gal(F/k(Si)),/9)Homcont(/3,/9)/3H^{1}(Gal(F/k(S_{i})),{\mathbb{Z}}/9)\cong Hom_{cont}({\mathbb{Z}}/3,{\mathbb{Z}}/9)\cong{\mathbb{Z}}/3

While the kernel for those SS doesn’t belongs to the discriminant locus is clearly zero by the same argument in 3-torsion case.

Now we notice that Γ\Gamma has no elements of order 9, this means S2(ξ)\partial^{2}_{S}(\xi^{{}^{\prime}}) can’t be mapped to 0 in TYB(1)H1(k(T),/3)\bigoplus_{T\in Y^{(1)}_{B}}H^{1}(k(T),{\mathbb{Z}}/3), hence a contradiction.

The above diagram chasing in fact gives us

Hnr2(k(Y)/Y,/3)Γker(C1)/αH/α.H^{2}_{nr}(k(Y)/Y,{\mathbb{Z}}/3)\cong\Gamma\cap\ker(\oplus\partial^{1}_{C})/\left<\alpha\right>\cong H/\left<\alpha\right>.

Next we determine classes in HH that are in Hnr2(k(Y)/k,/3)H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3). In particular, we show that the subgroup HH^{\prime} defined earlier is contained in Hnr2(k(Y)/k,/3)H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3). We do this by checking whether the classes in HH^{\prime} are unramified with respect to all divisorial valuations μ\mu of k(Y)k(Y) (and not just those that come from prime divisors on YY). Consider a class βH\beta\in H, viewed as an element in H2(K,/3)H^{2}(K,{\mathbb{Z}}/3). Denote by β\beta^{\prime} the image of β\beta in H2(k(Y),/3)H^{2}(k(Y),{\mathbb{Z}}/3). We aim to show that β\beta^{\prime} is unramified on YY if β\beta is in HH^{\prime}. Using the definition of HH, it is sufficient to check this for valuations whose centers on BB has codimension at least 1. In the following, we use 𝒪\mathscr{O} to denote the local ring of μ\mu in BB.

  1. Case 1:

    The center of μ\mu on BB is not contained in the discriminant locus: In this case, for any surface SS passing through the center of μ\mu, we have

    S2(β)=0.\partial^{2}_{S}(\beta)=0.

    Then [ABvBP20, Proposition 2.1] tells us β\beta is in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). Hence β=σ(β)\beta^{\prime}=\sigma(\beta) is also unramified with respect to μ\mu in this case.

  2. Case 2:

    The center of μ\mu on BB is contained in the discriminant locus, but not in the intersection of two or more components: Now the center is contained in SiS_{i} for a unique ii. Recall that the ithi^{\rm th} component xix_{i} of S2(β)\oplus\partial^{2}_{S}(\beta) is 0,10,1 or 22. If xi=0x_{i}=0, by an argument same as Case 1, β\beta is in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). Similarly, if xi=1x_{i}=1, βα\beta-\alpha is in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). Finally, if xi=2x_{i}=2, β2α\beta-2\alpha is in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). Notice that

    β=σ(β)=σ(βα)=σ(β2α).\beta^{\prime}=\sigma(\beta)=\sigma(\beta-\alpha)=\sigma(\beta-2\alpha).

    So in all three conditions, we have β\beta^{\prime} is unramified with respect to μ\mu in this case.

  3. Case 3:

    The center μ\mu on BB is a curve CC that is an irreducible component of SiSjS_{i}\cap S_{j}: In this case, we again check the possible values of xix_{i} and xjx_{j} in S2(β)\oplus\partial^{2}_{S}(\beta). We have the following cases:

    • Case 3(a):

      If xi=xjx_{i}=x_{j}, then the argument in Case 2 above gives us that at least one of β,βα\beta,\beta-\alpha or β2α\beta-2\alpha lies in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). So we are done in this situation.

    • Case 3(b):

      (xi,xj)=(0,1)(x_{i},x_{j})=(0,1) or (1,0)(1,0): By symmetry, we can assume (xi,xj)=(1,0)(x_{i},x_{j})=(1,0). Notice that

      3|(C1(γi)+C1(γj))3\left|\left(\partial^{1}_{C}(\gamma_{i})+\partial^{1}_{C}(\gamma_{j})\right)\right.

      by the exactness of the second row in the diagram. Then we must have

      C1(γi)=C1(γj)=0\partial^{1}_{C}(\gamma_{i})=\partial^{1}_{C}(\gamma_{j})=0

      This means that a rational function representing the class

      γiH1(k(Si),/3)=k(Si)×/k(Si)×3\gamma_{i}\in H^{1}(k(S_{i}),{\mathbb{Z}}/3)=k(S_{i})^{\times}/k(S_{i})^{\times 3}

      has a zero or a pole of order divisible by 3 along CC. Without loss of generality, we may assume that the function associated with γi\gamma_{i} is contained in the local ring 𝒪Si,C\mathscr{O}_{S_{i},C} of CC in SiS_{i}.We call this function fγif_{\gamma_{i}}. Let tt be a local parameter for CC in 𝒪Si,C\mathscr{O}_{S_{i},C}. Such a local parameter exists as CC is a Cartier divisor on SiS_{i}, which in turn follows since SiS_{i} is assumed to be factorial along CC. Then fγitμC(fγi)\displaystyle\frac{f_{\gamma_{i}}}{t^{\mu_{C}(f_{\gamma_{i}})}} is a unit, and hence any preimage in 𝒪\mathscr{O} is also a unit (See Remark 3.2 below). Call such a preimage uγiu_{\gamma_{i}}, which may be viewed as a rational function in KK. Assume πSi\pi_{S_{i}} is a local parameter of SiS_{i} in 𝒪\mathscr{O}. Consider the symbol algebra (uγi,πSi)H2(K,/3)(u_{\gamma_{i}},\pi_{S_{i}})\in H^{2}(K,{\mathbb{Z}}/3). Let SS be a surface containing CC. By lemma 2.1, we have

      S2(uγi,πSi)=(1)μS(uγi)μS(πSi)u¯γiμS(πSi)πSiμS(uγi)={u¯γiifS=Si0ifSSi\displaystyle\partial^{2}_{S}(u_{\gamma_{i}},\pi_{S_{i}})=(-1)^{\mu_{S}(u_{\gamma_{i}})\mu_{S}(\pi_{S_{i}})}{\frac{\bar{u}_{\gamma_{i}}^{\mu_{S}(\pi_{S_{i}})}}{\pi_{S_{i}}^{\mu_{S}(u_{\gamma_{i}})}}}=\left\{\begin{array}[]{cl}\bar{u}_{\gamma_{i}}&\text{if}\ S=S_{i}\\ 0&\text{if}\ S\neq S_{i}\\ \end{array}\right.

      On the other hand, u¯γi=γi\bar{u}_{\gamma_{i}}=\gamma_{i} by construction, so we have

      Si2(uγi,πSi)=γi=Si2(β)\partial^{2}_{S_{i}}(u_{\gamma_{i}},\pi_{S_{i}})=\gamma_{i}=\partial^{2}_{S_{i}}(\beta)
      Sj2(uγi,πSi)=0=Sj2(β)\partial^{2}_{S_{j}}(u_{\gamma_{i}},\pi_{S_{i}})=0=\partial^{2}_{S_{j}}(\beta)

      Also S2(β)=0\partial^{2}_{S}(\beta)=0 if SS is a surface passing through CC other than SiS_{i} and SjS_{j}.(In fact, by our assumption, such an SS is not in the discriminant locus and so this agrees with this conclusion.) Hence [ABvBP20, Proposition 2.1] tells us β(uγi,πSi)\beta-(u_{\gamma_{i}},\pi_{S_{i}}) is in the image of Hét2(𝒪,/3)H^{2}_{\text{\'{e}t}}(\mathscr{O},{\mathbb{Z}}/3). Hence

      μ2(σ(β(uγi,πSi)))=0\partial^{2}_{\mu}(\sigma(\beta-(u_{\gamma_{i}},\pi_{S_{i}})))=0

      It then suffices to show that

      μ2(σ((uγi,πSi)))=±u¯γiμ(πSi)=0H1(k(μ),/3)\partial^{2}_{\mu}(\sigma((u_{\gamma_{i}},\pi_{S_{i}})))=\pm\bar{u}_{\gamma_{i}}^{\mu(\pi_{S_{i}})}=0\in H^{1}(k(\mu),{\mathbb{Z}}/3)

      By assumption, u¯γi|C\bar{u}_{\gamma_{i}}|_{C} is trivial, hence so is u¯γiμ(πSi)\bar{u}_{\gamma_{i}}^{\mu(\pi_{S_{i}})} as the center of μ\mu is CC.

    • Case 3(c):

      (xi,xj)=(0,2)(x_{i},x_{j})=(0,2) or (2,0)(2,0): By symmetry, we can assume (xi,xj)=(2,0)(x_{i},x_{j})=(2,0). Now the proof is essentially same as Case 3(b), which shows that

      μ2(σ(β2(uγi,πSi)))=0.\partial^{2}_{\mu}(\sigma(\beta-2(u_{\gamma_{i}},\pi_{S_{i}})))=0.

      It follows that μ2(σ(β))=0.\partial^{2}_{\mu}(\sigma(\beta))=0. and so β\beta^{\prime} is unramified along μ\mu.

    • Case 3(d):

      (xi,xj)=(1,2)or(2,1)(x_{i},x_{j})=(1,2)\ \text{or}\ (2,1): Assume (xi,xj)=(2,1)(x_{i},x_{j})=(2,1). Then applying Case 3(b)3(b) to the class βα\beta-\alpha, we see that this case is also proved.

  4. Case 4:

    The center of μ\mu on BB is a point PCP\in C, here CC is as in case 3, and Si,SjS_{i},S_{j} are the only surfaces among the S1,,SnS_{1},\cdots,S_{n} that pass through PP. As we have seen in the discussion of Case 3, we can reduce to the case when (xi,xj)=(1,0)(x_{i},x_{j})=(1,0). Hence we again have C1(γi)=C1(γj)=0\partial^{1}_{C}(\gamma_{i})=\partial^{1}_{C}(\gamma_{j})=0. In fact, this is true for any curve CC^{\prime} that contains PP and is contained in SiSjS_{i}\cup S_{j}. Choose a function fγik(Si)f_{\gamma_{i}}\in k(S_{i}) representing the class γi\gamma_{i}. Then let C1,,CNC_{1},\cdots,C_{N} be all irreducible curves through PP that are either a zero or a pole for the function fγif_{\gamma_{i}}. Pick local equations tt_{\ell} of CC_{\ell} in 𝒪Si,P\mathscr{O}_{S_{i},P}, and consider the following rational function on SiS_{i}:

    fγi(t1μC1(fγi)tNμCN(fγi)).\frac{f_{\gamma_{i}}}{\left(t_{1}^{\mu_{C_{1}}(f_{\gamma_{i}})}\cdots t_{N}^{\mu_{C_{N}}(f_{\gamma_{i}})}\right)}.

    Since SiS_{i} is assumed to be factorial, in particular, normal at PP, the above rational function is a unit locally around PP. Hence it can be lifted to a unit in 𝒪\mathscr{O}. Then we can repeat the rest of the proof as in Case 3(b). (Notice that every element in k(P)k(P) is a cube since kk is algebraically closed, so the last step of Case 3(b) is automatically true.)

  5. Case 5:

    The center of μ\mu on BB is a point PP that lies on exactly three distinct surfaces Si,Sj,SlS_{i},S_{j},S_{l}: We consider the possible values of (xi,xj,xl)(x_{i},x_{j},x_{l}). If xi=xj=xlx_{i}=x_{j}=x_{l}, then one of β,βα\beta,\beta-\alpha or β2α\beta-2\alpha is unramified. By symmetry and up to subtraction by α\alpha or 2α2\alpha, the only remaining cases are (1,0,0)(1,0,0), (1,1,0)(1,1,0), and (2,1,0)(2,1,0). Notice that (2,1,0)=(1,1,0)+(1,0,0)(2,1,0)=(1,1,0)+(1,0,0), and that the case (1,1,0)(1,1,0) is equivalent to the case (2,0,0)(2,0,0). Hence, we only need to consider the case (1,0,0)(1,0,0), which is same as Case 4. Now the rest of the proof is same as in Case 4.

Remark 3.2.

