Rationality of Brauer-Severi surface bundles over rational 3-folds
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 geometry2020 Mathematics Subject Classification:
Primary 14E08,14M201. Introduction
This paper is motivated by the study of the stable rationality of conic bundles and Brauer-Severi surface bundles. Let be a projective variety over an algebraically closed field . We say that is rational if is birational to a projective space for some natural number . We call stably rational if there exists a natural number such that is rational. In [HPT18], an example of a quadratic surface bundle over that is not stably rational and has a nontrivial unramified Brauer group is constructed. This example is realized as a divisor in 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 via the second projection. This structure allowed the authors to establish a sufficient condition [ABvBP20, Thm. 2.6] for a conic bundle over to not be stably rational. Furthermore, they introduced a new example of a flat family of conic bundles over , 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 -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 . 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 [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 , 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 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 , 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,
Let be the function field of an integral scheme over the field of complex numbers, . The Kummer sequence provides a Galois cohomology sequence that identifies the 3-torsion subgroup of with the second Galois cohomology group of with constant coefficients :
By the Merkurjev-Suslin theorem [GS06, Thm. 2.5.7], is generated by the equivalence classes of cyclic algebras of order 3. These are denoted by and are defined by
where and is a primitive third root of unity.
Now let be a discrete valuation of with residue field , we have the following residue maps [GS06, Section 6.8]:
By Kummer theory, we can identify these residue maps as:
Lemma 2.1.
With notations as above, the two residue maps are defined by:
Proof.
See [IOOV17, Thm. 2.18] ∎
Definition 2.2.
The 3-torsion of unramified Brauer group of a field over another field , denoted by or , is the intersection of kernels of all residue maps , where take values in all divisorial valuations of which are trivial on .
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 be a field with characteristic not 3. Let be a regular, proper, integral variety over with function field , then we have
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 be a locally Noetherian scheme, in which 3 is invertible in the local rings. A Brauer-Severi surface bundle over is a flat projective morphism such that the fiber over every geometric point of is isomorphic to one of the following:
-
•
-
•
The union of three standard Hirzebruch surfaces , meeting transversally, such that any pair of them meets along a fiber of one and the -curve of the other.
-
•
An irreducible scheme whose underlying reduced subscheme is isomorphic to the cone over a twisted cubic curve.
Let be a Brauer-Severi surface bundle over a smooth projective rational threefold over a field whose generic fiber is smooth, and let denote its discriminant locus. We consider with its reduced closed subscheme structure in , and since is Noetherian, consists of finitely many irreducible components, say .
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 is good if the following conditions are satisfied:
-
(1)
Each irreducible component of is reduced. (This is assumed above.)
-
(2)
The fiber over a general for each irreducible component is geometrically the union of three standard Hirzebruch surfaces described in Definition 2.4.
-
(3)
The natural triple cover of induced by is irreducible.
-
(4)
By , the fiber over the generic point of is irreducible. Thus, there is a natural map from the cubic classes of function field of to the cubic classes of function field of . We assume the cubic extension over the function field of induced by generates the kernel of .
Remark 2.6.
Remark 2.7.
By Lemma 2.1, those are precisely those irreducible surfaces such that the Brauer class of the generic fiber have a nontrivial residue along . 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 be a smooth nonsplit Brauer-Severi surface over an arbitrary field (in particular, ). Then the pullback map induces an exact sequence:
where the kernel is generated by the Brauer class determined by . Furthermore, if characteristic of and contains a primitive -th root of unit, then the above exact sequence restricts to :
Proof.
We first prove the exactness of the first sequence. Recall that the kernel is given by Amitsur’s theorem [GS06, Thm. 5.4.1]. To show surjectivity of , consider the separable closure of , and let . Then we have the Hochschild-Serre spectral sequence
The low degree exact sequence reads
By the definition of a Brauer-Severi Variety, we know . Hence we have . Since we clearly have , it follows that is surjective.
For the second part, consider any . There exists a lift of such that . Therefore, . In other words, . Hence, surjects onto . Notice that by the description of the kernel in Amitsur’s theorem, we clearly have
To compute the cokernel, consider the short exact sequence of trivial -modules:
and its associated long exact sequence:
We claim the boundary maps in the above exact sequence are zero. Indeed, let be a primitive third root of unit. By the proof of [TOP17, Lemma 4.1], is given by the cup product with . By our assumption, is a cube in , hence . This shows .