In Case 3(b) in Theorem 3.1, we claimed that if x¯𝒪Si,C\bar{x}\in\mathscr{O}_{S_{i},C} is a unit, then any preimage xx in 𝒪=𝒪B,C\mathscr{O}=\mathscr{O}_{B,C} is also a unit. In fact, we have

𝒪Si,C𝒪/(πSi).\mathscr{O}_{S_{i},C}\cong\mathscr{O}/(\pi_{S_{i}}).

As x¯\bar{x} is a unit in 𝒪Si,C\mathscr{O}_{S_{i},C}, there exist a y¯𝒪Si,C\bar{y}\in\mathscr{O}_{S_{i},C} such that x¯y¯=1𝒪Si,C\bar{x}\bar{y}=1\in\mathscr{O}_{S_{i},C}. Hence there exist t𝒪t\in\mathscr{O}, such that

xy=1+πSit𝒪xy=1+\pi_{S_{i}}t\in\mathscr{O}

Notice that πSit\pi_{S_{i}}t is contained in the maximal ideal of 𝒪\mathscr{O}, so 1+πSit1+\pi_{S_{i}}t is a unit in 𝒪\mathscr{O}. Hence any preimage xx is also a unit in 𝒪\mathscr{O}.

We prove an immediate corollary in which we weaken the hypothesis about factoriality when n=2n=2. In this case, the discriminant locus has exactly two irreducible components. We prove that it is sufficient to have only one of them factorial at their intersection to make the unramified Brauer group nontrivial:

Corollary 3.3.

Assume n=2n=2. We continue with the same hypothesis as in the theorem except the following change: we replace the requirement (3)(3) by the following:
(3’) S1S_{1} is factorial at every point of S1S2S_{1}\cap S_{2} .
Then Hnr2(k(Y)/k,/3)H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3) is nontrivial and hence YY is not stably rational.

Proof.

In this case, the Brauer class β\beta in HH^{\prime} whose representative is (1,0)(1,0) can be lifted to a nontrivial unramified Brauer class in H2(k(Y)/k,/3)H^{2}(k(Y)/k,{\mathbb{Z}}/3)

4. Example

In this section, we will construct a Brauer-Severi surface bundle over 3{\mathbb{P}}^{3} that is stably non-rational. We use Corollary 3.3 for this purpose.

Example 4.1.

Consider the following two surfaces in 3=Proj[x0,x1,x2,x3]{\mathbb{P}}^{3}_{{\mathbb{C}}}={\rm Proj}\ {\mathbb{C}}[x_{0},x_{1},x_{2},x_{3}]:

S1:{x09+(x13x23)(x23x33)(x33x13)=0}S_{1}:\{x_{0}^{9}+(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-x_{3}^{3})(x_{3}^{3}-x_{1}^{3})=0\}
S2:{(x09+(x13x23)(x23x33)(x33x13))(x09x13x23x33)+x16x26x36=0}S_{2}:\{\left(x_{0}^{9}+(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-x_{3}^{3})(x_{3}^{3}-x_{1}^{3})\right)\left(x_{0}^{9}-x_{1}^{3}x_{2}^{3}x_{3}^{3}\right)+x_{1}^{6}x_{2}^{6}x_{3}^{6}=0\}

In the following, we use FS1,FS2F_{S_{1}},F_{S_{2}} to denote the equation defines S1S_{1}, S2S_{2} separately. We start by checking that both S1S_{1} and S2S_{2} are irreducible and reduced:

  • S1S_{1} is irreducible and reduced. This follows directly from the fact that the singular locus of S1S_{1} has dimension 0. In fact, S1S_{1} only singular at 12 isolated points:

    [0:1:0:0],[0:0:1:0],[0:0:0:1][0:1:0:0]\ ,\ [0:0:1:0]\ ,\ [0:0:0:1]
    [0:ω:1:1],[0:1:ω:1],[0:1:1:ω][0:\omega:1:1]\ ,\ [0:1:\omega:1]\ ,\ [0:1:1:\omega]
    [0:ω2:1:1],[0:1:ω2:1],[0:1:1:ω2][0:\omega^{2}:1:1],[0:1:\omega^{2}:1],\ [0:1:1:\omega^{2}]
    [0:ω2:ω:1],[0:ω:ω2:1],[0:1:1:1][0:\omega^{2}:\omega:1]\ ,[0:\omega:\omega^{2}:1]\ ,[0:1:1:1]

    Here ω\omega is a primitive 3rd3^{rd} roots of unity. If S1S_{1} is not reduced, then the singular locus would have dimension 2. If S1S_{1} is not irreducible, the singular locus would have dimension at least 1 by Bèzout theorem.

  • S2S_{2} is irreducible and reduced. We may rewrite the equation defining S2S_{2} as:

    x018+P(x1,x2,x3)x09x13x23x33P(x1,x2,x3)x_{0}^{18}+P(x_{1},x_{2},x_{3})x_{0}^{9}-x_{1}^{3}x_{2}^{3}x_{3}^{3}P(x_{1},x_{2},x_{3})

    where P(x1,x2,x3)=(x13x23)(x23x33)(x33x13)x13x23x33P(x_{1},x_{2},x_{3})=(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-x_{3}^{3})(x_{3}^{3}-x_{1}^{3})-x_{1}^{3}x_{2}^{3}x_{3}^{3}. We may consider the above polynomial as an element in [x0,x1,x2,x3]=[x1,x2,x3][x0]{\mathbb{C}}[x_{0},x_{1},x_{2},x_{3}]={\mathbb{C}}[x_{1},x_{2},x_{3}][x_{0}], which is a UFD. Hence to check it is irreducible, it is sufficient to use Eisenstein’s criterion: We need to find a prime ideal 𝔭\mathfrak{p} in [x1,x2,x3][x0]{\mathbb{C}}[x_{1},x_{2},x_{3}][x_{0}], such that

    P(x1,x2,x3)𝔭,P(x_{1},x_{2},x_{3})\in\mathfrak{p},
    x13x23x33P(x1,x2,x3)𝔭andx_{1}^{3}x_{2}^{3}x_{3}^{3}P(x_{1},x_{2},x_{3})\in\mathfrak{p}\ \text{and}
    x13x23x33P(x1,x2,x3)𝔭2.x_{1}^{3}x_{2}^{3}x_{3}^{3}P(x_{1},x_{2},x_{3})\notin\mathfrak{p}^{2}.

    It is evident that an appropriate prime ideal exists if P(x1,x2,x3)P(x_{1},x_{2},x_{3}) has an irreducible factor with multiplicity 1 and is coprime to x1x2x3x_{1}x_{2}x_{3}. In fact, any irreducible factor of P(x1,x2,x3)P(x_{1},x_{2},x_{3}) is inherently coprime to x1x2x3x_{1}x_{2}x_{3}. Therefore it suffices to provide a single regular point of P(x1,x2,x3)P(x_{1},x_{2},x_{3}) to show the existence of such an irreducible factor. Finally, we directly check that (1,1,0)(1,1,0) is a regular point of P(x1,x2,x3)P(x_{1},x_{2},x_{3}). Hence S2S_{2} is irreducible.

    Now as we have already shown S2S_{2} is irreducible, it is sufficient to find a smooth point in S2S_{2} to show it is reduced. Indeed, one can easily check that [(56)19:1:2:0][(-56)^{\frac{1}{9}}:1:2:0] is indeed a smooth point of S2S_{2}.

We choose rational triple covers of S1S_{1} and S2S_{2} defined by:

γ1=x23x33x03¯H1((S1),/3)(S1)×/(S1)×3\gamma_{1}=\overline{\frac{x_{2}^{3}-x_{3}^{3}}{x_{0}^{3}}}\in H^{1}({\mathbb{C}}(S_{1}),{\mathbb{Z}}/3)\cong{\mathbb{C}}(S_{1})^{\times}/{\mathbb{C}}(S_{1})^{\times 3}
γ2=x09x13x23x33x09¯H1((S2),/3)(S2)×/(S2)×3\gamma_{2}=\overline{\frac{x_{0}^{9}-x_{1}^{3}x_{2}^{3}x_{3}^{3}}{x_{0}^{9}}}\in H^{1}({\mathbb{C}}(S_{2}),{\mathbb{Z}}/3)\cong{\mathbb{C}}(S_{2})^{\times}/{\mathbb{C}}(S_{2})^{\times 3}

We claim the triple covers γ1,γ2\gamma_{1},\gamma_{2} are not trivial: In fact, by Lemma 2.1, the residue of γ1\gamma_{1} of a valuation centered at the point [0:ω:1:1][0:\omega:1:1] is 1/31\in{\mathbb{Z}}/3. Hence γ1\gamma_{1} is not trivial. To show γ2\gamma_{2} is not trivial is equivalent to show FS1F_{S_{1}} is not a cubic in the function field of S2S_{2}. And it’s true because the residue of FS1x09¯\overline{\frac{F_{S_{1}}}{x_{0}^{9}}} of a valuation centered at the point [0:0:1:1][0:0:1:1] is 1/31\in{\mathbb{Z}}/3. Hence γ1,γ2\gamma_{1},\gamma_{2} are not trivial.

Consider the corresponding Bloch-Ogus exact sequence:

0{0}Br((3))[3]{Br({\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}}))[3]}S(3)(1)H1(k(S),/3){\displaystyle\bigoplus_{S\in({\mathbb{P}}^{3}_{{\mathbb{C}}})^{(1)}}H^{1}(k(S),{\mathbb{Z}}/3)}C(3)(2)H0(k(C),/3){\displaystyle\bigoplus_{C\in({\mathbb{P}}^{3}_{{\mathbb{C}}})^{(2)}}H^{0}(k(C),{\mathbb{Z}}/3)}S2\scriptstyle{\oplus\partial^{2}_{S}}C1\scriptstyle{\oplus\partial^{1}_{C}}

We have C1(γ1)=C1(γ2)=0\oplus\partial^{1}_{C}(\gamma_{1})=\oplus\partial^{1}_{C}(\gamma_{2})=0. In fact, it is easy to check that for any curve CC such that γ1\gamma_{1}(or γ2\gamma_{2}) has a zero or pole along CC, the order is divided by 3. Hence

(1,,1,γ1,1,,1,γ2,1,)S(3)(1)H1(k(S),/3)(1,\cdots,1,\gamma_{1},1,\cdots,1,\gamma_{2},1,\cdots)\in\displaystyle\bigoplus_{S\in({\mathbb{P}}^{3}_{{\mathbb{C}}})^{(1)}}H^{1}(k(S),{\mathbb{Z}}/3)

can be lifted to a Brauer class [𝒜]=[(FS2(x23x33)x021,FS1x09)ω]Br((3))[3][\mathscr{A}]=[(\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x_{0}^{21}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}]\in\operatorname{Br}({\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}}))[3]. This can be directly checked by Lemma 2.1.

By Theorem 5.5, the cyclic algebra 𝒜\mathscr{A} gives out a Brauer-Severi surface bundle Y3Y\to{\mathbb{P}}^{3}_{{\mathbb{C}}}. This Brauer-Severi surface bundle has a good discriminant locus. We prove this by checking the conditions in Definition 2.5. Here is the list of corresponding arguments:

  1. (1)

    We already proved that S1S_{1} and S2S_{2} are reduced.

  2. (2)

    The behavior of a general fiber over S1S_{1} and S2S_{2} is given by [Mae97, Thm. 2.1].

  3. (3)

    The induced triple cover over S1S_{1} and S2S_{2} are irreducible because γ1\gamma_{1}, γ2\gamma_{2} are not trivial.

  4. (4)

    To show the last requirement in Definition 2.5 is true, we have the following commutative diagram:

    F×/F×3{F^{\times}/F^{\times 3}}F(u,v)×/F(u,v)×3{F(u,v)^{\times}/F(u,v)^{\times 3}}k(Si)×/k(Si)×3{k(S_{i})^{\times}/k(S_{i})^{\times 3}}k(Ti)×/k(Ti)×3{k(T_{i})^{\times}/k(T_{i})^{\times 3}}a\scriptstyle{a}b\scriptstyle{b}τi\scriptstyle{\tau_{i}}d\scriptstyle{d}

    Where TiT_{i} is defined right after the large diagram in Theorem 3.1. For S1S_{1}, bb is induced by the cubic extension defined by γ1\gamma_{1}, dd is induced by the cubic extension defined by FS2(x23x33)x021\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x^{21}_{0}}, which is equal to the cubic class defined by γ1\gamma_{1} ([Art82, Thm. 2.1]. ) Note that aa is injective, and ker(b)=<γ1>\ker(b)=<\gamma_{1}>. On the other hand, an easy diagram chasing as in part (5) in proof of Theorem 3.1 shows that ker(τ1)\ker(\tau_{1}) contains <γ1><\gamma_{1}> . This forces dd to be injective and ker(τ1)=<γ1>\ker(\tau_{1})=<\gamma_{1}>. Same argument works for S2S_{2}.

On the other hand, we list all irreducible components of S1S2S_{1}\cap S_{2}:

D=6D1+6D2+6D3D=6D_{1}+6D_{2}+6D_{3}

Where

D1={x1=0,x09x26x33+x23x36=0}D_{1}=\{x_{1}=0,x_{0}^{9}-x_{2}^{6}x_{3}^{3}+x_{2}^{3}x_{3}^{6}=0\}
D2={x2=0,x09x36x13+x33x16=0}D_{2}=\{x_{2}=0,x_{0}^{9}-x_{3}^{6}x_{1}^{3}+x_{3}^{3}x_{1}^{6}=0\}
D3={x3=0,x09x16x23+x13x26=0}D_{3}=\{x_{3}=0,x_{0}^{9}-x_{1}^{6}x_{2}^{3}+x_{1}^{3}x_{2}^{6}=0\}

One can easily check they are indeed irreducible using Eisenstein’s criterion by viewing those polynomials as elements in [x1,x2,x3][x0]{\mathbb{C}}[x_{1},x_{2},x_{3}][x_{0}]. Notice that D1D_{1} passes through only two singular points of S1S_{1}: [0:0:1:0][0:0:1:0] and [0:0:0:1][0:0:0:1]. It is straightforward to check D1D_{1} is indeed a Cartier divisor of S1S_{1}, even along these two singular points:

Lemma 4.2.

DiD_{i} are Cartier divisors of S1S_{1}.

Proof.

By symmetry, it is sufficient to check the behavior of D1D_{1} at singular points of S1S_{1}. Notice that D1D_{1} only passes through two singular points of S1S_{1}: [0:0:1:0][0:0:1:0] and [0:0:0:1][0:0:0:1]. Let P=[0:0:0:1]P=[0:0:0:1], then we have the local ring:

𝒪S1,P=([x0,x1,x2]/(x09+(x13x23)(x231)(1x13)))(x0,x1,x2)\mathscr{O}_{S_{1},P}=({\mathbb{C}}[x_{0},x_{1},x_{2}]/(x_{0}^{9}+(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-1)(1-x_{1}^{3})))_{(x_{0},x_{1},x_{2})}

By expanding the equation defining S1S_{1}, we have

x09x26+x23=x13(x13x23x26x13+1)𝒪S1,Px_{0}^{9}-x_{2}^{6}+x_{2}^{3}=x_{1}^{3}(x_{1}^{3}x_{2}^{3}-x_{2}^{6}-x_{1}^{3}+1)\in\mathscr{O}_{S_{1},P}

Notice x13x23x26x13+1x_{1}^{3}x_{2}^{3}-x_{2}^{6}-x_{1}^{3}+1 is a unit in 𝒪S1,P\mathscr{O}_{S_{1},P}, hence the ideal defining D1D_{1}, which is (x1,x09x26+x23)(x_{1},x_{0}^{9}-x_{2}^{6}+x_{2}^{3}), is generated by one element x1x_{1}. Similarly one can do the calculation for the point [0:0:1:0][0:0:1:0]. As a result, D1D_{1} is a Cartier divisor of S1S_{1}.

Finally, we need to check both γ1|Di\gamma_{1}|_{D_{i}} and γ2|Di\gamma_{2}|_{D_{i}} are trivial for i{1,2,3}i\in\{1,2,3\}. These are directly following from the choices of γ1\gamma_{1} and γ2\gamma_{2}. Hence in this example, using notations in Theorem 3.1, we have H=H=Γ=/3×/3H^{\prime}=H=\Gamma={\mathbb{Z}}/3\times{\mathbb{Z}}/3. By Corollary 3.3, the unramified Brauer group of YY contains a subgroup /3{\mathbb{Z}}/3, hence YY is not stably rational.

5. Flatness

In this section, we check the cyclic algebra

𝒜=(FS2(x23x33)x021,FS1x09)ω\mathscr{A}=(\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x_{0}^{21}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}

indeed gives us a Brauer-Severi surface bundle over 3{\mathbb{P}}^{3}_{{\mathbb{C}}} as in Definition 2.4. We keep the notation in Example 4.1 through out this section. The definition of a general Brauer-Severi scheme is given by Van den Bergh in [VdB88]. In [See99], Seelinger gave an alternating description of Brauer-Severi scheme which is easier to use in our case. See also Section 1 in [Mae97] for the discussion of the following definitions:

Definition 5.1.

Let Λ\Lambda be a sheaf of 𝒪3\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}} algebra that is torsion free and coherent as an 𝒪3\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}} module. We say Λ\Lambda is an 𝒪3\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}-order in 𝒜\mathscr{A} if Λ\Lambda contains 𝒪3\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}} and

Λ𝒪3(3)𝒜\Lambda\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}}{\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}})\cong\mathscr{A}
Definition 5.2.

For each point p3p\in{\mathbb{P}}^{3}_{{\mathbb{C}}}, let 𝒪3,p\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p} denote the regular local ring of 3{\mathbb{P}}^{3}_{{\mathbb{C}}} at pp. We say a finitely generated 𝒪3,p\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p} algebra Λp\Lambda_{p} is an 𝒪3,p\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}-order in 𝒜\mathscr{A}, if Λp\Lambda_{p} is torsion free and

Λp𝒪3,p(3)𝒜\Lambda_{p}\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}}{\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}})\cong\mathscr{A}
Remark 5.3.

In this paper, we always assume an order is locally free.

Recall that (3.4)(3.4) of [Mae97] describes an 𝒪3\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}-order which we again denote by Λ\Lambda in the following, we denote its localization at a point pp by Λp\Lambda_{p}.

Definition 5.4.

Let VΛV_{\Lambda} (respectively, VΛpV_{\Lambda_{p}}) be the functor from the category of 3{\mathbb{P}}^{3}_{{\mathbb{C}}}-schemes (respectively, Spec(𝒪3,p)\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p})-schemes) to the category of sets:

VΛ(S)={[z]Gn[(Λ𝒪3S)]|zu=NS(u)z,u(Λ𝒪3S)}V_{\Lambda}(S)=\{[z]\in G_{n}[(\Lambda\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}}S)^{\vee}]\ |\ z\cdot u=N_{S}(u)z\ ,\ \forall u\in(\Lambda\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}}S)^{*}\}
VΛp(S)={[z]Gn[(Λp𝒪3,pS)]|zu=NS(u)z,u(Λ𝒪3,pS)}V_{\Lambda_{p}}(S)=\{[z]\in G_{n}[(\Lambda_{p}\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}}S)^{\vee}]\ |\ z\cdot u=N_{S}(u)z\ ,\ \forall u\in(\Lambda\otimes_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}}S)^{*}\}

where \vee denotes the dual sheaf, * denotes the unit group , NSN_{S} is the reduced norm and GnG_{n} denotes the functor of Grassmannian of nn-quotients([VdB88, Def.1]). These functors are represented by schemes as these are closed subschemes of the Grassmannian, which we call the Brauer-Severi scheme (associated to Λ\Lambda, Λp\Lambda_{p}) and again denote them by VΛV_{\Lambda}, VΛpV_{\Lambda_{p}}.

Theorem 5.5.

Y=VΛY=V_{\Lambda} is a Brauer-Severi surface bundle over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}.

Proof.

According to Definition 5.4, for every closed point pp in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, we have the following commutative diagram of schemes:

VΛ×3Spec(𝒪3,p)VΛp{V_{\Lambda}\times_{{\mathbb{P}}^{3}_{{\mathbb{C}}}}\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p})\cong V_{\Lambda_{p}}}VΛ{V_{\Lambda}}Spec(𝒪3,p){\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p})}3{{\mathbb{P}}^{3}_{{\mathbb{C}}}}πp\scriptstyle{\pi_{p}}π\scriptstyle{\pi}

We first show π\pi is a flat morphism. In order to do so, it suffices to show πp\pi_{p} is flat for all closed points p3p\in{\mathbb{P}}^{3}_{{\mathbb{C}}}. Indeed, if this is done, the flat locus of π\pi would be an open subset of 3{\mathbb{P}}^{3}_{{\mathbb{C}}} containing all closed points, hence is equal to 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. Furthermore, by the "Miracle flatness" theorem [Sta23] and the fact that Spec(𝒪3,p)\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}) is regular, it suffices to show each VΛpV_{\Lambda_{p}} is Cohen-Macaulay and each fiber of πp\pi_{p} has the same dimension. We do this by a case-by-case argument for all closed points in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}:

  1. Case 1:

    pS1S2p\notin S_{1}\cup S_{2}. It is well know that Λ\Lambda is an Azumaya algebra outside of discriminant locus [Art82]. All fibers of πp\pi_{p} are smooth Brauer-Severi surfaces and furthermore VΛpV_{\Lambda_{p}} is regular, hence Cohen-Macaulay. By the "Miracle flatness" theorem, πp\pi_{p} is flat in this case.

  2. Case 2:

    pS1S2p\in S_{1}\cup S_{2} and pS1S2p\notin S_{1}\cap S_{2} and pS1{x23x33=0}p\notin S_{1}\cap\{x_{2}^{3}-x_{3}^{3}=0\} . Following ideas from Artin [Art82], we may write Λp\Lambda_{p} as the symbol algebra (fp,gp)ω(f_{p},g_{p})_{\omega}. That is, Λp\Lambda_{p} over Spec(𝒪3,p)\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}) is generated by x,yx,y subject to the relations

    x3=fp,y3=gp,xy=ωyx.x^{3}=f_{p},y^{3}=g_{p},xy=\omega yx.

    Since pS1{x23x33=0}p\notin S_{1}\cap\{x_{2}^{3}-x_{3}^{3}=0\}, it follows that fpf_{p} is a unit in 𝒪3,p\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}. Let Rp=𝒪3,p[T]/(T3fp)R_{p}=\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}[T]/(T^{3}-f_{p}), then

    Spec(Rp)𝜏Spec(𝒪3,p)\operatorname{Spec}(R_{p})\xrightarrow[]{\tau}\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p})

    is an étale neighborhood of Spec(𝒪3,p)\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}) with τ\tau faithfully flat as it surjects on the underlying topological space. By faithfully flat descent, it suffices to show VΛpRpV_{\Lambda_{p}}\otimes R_{p} is flat over Spec(Rp)\operatorname{Spec}(R_{p}). In [Art82], Artin noticed VΛpRpV_{\Lambda_{p}}\otimes R_{p} can be viewed as a subalgebra of the 3 by 3 matrices algebra over RpR_{p} by setting

    x=[T000Tω000Tω2],y=[010001gp00]x=\begin{bmatrix}T&0&0\\ 0&T\omega&0\\ 0&0&T\omega^{2}\end{bmatrix},y=\begin{bmatrix}0&1&0\\ 0&0&1\\ g_{p}&0&0\end{bmatrix}

    And VΛpRpV_{\Lambda_{p}}\otimes R_{p} can be embedded into Rp2×Rp2×Rp2{\mathbb{P}}^{2}_{R_{p}}\times{\mathbb{P}}^{2}_{R_{p}}\times{\mathbb{P}}^{2}_{R_{p}} by the following 9 equations with a cyclic permutations in indices:

    gpξ11ξ22=ξ12ξ21g_{p}\xi_{11}\xi_{22}=\xi_{12}\xi_{21}
    gpξ11ξ23=ξ13ξ21g_{p}\xi_{11}\xi_{23}=\xi_{13}\xi_{21}
    gpξ11ξ32=gpξ12ξ31g_{p}\xi_{11}\xi_{32}=g_{p}\xi_{12}\xi_{31}
    gpξ11ξ33=ξ13ξ31g_{p}\xi_{11}\xi_{33}=\xi_{13}\xi_{31}
    ξ12ξ23=ξ13ξ22\xi_{12}\xi_{23}=\xi_{13}\xi_{22}
    gpξ12ξ33=ξ13ξ32g_{p}\xi_{12}\xi_{33}=\xi_{13}\xi_{32}
    gpξ21ξ32=gp2ξ22ξ31g_{p}\xi_{21}\xi_{32}=g^{2}_{p}\xi_{22}\xi_{31}
    gpξ21ξ33=gp2ξ23ξ31g_{p}\xi_{21}\xi_{33}=g^{2}_{p}\xi_{23}\xi_{31}
    gpξ22ξ33=ξ23ξ32g_{p}\xi_{22}\xi_{33}=\xi_{23}\xi_{32}

    Here we use [ξ11:ξ12:ξ13],[ξ21:ξ22:ξ23],[ξ31:ξ32:ξ33][\xi_{11}:\xi_{12}:\xi_{13}],[\xi_{21}:\xi_{22}:\xi_{23}],[\xi_{31}:\xi_{32}:\xi_{33}] to denote the coordinates in Rp2×Rp2×Rp2{\mathbb{P}}^{2}_{R_{p}}\times{\mathbb{P}}^{2}_{R_{p}}\times{\mathbb{P}}^{2}_{R_{p}}. Note that even though Artin’s original calculation assume the local ring is a DVR, [Art82, Prop 3.6] does work for any regular local rings [Mae97, Thm 2.1]. If gpg_{p} is part of a regular system of parameters of RpR_{p}, then sections 4 of [Art82] tells us VΛpRpV_{\Lambda_{p}}\otimes R_{p} is indeed regular. If pp is a singular point of S1S_{1} or S2S_{2} which doesn’t lie in S1S2S_{1}\cap S_{2}), VΛpRpV_{\Lambda_{p}}\otimes R_{p} is not regular. However, from the above equations, a direct calculations show that on each standard affine chart (e.g. {ξ11=ξ21=ξ31=1}\{\xi_{11}=\xi_{21}=\xi_{31}=1\}), VΛpRpV_{\Lambda_{p}}\otimes R_{p} can be defined by 4 equations. Hence VΛpRpV_{\Lambda_{p}}\otimes R_{p} has an open cover with each a complete intersection in 𝔸Rp6{\mathbb{A}}^{6}_{R_{p}}, furthermore the coordinate ring of each affine chart is again a complete intersection as a {\mathbb{C}}-algebra by counting dimensions. Hence VΛpRpV_{\Lambda_{p}}\otimes R_{p} is Cohen-Macaulay.

    Consider the points (not necessarily closed) qSpec(Rp)q\in\operatorname{Spec}(R_{p}). If gpmqg_{p}\in m_{q}, the fiber over qq is the union of three standard Hirzebruch surfaces 𝔽1\mathbb{F}_{1}, meeting transversally, such that any pair of them meet along a fiber of one and the (1)(-1)-curve of the other ([Art82, Prop. 3.10]). If gpmqg_{p}\notin m_{q}, the fiber over qq is completely determined by [ξ11:ξ12:ξ13][\xi_{11}:\xi_{12}:\xi_{13}], hence is isomorphic to Rp2{\mathbb{P}}^{2}_{R_{p}}. So, in particular, the closed fiber is the union of three 𝔽1\mathbb{F}_{1} as desired and all fibers have same relative dimension. Again by the "Miracle flatness" theorem, VΛpRpV_{\Lambda_{p}}\otimes R_{p} is flat over Spec(Rp)\operatorname{Spec}(R_{p}). As τ\tau is faithfully flat, πp\pi_{p} is flat in this case.

  3. Case 3:

    pS1S2p\in S_{1}\cap S_{2} or pS1{x23x33=0}p\in S_{1}\cap\{x_{2}^{3}-x_{3}^{3}=0\}. We again write Λp=(fp,gp)ω\Lambda_{p}=(f_{p},g_{p})_{\omega}. A same calculation as in [Mae97, Prop. (2.2),Lemma (2.3)] shows that each fiber over Spec(𝒪3,p)\operatorname{Spec}(\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}}},p) has the same relative dimension and the closed fiber is a cone over a twisted cubic as described in Definition 2.4. Furthermore, in [Mae97, Lemma 2.4], Maeda shows the following facts:

    1. (a)

      VΛpV_{\Lambda_{p}} has an open affine cover

      VΛp=U1U2U3V_{\Lambda_{p}}=U_{1}\cup U_{2}\cup U_{3}

      where U1U_{1} and U2U_{2} are hypersurfaces in 𝔸𝒪3,p3{\mathbb{A}}^{3}_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}}.

    2. (b)

      U3U_{3} is a (3,3)(3,3)- complete intersection in 𝔸𝒪3,p4{\mathbb{A}}^{4}_{\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}}.

    Notice that here U1,U2,U3U_{1},U_{2},U_{3} do not have to be regular as fp,gpf_{p},g_{p} are not part of a local parameters in the maximal ideal of the local ring at pp, for some pp. For example, when p=[0:0:1:0]p=[0:0:1:0]. However, we can still conclude that U1,U2,U3U_{1},U_{2},U_{3} are Cohen-Macaulay since they are complete intersection, hence so is VΛpV_{\Lambda_{p}}. So we have πp\pi_{p} is flat by the "Miracle flatness" theorem.

As the base field is {\mathbb{C}}, the argument above shows that the fiber over each geometric point is indeed one of the three cases in Definition 2.4. This shows that VΛV_{\Lambda} is a Brauer-Severi surface bundle over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. We denote it by YY in the following sections of this paper as before. ∎

Now we explain which surfaces in 3{\mathbb{P}}^{3}_{{\mathbb{C}}} admit an associated Brauer-Severi surface bundle:

Definition 5.6.

Let SS be a reduced surface in 3{\mathbb{P}}^{3}_{{\mathbb{C}}} with irreducible components S=S1S2SmS=S_{1}\cup S_{2}\cup\cdots\cup S_{m}. Then we say SS admits a nontrivial triple cover étale in codimension 1 if there is nontrivial element in

i=1nH1((Si),/3)ker(C(C1))\bigoplus_{i=1}^{n}H^{1}({\mathbb{C}}(S_{i}),{\mathbb{Z}}/3)\cap\ker(\bigoplus_{C}(\partial^{1}_{C}))

Where CC runs over all irreducible curves in 3{\mathbb{P}}^{3}, C1\partial^{1}_{C} is the residue map as in Definition 2.1.

It is clear that any surface admits a nontrivial triple cover étale in codimension 1 will give us a 3-torsion Brauer class in (3){\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}}) by Bloch-Ogus sequence as discussed in Example 4.1.

So with the proof of Theorem 5.5, we have:

Corollary 5.7.

Let S3S\subset{\mathbb{P}}^{3}_{{\mathbb{C}}} be a reduced surface which admits a nontrivial triple cover étale in codimension 1 (Definition 5.6). Assume the 3-torsion Brauer class given by the Bloch-Ogus sequence is represented by a cyclic algebra 𝒜\mathscr{A} of degree 3. Then there exists a Brauer-Severi surface bundle YS3Y_{S}\to{\mathbb{P}}^{3}_{{\mathbb{C}}} with discriminant locus SS associated to 𝒜\mathscr{A}. Furthermore, YSY_{S} is reduced. If the discriminant locus SS is good (Definition 2.5, indeed here we only need that part (3)(3) of this definition holds), then YSY_{S} is also irreducible, hence integral.

Proof.

The first part of this corollary directly follows from a similar discussion of local structures as in Theorem 5.5. Next, we show that YSY_{S} is reduced. Indeed, the map YS3Y_{S}\to{\mathbb{P}}^{3}_{{\mathbb{C}}} is projective, hence closed. Then for any point yYSy\in Y_{S}, there is a point yy^{\prime} lying in a closed fiber such that yy specializes to yy^{\prime}. Since any localization of a reduced ring is again reduced, it suffices to check the local ring 𝒪YS,y\mathscr{O}_{Y_{S},y^{\prime}} is reduced. Further more it suffices to assume yy^{\prime} is a closed point. This can be directly checked using the explicit equations given in the proof of Theorem 5.5. (Details are discussed in Lemma A.6.)

Now assume that the discriminant locus SS is good (Definition 2.5). Let π\pi denote the structure morphism YS3Y_{S}\to{\mathbb{P}}^{3}_{{\mathbb{C}}}. Consider the restricted Brauer-Severi surface bundle:

π:π1(3S)3S,\pi^{{}^{\prime}}:\pi^{-1}({\mathbb{P}}^{3}_{{\mathbb{C}}}-S)\to{\mathbb{P}}^{3}_{{\mathbb{C}}}-S,

the base here is clearly irreducible. Since each fiber is a smooth Brauer-Severi surface, of the same dimension, and π\pi^{{}^{\prime}} is projective, hence closed, we conclude that π1(3S)\pi^{-1}({\mathbb{P}}^{3}_{{\mathbb{C}}}-S) is irreducible. Now if YSY_{S} is reducible, then by the argument above, π1(S)\pi^{-1}(S) is reducible. Hence there exist an irreducible component SiS_{i} of SS, such that π1(Si)\pi^{-1}(S_{i}) is reducible (note that YSY_{S} is connected). But this is a contradiction to part (3)(3) of Definition 2.5. ∎

6. The Specialization method and Desingularization

In this section, we apply a specialization method introduced by Voisin in [Voi15]. It was further developed by Colliot-Thélène and Pirutka in [CTP16] and modified by Schreieder in [Sch19a, Proposition 26]. We use this last version below as it is most suitabe to our example. We also refer to [HPT18, Section 2] for a brief introduction of the Specialization method. The main difficulty of applying this to Example 4.1 is that we need to construct an explicit desingularization

f:Y~Yf:\tilde{Y}\to Y

so that for all field extension L/L/{\mathbb{C}}, ff induces an isomorphism :

f:CH0(Y~L)CH0(YL)f_{*}:\operatorname{CH}_{0}(\tilde{Y}_{L})\to\operatorname{CH}_{0}(Y_{L})

Recently, Schreieder gave an alternate approach in a series of papers: [Sch19a, Proposition 26] and [Sch19b]. Instead of constructing such a desingularization, Schreieder’s result allows a purely cohomological criteria. Guided by his idea, we have the following known lemma(e.g. [Sch21, Proposition 4.8(a)]):

Lemma 6.1.

Let YY be a projective variety over a field kk. Let EYE\subset Y be an irreducible subvariety such that the local ring of YY at the generic point ηE\eta_{E} of EE, denoted by 𝒪Y,ηE\mathscr{O}_{Y,\eta_{E}}, is a regular local ring. Then there exists a restriction map:

ResEY:Hnr2(k(Y)/k,/3)H2(k(E),/3)\operatorname{Res}^{Y}_{E}:H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3)\to H^{2}(k(E),{\mathbb{Z}}/3)
Proof.

Let αHnr2(k(Y)/k,/3)\alpha\in H^{2}_{nr}(k(Y)/k,{\mathbb{Z}}/3) be an unramified Brauer class (Def 2.2). Notice that by assumption, 𝒪Y,ηE\mathscr{O}_{Y,\eta_{E}} is a regular local ring with residue field k(E)k(E) and fraction field k(Y)k(Y). We have the following diagram:

0{0}H2(𝒪Y,ηE,/3){H^{2}(\mathscr{O}_{Y,\eta_{E}},{\mathbb{Z}}/3)}H2(k(E),/3){H^{2}(k(E),{\mathbb{Z}}/3)}H2(k(Y),/3){H^{2}(k(Y),{\mathbb{Z}}/3)}H1(k(ν),/3){\bigoplus H^{1}(k(\nu),{\mathbb{Z}}/3)}ν2\scriptstyle{\oplus\partial^{2}_{\nu}}

here the left column is given by [CT95, Theorem 3.8.3]. The horizontal map is given by the functoriality in étale cohomology. Now that α\alpha is an unramified Brauer class, it is killed by ν2\oplus\partial^{2}_{\nu}. Hence α\alpha comes from a class in H2(𝒪Y,ηE,/3)H^{2}(\mathscr{O}_{Y,\eta_{E}},{\mathbb{Z}}/3), which can be further mapped to H2(k(E),/3)H^{2}(k(E),{\mathbb{Z}}/3) by the horizontal map. ∎

We use notation in Example 4.1 and Lemma 6.1. Let UYU\subset Y be the smooth locus of YY. Let α1Br((3))[3]\alpha_{1}\in Br({\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}}))[3] be the Brauer class which has nontrivial residue γ1\gamma_{1} along S1S_{1}, and trivial residues everywhere else. By Lemma 2.1, α1\alpha_{1} can be represents by the cyclic algebra

(x23x33x03,FS1x09)ω.(\frac{x_{2}^{3}-x_{3}^{3}}{x_{0}^{3}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}.

By arguments in Example 4.1, α1\alpha_{1} can be lifted to a nontrivial unramified Brauer class

α~1Hnr2((Y)/,/3)\tilde{\alpha}_{1}\in H^{2}_{nr}({\mathbb{C}}(Y)/{\mathbb{C}},{\mathbb{Z}}/3)

Now we prove that the second hypothesis in [Sch19a, Proposition 26] is true in our case:

Lemma 6.2.

Let π:Y3\pi:Y\to{\mathbb{P}}^{3}_{{\mathbb{C}}} be the Brauer-Severi surface bundle in Example 4.1. Let UU be the smooth locus of YY. Then there exists a resolution of singularities f:Y~Yf:\tilde{Y}\to Y, such that for each irreducible component EE of Y~f1(U)\tilde{Y}-f^{-1}(U), ResEY~(α~1)\operatorname{Res}^{\tilde{Y}}_{E}(\tilde{\alpha}_{1}) is trivial.

Proof.

The existence of resolution of singularities of YY is guaranteed by Hironaka’s theorem [Hir64]. We can further assume without loss of generality that each EE is a prime divisor of Y~\tilde{Y}.

Recall that YY has singular locus of codimension at least 2. So for every irreducible component EE of Y~f1(U)\tilde{Y}-f^{-1}(U), f(E)f(E) has dimension at most 33. On the other hand, since that the generic fiber of π\pi is smooth, and the generic fiber over each irreducible component of the discriminant locus (namely S1S2S_{1}\cup S_{2}) is the union of three standard Hirzebruch surfaces 𝔽1\mathbb{F}_{1} described in Definition 2.4. We know π(f(E))\pi(f(E)) has dimension at most 1 (This follows from that local model of π\pi over S1S_{1} and S2S_{2} is smooth, see the discussion in Theorem 5.5). In other words, each EE in Y~\tilde{Y} would dominate a curve or a point in S1S2S_{1}\cup S_{2}.

In the following of this proof, let K=(3)K={\mathbb{C}}({\mathbb{P}}^{3}_{{\mathbb{C}}}) be the function field of 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. Denote by pEp_{E} the generic point of π(f(E))\pi(f(E)), by KPEK_{P_{E}} the field of fractions of the regular complete local ring 𝒪^3,pE\hat{\mathscr{O}}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p_{E}}. We give a case by case argument according to the generic point pE3p_{E}\in{\mathbb{P}}^{3}_{{\mathbb{C}}} of π(f(E))\pi(f(E)):

  • Case 1:

    pES1p_{E}\notin S_{1}. Then ResEY~(α~1)=0\operatorname{Res}^{\tilde{Y}}_{E}(\tilde{\alpha}_{1})=0 simply follows from the fact that pEp_{E} does not belongs to the discriminant locus of α1\alpha_{1} (see e.g. the proof of [Sch19b, Proposition 5.1(2)]).

  • Case 2:

    pES1S2p_{E}\in S_{1}-S_{2}. Consider the following commutative diagram coming from functoriality:

    H2(K(YK),/3){H^{2}(K(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}})}H2(KpE(YK),/3){H^{2}(K_{p_{E}}(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}})}H2(K,/3){H^{2}(K,{\mathbb{Z}}/3{\mathbb{Z}})}H2(KpE,/3){H^{2}(K_{p_{E}},{\mathbb{Z}}/3{\mathbb{Z}})}

    Because α~1\tilde{\alpha}_{1} is unramified, by [Pir23, Proposition 2.5], it suffices to check α1=0\alpha_{1}=0 in H2(KpE(YK),/3)H^{2}(K_{p_{E}}(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}}). Indeed, as YKY_{K} is the Brauer-Severi surface associate to the cyclic algebra 𝒜=(FS2(x23x33)x021,FS1x09)ω\mathscr{A}=(\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x_{0}^{21}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega} and FS2F_{S_{2}} is a nonzero cube when FS1=0F_{S_{1}}=0 in the residue field of 𝒪3,pE\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p_{E}}. By Cohen’s structure theorem [Coh46, Theorem 15], the residue field embeds into KPEK_{P_{E}}. Hence, after taking base changes to KpEK_{p_{E}},

    (FS2(x23x33)x021,FS1x09)ω(x23x33x03,FS1x09)ω.(\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x_{0}^{21}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}\cong(\frac{x_{2}^{3}-x_{3}^{3}}{x_{0}^{3}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}.

    This implies that YKY_{K} is also the Brauer-Severi surface associate to α1\alpha_{1}, we conclude that α1=0\alpha_{1}=0 in H2(KpE(YK),/3)H^{2}(K_{p_{E}}(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}}) by Amitsur’s theorem [GS06, Theorem 5.4.1].

  • Case 3:

    pES1S2p_{E}\in S_{1}\cap S_{2} is a closed point, and pE{x23x33=0}p_{E}\notin\{x_{2}^{3}-x_{3}^{3}=0\}. Then notice α1\alpha_{1} can be represented by the cyclic algebra

    (x23x33x03,FS1x09)ω((x2x3)6(x23x33)2,FS1(x23x33)3)ω.(\frac{x_{2}^{3}-x_{3}^{3}}{x_{0}^{3}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}\cong(\frac{(x_{2}-x_{3})^{6}}{(x_{2}^{3}-x_{3}^{3})^{2}},\frac{F_{S_{1}}}{(x_{2}^{3}-x_{3}^{3})^{3}})_{\omega}.

    Then (x2x3)6(x23x33)2\frac{(x_{2}-x_{3})^{6}}{(x_{2}^{3}-x_{3}^{3})^{2}} is nonzero in the residue field of 𝒪3,pE\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p_{E}}, which is {\mathbb{C}}. Hence (x2x3)6(x23x33)2\frac{(x_{2}-x_{3})^{6}}{(x_{2}^{3}-x_{3}^{3})^{2}} is a cube in the residue field as {\mathbb{C}} is algebraically closed. By Cohen’s structure theorem, the residue field {\mathbb{C}} embeds into KPEK_{P_{E}}, hence (x2x3)6(x23x33)2\frac{(x_{2}-x_{3})^{6}}{(x_{2}^{3}-x_{3}^{3})^{2}} is also a cube in KPEK_{P_{E}}. This shows that α1\alpha_{1} is trivial in H2(KPE,/3)H^{2}(K_{P_{E}},{\mathbb{Z}}/3{\mathbb{Z}}). By the same commutative diagram as in case 2, it is clear that α1=0\alpha_{1}=0 in H2(KpE(YK),/3)H^{2}(K_{p_{E}}(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}}). We then have ResEY~(α~1)=0\operatorname{Res}^{\tilde{Y}}_{E}(\tilde{\alpha}_{1})=0 by [Pir23, Proposition 2.5].

  • Case 4:

    pEp_{E} is one of [0:0:1:1][0:0:1:1],[0:0:1:ω][0:0:1:\omega],[0:0:1:ω2][0:0:1:\omega^{2}]. In these cases, by the discussions in Case 4 of the proof of Theorem 3.1, we can choose another appropriate representing algebras of α1\alpha_{1}:

    (x26(x13x23)(x33x13),FS1x09)ω.(\frac{x_{2}^{6}}{(x_{1}^{3}-x_{2}^{3})(x_{3}^{3}-x_{1}^{3})},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}.

    It is straight forward to check this is a representing algebra of α1\alpha_{1} by applying Lemma 2.1. And x26(x13x23)(x33x13)\frac{x_{2}^{6}}{(x_{1}^{3}-x_{2}^{3})(x_{3}^{3}-x_{1}^{3})} is a nontrivial unit in the residue field of 𝒪3,pE\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p_{E}}, which is {\mathbb{C}}. Hence the remaining proof can be done exactly same as in Case 3.

  • Case 5:

    pE=[0:1:0:0]p_{E}=[0:1:0:0]. In this case, the proof are the same as in Case 4, by using the following representing algebra of α1\alpha_{1}:

    (x16(x13x23)(x33x13),FS1x09)ω.(\frac{x_{1}^{6}}{(x_{1}^{3}-x_{2}^{3})(x_{3}^{3}-x_{1}^{3})},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}.
  • Case 6:

    pEp_{E} is one of the generic point of D1D_{1}, D2D_{2} and D3D_{3}(Example 4.1). Note that by the defining equations of D1,D2D_{1},D_{2} and D3D_{3}, x23x33x03\frac{x_{2}^{3}-x_{3}^{3}}{x_{0}^{3}} is always a nontrivial cube in the residue field of 𝒪3,pE\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p_{E}}. Again as the residue field embeds into KpEK_{p_{E}}, we get α1=0\alpha_{1}=0 in H2(KpE(YK),/3)H^{2}(K_{p_{E}}(Y_{K}),{\mathbb{Z}}/3{\mathbb{Z}}).

This completes the proof. ∎

7. Main result

With Example 4.1, we prove Theorem 1.1: See 1.1

Proof.

By [Sch19a, Proposition 26] and Lemma 6.2,we know the Brauer-Severi surface bundle constructed in Example 4.1 can be used as a reference variety.

To finish the proof, we need to construct a flat family of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}_{{\mathbb{C}}} with Example 4.1 as one closed fiber with smooth general fiber. Start with the cyclic algebra from Example 4.1:

𝒜=(FS2(x23x33)x021,FS1x09)ω\mathscr{A}=(\frac{F_{S_{2}}(x_{2}^{3}-x_{3}^{3})}{x_{0}^{21}},\frac{F_{S_{1}}}{x_{0}^{9}})_{\omega}

We consider two regular surfaces in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}:

G1={x09x19+x28x3+x38x2=0}G_{1}=\{x_{0}^{9}-x_{1}^{9}+x_{2}^{8}x_{3}+x_{3}^{8}x_{2}=0\}
G2={x021+x121+x221x321=0}G_{2}=\{x_{0}^{21}+x_{1}^{21}+x_{2}^{21}-x_{3}^{21}=0\}

By Lemma A.3, both G1G_{1} and G2G_{2} are regular surfaces in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, and they intersect transversally. Consider the following pencil of cyclic algebras:

𝒜[t0:t1]=(t0FS2(x23x33)+t1(G2FS2(x23x33))x021,t0FS1+t1(G1FS1)x09)ω\mathscr{A}_{[t_{0}:t_{1}]}=(\frac{t_{0}F_{S_{2}}(x_{2}^{3}-x_{3}^{3})+t_{1}(G_{2}-F_{S_{2}}(x_{2}^{3}-x_{3}^{3}))}{x_{0}^{21}},\frac{t_{0}F_{S_{1}}+t_{1}(G_{1}-F_{S_{1}})}{x_{0}^{9}})_{\omega}

We denote

t0FS2(x23x33)+t1(G2FS2(x23x33))t_{0}F_{S_{2}}(x_{2}^{3}-x_{3}^{3})+t_{1}(G_{2}-F_{S_{2}}(x_{2}^{3}-x_{3}^{3}))

and

t0FS1+t1(G1FS1)t_{0}F_{S_{1}}+t_{1}(G_{1}-F_{S_{1}})

by FS2[t0:t1]F_{S_{2}}^{[t_{0}:t_{1}]} and FS1[t0:t1]F_{S_{1}}^{[t_{0}:t_{1}]} respectively. By Lemma A.4 and Lemma A.5, when [t0:t1][1:0][t_{0}:t_{1}]\neq[1:0], both FS2[t0:t1]=0F_{S_{2}}^{[t_{0}:t_{1}]}=0 and FS1[t0:t1]=0F_{S_{1}}^{[t_{0}:t_{1}]}=0 are irreducible surfaces in 3{\mathbb{P}}^{3}. Using Lemma 2.1, the induced triple covers are given by

γ1[t0:t1]=FS2[t0:t1]x021,γ2[t0:t1]=x09FS1[t0:t1]\gamma_{1}^{[t_{0}:t_{1}]}=\frac{F_{S_{2}}^{[t_{0}:t_{1}]}}{x_{0}^{21}},\gamma_{2}^{[t_{0}:t_{1}]}=\frac{x_{0}^{9}}{F_{S_{1}}^{[t_{0}:t_{1}]}}

Similar to the discussion in Example 4.1, note that the residue of γ1[t0:t1]\gamma_{1}^{[t_{0}:t_{1}]} of the valuation centered at the point [0:ξ:1:1][0:\xi:1:1] is 1/31\in{\mathbb{Z}}/3, where ξ\xi satisfies ξ92=0\xi^{9}-2=0. And the residue of γ2[t0:t1]\gamma_{2}^{[t_{0}:t_{1}]} of the valuation centered at the point [0:ψ:1:1][0:\psi:1:-1] is 2/32\in{\mathbb{Z}}/3, where ψ21+2=0\psi^{21}+2=0. Hence both γ1[t0:t1]\gamma_{1}^{[t_{0}:t_{1}]} and γ2[t0:t1]\gamma_{2}^{[t_{0}:t_{1}]} are irreducible. By Corollary 5.7, for any [t0:t1]1[t_{0}:t_{1}]\in{\mathbb{P}}^{1}_{{\mathbb{C}}}, there exists an integral Brauer-Severi surface bundle 𝒴[t0:t1]3\mathscr{Y}_{[t_{0}:t_{1}]}\to{\mathbb{P}}^{3}_{{\mathbb{C}}}.

By viewing 𝒜[t0:t1]\mathscr{A}_{[t_{0}:t_{1}]} as a simple algebra over 1×3{\mathbb{P}}^{1}_{{\mathbb{C}}}\times{\mathbb{P}}^{3}_{{\mathbb{C}}} and applying the construction of Theorem 5.5 again, we have a Brauer-Severi surface bundle over 1×3{\mathbb{P}}^{1}_{{\mathbb{C}}}\times{\mathbb{P}}^{3}_{{\mathbb{C}}} which can be viewed as a 1 dimensional family of of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. Let 𝒴1\mathscr{Y}\to{\mathbb{P}}^{1}_{{\mathbb{C}}} denote this family of Brauer-Severi surface bundles. We claim this is indeed a flat family. By [Har77, Proposition 9.7], It is sufficient to check 𝒴\mathscr{Y} is integral. 𝒴\mathscr{Y} is irreducible because each closed fiber is an irreducible variety and the morphism 𝒴1\mathscr{Y}\to{\mathbb{P}}^{1}_{{\mathbb{C}}} is projective, hence closed. Now let 𝒴^\hat{\mathscr{Y}} be the closed sub-scheme of 𝒴\mathscr{Y} with the same underlying topological space equipped with the reduced scheme structure, we have the following Cartesian diagram:

𝒴^[t0:t1]{\hat{\mathscr{Y}}_{[t_{0}:t_{1}]}}𝒴^{\hat{\mathscr{Y}}}𝒴[t0:t1]{\mathscr{Y}_{[t_{0}:t_{1}]}}𝒴{\mathscr{Y}}{[t0:t1]}{\{[t_{0}:t_{1}]\}}1{{\mathbb{P}}^{1}_{{\mathbb{C}}}}i[t0:t1]\scriptstyle{i_{[t_{0}:t_{1}]}}i\scriptstyle{i}

On one hand, i[t0:t1]i_{[t_{0}:t_{1}]} is a homeomorphism on topological spaces as the pullback of schemes by monomorphism coincide with topological pullback according to the explicit construction of fiber product of ringed spaces. On the other hand, i[t0:t1]i_{[t_{0}:t_{1}]} is a closed immersion as a base change of the closed immersion ii. Since 𝒴[t0:t1]\mathscr{Y}_{[t_{0}:t_{1}]} is reduced as discussed above, we know i[t0:t1]i_{[t_{0}:t_{1}]} is an isomorphism.

Finally, by replacing 𝒴\mathscr{Y} by 𝒴^\hat{\mathscr{Y}} if necessary, we get a 1 dimensional flat family of Brauer-Severi surface bundles over 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, with a special fiber 𝒴[1:0]\mathscr{Y}_{[1:0]} (Example 4.1 and Lemma 6.2) and a regular fiber 𝒴[1:1]\mathscr{Y}_{[1:1]} ([Mae97, Theorem 2.1]), hence we are done.

Appendix A calculations

Lemma A.1.

Let P1=P1(x1,x2,x3)=(x13x23)(x23x33)(x33x13)P_{1}=P_{1}(x_{1},x_{2},x_{3})=(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-x_{3}^{3})(x_{3}^{3}-x_{1}^{3}), let P2=P2(x1,x2,x3)=x13x23x33P_{2}=P_{2}(x_{1},x_{2},x_{3})=x_{1}^{3}x_{2}^{3}x_{3}^{3}. Then for any [t1:t2][0:1][t_{1}:t_{2}]\neq[0:1] or [1:0][1:0] or [1:27][1:\sqrt{-27}] or [1:27][1:-\sqrt{-27}], singular locus of the curve t1P1+t2P2=0t_{1}P_{1}+t_{2}P_{2}=0 in 2=Proj[x1,x2,x3]{\mathbb{P}}^{2}_{{\mathbb{C}}}=\operatorname{Proj}{\mathbb{C}}[x_{1},x_{2},x_{3}] are precisely the three points:

[1:0:0],[0:1:0],[0:0:1][1:0:0],[0:1:0],[0:0:1]
Proof.

Taking partial derivatives:

t1P1x1+t2P2x1=0t_{1}\frac{\partial P_{1}}{\partial x_{1}}+t_{2}\frac{\partial P_{2}}{\partial x_{1}}=0
t1P1x2+t2P2x2=0t_{1}\frac{\partial P_{1}}{\partial x_{2}}+t_{2}\frac{\partial P_{2}}{\partial x_{2}}=0
t1P1x3+t2P2x3=0t_{1}\frac{\partial P_{1}}{\partial x_{3}}+t_{2}\frac{\partial P_{2}}{\partial x_{3}}=0

Consider a point with x1=0x_{1}=0 lying in the curve. Then P2=0P_{2}=0, and this forces P1=x23x33(x33x23)=0P_{1}=x_{2}^{3}x_{3}^{3}(x_{3}^{3}-x_{2}^{3})=0 as t10t_{1}\neq 0. If one of x2x_{2} or x3x_{3} is 0, then we get the singularities listed in the statement of this lemma. If not, we have x23x33=0x_{2}^{3}-x_{3}^{3}=0, and the above partial derivatives can be simplified to obtain x2=0x_{2}=0, and hence x3=0x_{3}=0, which is a contradiction. Similar calculations work when x2=0x_{2}=0 or x3=0x_{3}=0. So all singularities when some xix_{i} equal to 0 are precisely the three points listed above.

In the following we assume all of x1x_{1}, x2x_{2} and x3x_{3} are nonzero. It is clear that for iji\neq j, we have xiP2xi=xjP2xjx_{i}\frac{\partial P_{2}}{\partial x_{i}}=x_{j}\frac{\partial P_{2}}{\partial x_{j}}. The above partial derivatives also tell us

x1P1x1=x2P1x2=x3P1x3.x_{1}\frac{\partial P_{1}}{\partial x_{1}}=x_{2}\frac{\partial P_{1}}{\partial x_{2}}=x_{3}\frac{\partial P_{1}}{\partial x_{3}}.

By Euler’s theorem on homogeneous polynomial, we know

9P1=x1P1x1+x2P1x2+x3P1x3.9P_{1}=x_{1}\frac{\partial P_{1}}{\partial x_{1}}+x_{2}\frac{\partial P_{1}}{\partial x_{2}}+x_{3}\frac{\partial P_{1}}{\partial x_{3}}.

Hence we have

x1P1x1=3P1,x_{1}\frac{\partial P_{1}}{\partial x_{1}}=3P_{1},

which simplifies to

(x23x33)(x23x33x16)=0(x_{2}^{3}-x_{3}^{3})(x_{2}^{3}x_{3}^{3}-x_{1}^{6})=0

Similarly,

(x33x13)(x13x33x26)=0(x_{3}^{3}-x_{1}^{3})(x_{1}^{3}x_{3}^{3}-x_{2}^{6})=0
(x13x23)(x13x23x36)=0(x_{1}^{3}-x_{2}^{3})(x_{1}^{3}x_{2}^{3}-x_{3}^{6})=0

By our assumption here, P20P_{2}\neq 0, so P10P_{1}\neq 0. Hence xi3xj3x_{i}^{3}\neq x_{j}^{3}, so we have

x23x33x16=0x_{2}^{3}x_{3}^{3}-x_{1}^{6}=0
x13x33x26=0x_{1}^{3}x_{3}^{3}-x_{2}^{6}=0
x13x23x36=0x_{1}^{3}x_{2}^{3}-x_{3}^{6}=0

Hence the original partial derivatives simplify to

3t1(x33x23)+t2x13=03t_{1}(x_{3}^{3}-x_{2}^{3})+t_{2}x_{1}^{3}=0
3t1(x23x13)+t2x33=03t_{1}(x_{2}^{3}-x_{1}^{3})+t_{2}x_{3}^{3}=0
3t1(x13x33)+t2x23=03t_{1}(x_{1}^{3}-x_{3}^{3})+t_{2}x_{2}^{3}=0

Viewing this as three linear equations of xi3x_{i}^{3}, we can calculate the determinant of the coefficient matrix as t2(27t12+t22)-t_{2}(27t_{1}^{2}+t_{2}^{2}). Hence when [t1:t2][t_{1}:t_{2}] is not one of the four cases listed in the statement, the determinant is non-zero, we know x13=x23=x03=0x_{1}^{3}=x_{2}^{3}=x_{0}^{3}=0 is the only solution, which is impossible. ∎

Lemma A.2.

Using notations in Example 4.1, singular locus of S2S_{2} consists of three lines:

L1={x0=0,x1=0}L_{1}=\{x_{0}=0,x_{1}=0\}
L2={x0=0,x2=0}L_{2}=\{x_{0}=0,x_{2}=0\}
L3={x0=0,x3=0}L_{3}=\{x_{0}=0,x_{3}=0\}

Each DiD_{i} exactly passes through 5 singular points of S2S_{2}, they are as follows:

D1:[0:0:0:1],[0:0:1:0],[0:0:1:1],[0:0:1:ω],[0:0:1:ω2]D_{1}:[0:0:0:1],[0:0:1:0],[0:0:1:1],[0:0:1:\omega],[0:0:1:\omega^{2}]
D2:[0:0:0:1],[0:1:0:0],[0:1:0:1],[0:1:0:ω],[0:1:0:ω2]D_{2}:[0:0:0:1],[0:1:0:0],[0:1:0:1],[0:1:0:\omega],[0:1:0:\omega^{2}]
D3:[0:0:1:0],[0:1:0:0],[0:1:1:0],[0:1:ω:0],[0:1:ω2:0]D_{3}:[0:0:1:0],[0:1:0:0],[0:1:1:0],[0:1:\omega:0],[0:1:\omega^{2}:0]
Proof.

Let P1=P1(x1,x2,x3)=(x13x23)(x23x33)(x33x13)P_{1}=P_{1}(x_{1},x_{2},x_{3})=(x_{1}^{3}-x_{2}^{3})(x_{2}^{3}-x_{3}^{3})(x_{3}^{3}-x_{1}^{3}), and let P2=P2(x1,x2,x3)=x13x23x33P_{2}=P_{2}(x_{1},x_{2},x_{3})=x_{1}^{3}x_{2}^{3}x_{3}^{3}. Then any singular point of S2S_{2} satisfies the one of the two systems of equations:

{x0=0P2P1x1+(P12P2)P2x1=0P2P1x2+(P12P2)P2x2=0P2P1x3+(P12P2)P2x3=0.\begin{cases}x_{0}=0\\ P_{2}\frac{\partial P_{1}}{\partial x_{1}}+(P_{1}-2P_{2})\frac{\partial P_{2}}{\partial x_{1}}=0\\ P_{2}\frac{\partial P_{1}}{\partial x_{2}}+(P_{1}-2P_{2})\frac{\partial P_{2}}{\partial x_{2}}=0\\ P_{2}\frac{\partial P_{1}}{\partial x_{3}}+(P_{1}-2P_{2})\frac{\partial P_{2}}{\partial x_{3}}=0\\ \end{cases}\,. (A.1)

or

{x09=12(P1P2)(P1+P2)P1x1+(P13P2)P2x1=0(P1+P2)P1x2+(P13P2)P2x2=0(P1+P2)P1x3+(P13P2)P2x3=0.\begin{cases}x_{0}^{9}=-\frac{1}{2}(P_{1}-P_{2})\\ (P_{1}+P_{2})\frac{\partial P_{1}}{\partial x_{1}}+(P_{1}-3P_{2})\frac{\partial P_{2}}{\partial x_{1}}=0\\ (P_{1}+P_{2})\frac{\partial P_{1}}{\partial x_{2}}+(P_{1}-3P_{2})\frac{\partial P_{2}}{\partial x_{2}}=0\\ (P_{1}+P_{2})\frac{\partial P_{1}}{\partial x_{3}}+(P_{1}-3P_{2})\frac{\partial P_{2}}{\partial x_{3}}=0\\ \end{cases}\,. (A.2)

In case ((A.1).1), clearly points in L1L2L3L_{1}\cup L_{2}\cup L_{3} are solutions of this system of equations. Assume xi0,i=1,2,3x_{i}\neq 0,i=1,2,3. Then multiply the second equation of ((A.1).1) by x1x_{1}, multiply the third equation of ((A.1).1) by x2x_{2}, multiply the last equation of ((A.1).1) by x3x_{3} and add the resulting equations. By Euler’s theorem on homogeneous polynomial, we have

P1P2=0P_{1}-P_{2}=0

and the partial derivatives can be simplified as

x0=0x_{0}=0
P1x1P2x1=0\frac{\partial P_{1}}{\partial x_{1}}-\frac{\partial P_{2}}{\partial x_{1}}=0
P1x2P2x2=0\frac{\partial P_{1}}{\partial x_{2}}-\frac{\partial P_{2}}{\partial x_{2}}=0
P1x3P2x3=0\frac{\partial P_{1}}{\partial x_{3}}-\frac{\partial P_{2}}{\partial x_{3}}=0

Hence by Lemma A.1, the set of all singularities in this case is identified to L1L2L3L_{1}\cup L_{2}\cup L_{3}.

In case ((A.2).2), it is easy to check if some xi=0,i=1,2,3x_{i}=0,i=1,2,3, then either we are reduced to case ((A.1).1) or we obtain that all of them are 0. So we again assume xi0,i=1,2,3x_{i}\neq 0,i=1,2,3, and use the same trick as the previous case. By Euler’s theorem on homogeneous polynomial, we have

0=P12+2P1P23P22=(P1P2)(P1+3P2)0=P_{1}^{2}+2P_{1}P_{2}-3P_{2}^{2}=(P_{1}-P_{2})(P_{1}+3P_{2})

If P1P2=0P_{1}-P_{2}=0, then x0=0x_{0}=0, we are reduced to the case ((A.1).1). So the only new possibility is P1+3P2=0P_{1}+3P_{2}=0. Then the partial derivatives are exactly the partial derivatives for the curve P1+3P2=0P_{1}+3P_{2}=0. Again by Lemma A.1, the set of singularities of S2S_{2} are L1L2L3L_{1}\cup L_{2}\cup L_{3}. ∎

Lemma A.3.

Let

G1={x09x19+x28x3+x38x2=0}G_{1}=\{x_{0}^{9}-x_{1}^{9}+x_{2}^{8}x_{3}+x_{3}^{8}x_{2}=0\}
G2={x021+x121+x221x321=0}G_{2}=\{x_{0}^{21}+x_{1}^{21}+x_{2}^{21}-x_{3}^{21}=0\}

Then G1G_{1} and G2G_{2} are regular surface in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. And furthermore G1G_{1} and G2G_{2} intersect transversally.

Proof.

G2G_{2} is clearly a regular surface in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}. Taking partial derivatives of defining equation of G1G_{1}, the singular points are defined by the equations:

9x08=09x_{0}^{8}=0
9x18=0-9x_{1}^{8}=0
8x27x3+x38=08x_{2}^{7}x_{3}+x_{3}^{8}=0
8x37x2+x28=08x_{3}^{7}x_{2}+x_{2}^{8}=0

This gives x0=x1=0x_{0}=x_{1}=0. Note that if one of x2x_{2} or x3x_{3} is 0, so is the other. Assume x20x_{2}\neq 0 and x30x_{3}\neq 0, we get 8x27+x37=0,8x37+x27=08x_{2}^{7}+x_{3}^{7}=0,8x_{3}^{7}+x_{2}^{7}=0. Then again x2=x3=0x_{2}=x_{3}=0, a contradiction. Hence G1G_{1} is also a regular surface in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}.

To check G1G_{1} intersects G2G_{2} transversally, we prove by contradiction: Assume there is a point P=[x0:x1:x2:x3]G1G2P=[x_{0}:x_{1}:x_{2}:x_{3}]\in G_{1}\cap G_{2}, such that there exists a nonzero complex number kk, with

21x020=k(9x08)21x_{0}^{20}=k(9x_{0}^{8})
21x120=k(9x18)-21x_{1}^{20}=k(9x_{1}^{8})
21x220=k(8x27x3+x38)21x_{2}^{20}=k(8x_{2}^{7}x_{3}+x_{3}^{8})
21x320=k(8x37x2+x28)-21x_{3}^{20}=k(8x_{3}^{7}x_{2}+x_{2}^{8})

Where the left side of each equations is the partial derivatives of defining equations of G1G_{1}, and the right hand side is kk times the partial derivatives of defining equations of G2G_{2}. We split into several cases:

  1. Case 1:

    x2=0x_{2}=0 or x3=0x_{3}=0. In this case, we clearly have x2=x3=0x_{2}=x_{3}=0, hence x00x_{0}\neq 0 and x10x_{1}\neq 0. So the partial derivatives with respect x0x_{0} and x1x_{1} tells us:

    x012=921k=x112x_{0}^{12}=\frac{9}{21}k=-x_{1}^{12}

    As PG1G2P\in G_{1}\cap G_{2}, we also have

    x09x19=0x_{0}^{9}-x_{1}^{9}=0
    x021+x121=0x_{0}^{21}+x_{1}^{21}=0

    This forces x0=x1=0x_{0}=x_{1}=0, which makes this case impossible.

  2. Case 2:

    x20x_{2}\neq 0 and x30x_{3}\neq 0. Then the partial derivatives with respect to x2x_{2} and x3x_{3} shows that

    x2208x27x3+x38=k21=x3208x37x2+x28\frac{x_{2}^{20}}{8x_{2}^{7}x_{3}+x_{3}^{8}}=\frac{k}{21}=-\frac{x_{3}^{20}}{8x_{3}^{7}x_{2}+x_{2}^{8}}

    (note that the denominator is nonzero, otherwise by the partial derivatives with respect to x2x_{2} and x3x_{3}, one of x2x_{2} or x3x_{3} is 0. This contradicts to the assumption.) This can be simplified to

    x2218x27+x37=x3218x37+x27\frac{x_{2}^{21}}{8x_{2}^{7}+x_{3}^{7}}=-\frac{x_{3}^{21}}{8x_{3}^{7}+x_{2}^{7}}

    Since x30x_{3}\neq 0, we assume x3=1x_{3}=1 without lose of generality. Let t=x27t=x_{2}^{7}, we see the above relation shows that tt is a root of the following polynomial:

    t4+8t3+8t+1=0t^{4}+8t^{3}+8t+1=0

    Now we have three subcases:

    1. Sub-case 1:

      x0=0x_{0}=0 and x1=0x_{1}=0. Then the defining polynomial of G2G_{2} tells us

      t31=0t^{3}-1=0

      This contradicts to the relation: t4+8t3+8t+1=0t^{4}+8t^{3}+8t+1=0.

    2. Sub-case 2:

      x00x_{0}\neq 0 and x10x_{1}\neq 0. In this case, a similar argument as in Case 1 shows that

      x012+x112=0x_{0}^{12}+x_{1}^{12}=0

      So we may assume x0=τx1x_{0}=\tau x_{1}, where τ\tau satisfies τ12+1=0\tau^{12}+1=0. Plug these information into defining equations of G1G_{1} and G2G_{2}, we get:

      G1:(τ91)x19+x2x3(t+1)=0G_{1}:(\tau^{9}-1)x_{1}^{9}+x_{2}x_{3}(t+1)=0
      G2:(τ91)x121+t31=0G_{2}:-(\tau^{9}-1)x_{1}^{21}+t^{3}-1=0

      Note that τ910\tau^{9}-1\neq 0, and hence the above two equations shows that t+10t+1\neq 0 and t310t^{3}-1\neq 0. Taking ratio of the above two equations, we have:

      x112=t31x2x3(t+1)-x_{1}^{12}=\frac{t^{3}-1}{x_{2}x_{3}(t+1)}

      Compare the above relation with G1G_{1}, we have:

      x13=(τ91)(t31)x22(t+1)2x_{1}^{3}=\frac{(\tau^{9}-1)(t^{3}-1)}{x_{2}^{2}(t+1)^{2}}

      Hence we get:

      t31τ91=x121=(τ91)7(t31)7t2(t+1)14\frac{t^{3}-1}{\tau^{9}-1}=x_{1}^{21}=\frac{(\tau^{9}-1)^{7}(t^{3}-1)^{7}}{t^{2}(t+1)^{14}}

      This is simplified to

      (τ91)8(t31)6t2(t+1)14=0(\tau^{9}-1)^{8}(t^{3}-1)^{6}-t^{2}(t+1)^{14}=0

      That is

      (τ91)8=t2(t+1)14(t31)6(\tau^{9}-1)^{8}=\frac{t^{2}(t+1)^{14}}{(t^{3}-1)^{6}}

      Since tt satisfies t4+8t3+8t+1=0t^{4}+8t^{3}+8t+1=0, we list all roots of this polynomial:

      t1=232152+62t_{1}=-2-\frac{3}{\sqrt{2}}-\sqrt{\frac{15}{2}+6\sqrt{2}}
      t2=232+152+62t_{2}=-2-\frac{3}{\sqrt{2}}+\sqrt{\frac{15}{2}+6\sqrt{2}}
      t3=2+321152+62t_{3}=-2+\frac{3}{\sqrt{2}}-\sqrt{-1}\sqrt{-\frac{15}{2}+6\sqrt{2}}
      t4=2+32+1152+62t_{4}=-2+\frac{3}{\sqrt{2}}+\sqrt{-1}\sqrt{-\frac{15}{2}+6\sqrt{2}}

      By taking norm of both sides of (τ91)8=t2(t+1)14(t31)6(\tau^{9}-1)^{8}=\frac{t^{2}(t+1)^{14}}{(t^{3}-1)^{6}} for each value of tt above, we see all four possible values of tt are impossible. (Indeed, one can check the norm of the left hand side has two possible estimated values: 0.1180.118 or 135.882135.882, while the norm of the right hand side has two possible estimated values: 0.002380.00238 or 14.277814.2778.)

    3. Sub-case 3:

      One of x0x_{0} or x1x_{1} is 0, and the other is nonzero. A similar discussion as in Sub-case 2 gives us

      1=t2(t+1)14(t31)61=\frac{t^{2}(t+1)^{14}}{(t^{3}-1)^{6}}

      Again by taking norms of both sides, we see this is also impossible.

This completes the calculation. ∎

Lemma A.4.

Use notations of Example 4.1 and Lemma A.2, let

G1={x09x19+x28x3+x38x2=0}G_{1}=\{x_{0}^{9}-x_{1}^{9}+x_{2}^{8}x_{3}+x_{3}^{8}x_{2}=0\}

Then for any [t0:t1][1:0]1[t_{0}:t_{1}]\neq[1:0]\in{\mathbb{P}}^{1}_{{\mathbb{C}}},

FS1[t0:t1]=t0FS1+t1(G1FS1)=0F_{S_{1}}^{[t_{0}:t_{1}]}=t_{0}F_{S_{1}}+t_{1}(G_{1}-F_{S_{1}})=0

defines irreducible surfaces in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}.

Proof.

We can view FS1[t0:t1]F_{S_{1}}^{[t_{0}:t_{1}]} as a polynomial in x0x_{0}:

FS1[t0:t1]=t0x09+t1(x19+x28x3+x38x2)+(t0t1)P1=0F_{S_{1}}^{[t_{0}:t_{1}]}=t_{0}x_{0}^{9}+t_{1}(-x_{1}^{9}+x_{2}^{8}x_{3}+x_{3}^{8}x_{2})+(t_{0}-t_{1})P_{1}=0

Hence by the Eisenstein’s criterion, it suffices to show the curve in 2{\mathbb{P}}^{2}_{{\mathbb{C}}} defined by the constant term

t1(x19+x28x3+x38x2)+(t0t1)P1=0t_{1}(-x_{1}^{9}+x_{2}^{8}x_{3}+x_{3}^{8}x_{2})+(t_{0}-t_{1})P_{1}=0

has a regular point. It is clear that [0:1:0][0:1:0] is such a point. ∎

Lemma A.5.

Use notations of Example 4.1 and Lemma A.2, let

G2={x021+x121+x221x321=0}G_{2}=\{x_{0}^{21}+x_{1}^{21}+x_{2}^{21}-x_{3}^{21}=0\}

Then for any [t0:t1][1:0]1[t_{0}:t_{1}]\neq[1:0]\in{\mathbb{P}}^{1}_{{\mathbb{C}}},

FS2[t0:t1]=t0FS2(x23x33)+t1(G2FS2(x23x33))=0F_{S_{2}}^{[t_{0}:t_{1}]}=t_{0}F_{S_{2}}(x_{2}^{3}-x_{3}^{3})+t_{1}(G_{2}-F_{S_{2}}(x_{2}^{3}-x_{3}^{3}))=0

define irreducible surfaces in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}.

Proof.

View FS2[t0:t1]F_{S_{2}}^{[t_{0}:t_{1}]} as a polynomial of x0x_{0}. As any factor of a homogeneous polynomial is also homogeneous, it suffices to show the constant term with respect to x0x_{0} is itself irreducible. That is we need to show that

t1x121(t0t1)(x23x33)P2(P1P2)+t1(x221x321)=0t_{1}x_{1}^{21}-(t_{0}-t_{1})(x_{2}^{3}-x_{3}^{3})P_{2}(P_{1}-P_{2})+t_{1}(x_{2}^{21}-x_{3}^{21})=0

is irreducible. View the above polynomial as a polynomial of x1x_{1}, using Eisenstein criterion with the prime factor (x2x3)(x_{2}-x_{3}), we get the conclusion. ∎

Lemma A.6.

With notations as in the proof of Theorem 5.5 and Corollary 5.7, we have that YSY_{S} is reduced.

Proof.

As stated in the proof of Corollary 5.7, it suffices to check that over any closed point p3p\in{\mathbb{P}}^{3}_{{\mathbb{C}}}, and any point yy lying in the fiber over pp, the local ring 𝒪VΛp,y𝒪YS,y\mathscr{O}_{V_{\Lambda_{p}},y}\cong\mathscr{O}_{Y_{S},y} is reduced.

In the proof of Theorem 5.5, we provide open affine covers for each local model VΛpV_{\Lambda_{p}}. Hence it suffice to show the coordinate ring of each open affine set appearing in these open affine covers is reduced. We discuss the cases as in the proof of Theorem 5.5 separately. First, Case 1 is trivial as the local model is regular.

In Case 2, we consider the the affine chart {ξ11=ξ21=ξ31=1}\{\xi_{11}=\xi_{21}=\xi_{31}=1\}. Then its coordinate ring is

Rp[ξ12,ξ13,ξ22,ξ23,ξ32,ξ33]/(gpξ22ξ12,gpξ23ξ13,gpξ32gpξ12,gpξ33ξ13)R_{p}[\xi_{12},\xi_{13},\xi_{22},\xi_{23},\xi_{32},\xi_{33}]/(g_{p}\xi_{22}-\xi_{12},g_{p}\xi_{23}-\xi_{13},g_{p}\xi_{32}-g_{p}\xi_{12},g_{p}\xi_{33}-\xi_{13})

It is easy to check that this defining ideal is radical. All the other affine charts can be checked similarly.

In Case 3, by [Mae97, Lemma 2.4], the local model has an open affine cover consisting of three open affine charts. The first two affine charts are both hypersurfaces in 𝔸RP3{\mathbb{A}}^{3}_{R_{P}} and since each is defined by an irreducible polynomial, each affine chart is reduced. For the third chart, we need to be careful since it is not a hypersurface. Its coordinate ring is given by

Rp[x,y,z,w]/(F1,F2),R_{p}[x,y,z,w]/(F_{1},F_{2}),
F1=y3ωx2w+(1ω)xyzfp,F_{1}=y^{3}-\omega x^{2}w+(1-\omega)xyz-f_{p},
F2=z3ω2xw2(1ω)xyzgp.F_{2}=z^{3}-\omega^{2}xw^{2}-(1-\omega)xyz-g_{p}.

for some fp.gpRpf_{p}.g_{p}\in R_{p}. Here ω\omega is a primitive third root of unity. Recall that Rp=𝒪3,p=[x¯1,x¯2,x¯3]S0R_{p}=\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}={\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3}]_{S_{0}}, where x¯i\bar{x}_{i} is a coordinate for some standard affine chart in 3{\mathbb{P}}^{3}_{{\mathbb{C}}}, and S0S_{0} is the multiplicative set [x¯1,x¯2,x¯3]mp{\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3}]-m_{p} with mpm_{p} the maximal ideal associates to pp. Hence:

Rp[x,y,z,w]/(F1,F2)([x¯1,x¯2,x¯3,x,y,z,w]/(F1,F2))S0.\displaystyle R_{p}[x,y,z,w]/(F_{1},F_{2})\cong({\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w]/(F_{1},F_{2}))_{S_{0}}.

We check that this algebra is reduced using Serre’s criterion. Namely, we verify whether our ring satisfies (R0)(R_{0}) and (S1)(S_{1}) [Sta24]. Note that F1,F2F_{1},F_{2} do not have common factors in the polynomial ring [x¯1,x¯2,x¯3,x,y,z,w]{\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w]. Hence F1,F2F_{1},F_{2} form a regular sequence, and so We have that [x¯1,x¯2,x¯3,x,y,z,w]/(F1,F2){\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w]/(F_{1},F_{2}) is a complete intersection. In particular, this affine chart is Cohen-Macaulay. This shows [x¯1,x¯2,x¯3,x,y,z,w]/(F1,F2))S0{\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w]/(F_{1},F_{2}))_{S_{0}} is also Cohen-Macaulay and hence satisfies Serre’s condition (S1)(S_{1}). On the other hand, one easily checks that F1F_{1} and F2F_{2} intersect transversally by showing that the rows of the jacobian matrix are never proportional along their intersection whenever both of them are nonzero. Note that this follows immediately since the part of the jacobian matrix corresponding to the variables x,y,z,x,y,z, and ww already satisfies this property. Hence the 2×72\times 7 Jacobian matrix of F1,F2F_{1},F_{2} is not of full rank if and only if at least one of the two rows is zero. By the Jacobian criterion, these are precisely the singular points. We easily see that these points correspond to prime ideals in [x¯1,x¯2,x¯3,x,y,z,w]{\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w] containing one of the following ideals: (x,y,fp)(x,y,f_{p}),(z,x,y,gp)(z,x,y,g_{p}),(z,x,w,gp)(z,x,w,g_{p}) or (z,y,w,gp)(z,y,w,g_{p}). Hence singular set of [x¯1,x¯2,x¯3,x,y,z,w]/(F1,F2){\mathbb{C}}[\bar{x}_{1},\bar{x}_{2},\bar{x}_{3},x,y,z,w]/(F_{1},F_{2}) has codimension at least 3, which remains true by passing to the localization with respect to S0S_{0} as fp,gpf_{p},g_{p} lives in the maximal ideal of 𝒪3,p\mathscr{O}_{{\mathbb{P}}^{3}_{{\mathbb{C}}},p}. Thus Rp[x,y,z,w]/(F1,F2)R_{p}[x,y,z,w]/(F_{1},F_{2}) is regular in codimension 0, namely (R0)(R_{0}). Hence this affine chart is also reduced. This completes the proof. ∎

References

  • [ABvBP20] Asher Auel, Christian Böhning, Hans-Christian Graf von Bothmer, and Alena Pirutka. Conic bundle fourfolds with nontrivial unramified Brauer group. J. Algebraic Geom., 29(2):285–327, 2020.
  • [AM72] M. Artin and D. Mumford. Some Elementary Examples of Unirational Varieties Which are Not Rational. Proceedings of the London Mathematical Society, s3-25(1):75–95, 07 1972.
  • [Art82] M. Artin. Left ideals in maximal orders. In Brauer groups in ring theory and algebraic geometry (Wilrijk, 1981), volume 917 of Lecture Notes in Math., pages 182–193. Springer, Berlin-New York, 1982.
  • [BO74] Spencer Bloch and Arthur Ogus. Gersten’s conjecture and the homology of schemes. Ann. Sci. École Norm. Sup. (4), 7:181–201 (1975), 1974.
  • [Coh46] I. S. Cohen. On the structure and ideal theory of complete local rings. Transactions of the American Mathematical Society, 59(1):54–106, 1946.
  • [CT95] J.-L. Colliot-Thélène. Birational invariants, purity and the Gersten conjecture. In KK-theory and algebraic geometry: connections with quadratic forms and division algebras (Santa Barbara, CA, 1992), volume 58 of Proc. Sympos. Pure Math., pages 1–64. Amer. Math. Soc., Providence, RI, 1995.
  • [CTO89] Jean-Louis Colliot-Thélène and Manuel Ojanguren. Variétés unirationnelles non rationnelles: au-delà de l’exemple d’artin et mumford. Inventiones mathematicae, 97(1):141–158, Feb 1989.
  • [CTP16] Jean-Louis Colliot-Thélène and Alena Pirutka. Hypersurfaces quartiques de dimension 3: non-rationalité stable. Ann. Sci. Éc. Norm. Supér. (4), 49(2):371–397, 2016.
  • [CTS21] Jean-Louis Colliot-Thélène and Alexei N. Skorobogatov. The Brauer-Grothendieck group, volume 71 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer, Cham, [2021] ©2021.
  • [GS06] Philippe Gille and Tamás Szamuely. Central simple algebras and Galois cohomology, volume 101 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2006.
  • [Har77] Robin Hartshorne. Algebraic geometry, volume No. 52 of Graduate Texts in Mathematics. Springer-Verlag, New York-Heidelberg, 1977.
  • [Hir64] Heisuke Hironaka. Resolution of singularities of an algebraic variety over a field of characteristic zero. I, II. Ann. of Math. (2), 79:109–203; 79 (1964), 205–326, 1964.
  • [HPT18] Brendan Hassett, Alena Pirutka, and Yuri Tschinkel. Stable rationality of quadric surface bundles over surfaces. Acta Math., 220(2):341–365, 2018.
  • [IOOV17] Colin Ingalls, Andrew Obus, Ekin Ozman, and Bianca Viray. Unramified Brauer classes on cyclic covers of the projective plane. In Brauer groups and obstruction problems, volume 320 of Progr. Math., pages 115–153. Birkhäuser/Springer, Cham, 2017. With an appendix by Hugh Thomas.
  • [KT19] Andrew Kresch and Yuri Tschinkel. Models of Brauer-Severi surface bundles. Mosc. Math. J., 19(3):549–595, 2019.
  • [KT20] Andrew Kresch and Yuri Tschinkel. Stable rationality of Brauer-Severi surface bundles. Manuscripta Math., 161(1-2):1–14, 2020.
  • [Mae97] Takashi Maeda. On standard projective plane bundles. J. Algebra, 197(1):14–48, 1997.
  • [NSW08] Jürgen Neukirch, Alexander Schmidt, and Kay Wingberg. Cohomology of Number Fields. Grundlehren der mathematischen Wissenschaften. Springer Berlin, Heidelberg, Berlin, Heidelberg, 2 edition, 2008. Published: 18 February 2008 (Hardcover), 26 September 2013 (eBook), 23 August 2016 (Softcover).
  • [Pir18] Alena Pirutka. Varieties that are not stably rational, zero-cycles and unramified cohomology. In Algebraic geometry: Salt Lake City 2015, volume 97.2 of Proc. Sympos. Pure Math., pages 459–483. Amer. Math. Soc., Providence, RI, 2018.
  • [Pir23] Alena Pirutka. Cubic surface bundles and the brauer group, 2023.
  • [Sch19a] Stefan Schreieder. On the rationality problem for quadric bundles. Duke Math. J., 168(2):187–223, 2019.
  • [Sch19b] Stefan Schreieder. Stably irrational hypersurfaces of small slopes. J. Amer. Math. Soc., 32(4):1171–1199, 2019.
  • [Sch21] Stefan Schreieder. Unramified cohomology, algebraic cycles and rationality, 2021.
  • [See99] George F. Seelinger. Brauer-Severi schemes of finitely generated algebras. Israel J. Math., 111:321–337, 1999.
  • [Sta23] The Stacks project authors. The stacks project. https://stacks.math.columbia.edu/tag/00R4, 2023.
  • [Sta24] The Stacks project authors. The stacks project. https://stacks.math.columbia.edu, 2024.
  • [TOP17] ADAM TOPAZ. Abelian-by-central galois groups of fields i: A formal description. Transactions of the American Mathematical Society, 369(4):pp. 2721–2745, 2017.
  • [VdB88] Michel Van den Bergh. The Brauer-Severi scheme of the trace ring of generic matrices. In Perspectives in ring theory (Antwerp, 1987), volume 233 of NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., pages 333–338. Kluwer Acad. Publ., Dordrecht, 1988.
  • [Voi15] Claire Voisin. Unirational threefolds with no universal codimension 22 cycle. Invent. Math., 201(1):207–237, 2015.