To show , consider the following commutative diagram given in [GS06, Lemma 7.5.10]:
Notations in the above diagram are explained in [GS06, Lemma 7.5.10]. Notice since are trivial -modules, we have the following isomorphism ([TOP17, Lemma 4.1]):
Since the upper horizontal map in the commutative diagram is given by the symbol product with , it follows that this map is 0. By the Merkurjev-Suslin theorem ([GS06, Theorem 8.6.5]), the left vertical map is surjective, hence . Given that cup products "commute" with boundary homomorphisms ([NSW08, Proposition 1.4.3]), we have the commutative diagram involving :
From this, it is clear that .
Next notice that has characteristic , so we have when is a power of 3. Consider the following commutative diagram:
Then the snake lemma gives us
Since is multiplication by 3 and , . The map is also zero as maps onto . Hence we have ∎
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 be an algebraically closed field of characteristic and let be a Brauer-Severi surface bundle over a smooth projective threefold over with a smooth generic fiber. Assume and . (For example, take .) Let be the Brauer class over corresponding to the generic fiber of , and it can be represented by a cyclic algebra of index 3. Assume the discriminant locus is good [Def. 2.5] with irreducible components . Further suppose the following conditions also hold:
-
(1)
Any irreducible curve in is contained in at most two surfaces from the set .
-
(2)
Through any point of , there pass at most three surfaces from the set .
-
(3)
For all , and are factorial at every point of .
Let . Let be the subgroup of given by . We write elements of as with .
Let consist of those elements such that whenever there exists an irreducible component of , such that
Let be the subgroup consisting of elements such that whenever there exists an irreducible components of , such that
Then contains the subquotient .
Proof.
First, note that under these assumptions, is necessarily integral(see 5.7), hence we can talk about its function field . We have the following commutative diagram:
We make some observations related to this diagram:
-
(1)
By definition, denotes all those classes in which are unramified with respect to divisorial valuations corresponding to prime divisors on , Since the singular locus of has codimension , we can also characterize as all those classes in that are unramified with respect to divisorial valuations which have centers on Y which are not contained in [ABvBP20, Cor. 2.2].
-
(2)
denotes those classes in which are killed by residue maps associated to divisorial valuations that are trivial on , hence correspond to prime divisors of dominating the base . We use to denote all prime divisors in that do not dominate the base . Then the upper row is exact by definition.
-
(3)
The second row is obtained from Bloch-Ogus complex [BO74], which is exact under the assumptions
-
(4)
The left vertical row is exact by Lemma 2.8, because we have
where is the Brauer-Severi surface (over ) corresponding to the generic fiber . Hence is smooth and we have the last isomorphism in the above statement.
-
(5)
In the right vertical row, the map is induced by the field extensions , coincides with the induced map
If is not contained in the discriminant locus, the generic fiber of is geometrically integral. Then is algebraically closed in , and thus the induced map above is injective. If is a component of the discriminant locus, then after taking the base change to the cubic extension defined by the residue class , the generic fiber of is a union of three Hirzebruch surfaces , meeting transversally so that any pair of them meet along a fiber of one and a -curve of the other. (This is correct because is indeed the preimage of under , with the third assumption in Definition 2.5). In this case, the low degree long exact sequence from the Hochschild-Serre spectral sequence
implies the kernel of the natural map is generated by . By the last assumption in Definition 2.5, we know .
Then we can prove that lies in the image of . In fact, let and denote by again its the image in . Then is killed by . If is not in the image of , it lifts to a class by Lemma 2.8. We have the following exact sequence which is similar to the second row in above diagram with coefficients :
Hence at least one residue must have order 9 (since is injective both for 3-torsion and 9-torsion cases ). On the other hand,
This is correct because (again we use to denote a separable closure of ) in the long exact sequence associate to
We have
As we can calculate ètale cohomology of spectrum of a field using Galois cohomology, we have the kernel:
While the kernel for those doesn’t belongs to the discriminant locus is clearly zero by the same argument in 3-torsion case.
Now we notice that has no elements of order 9, this means can’t be mapped to 0 in , hence a contradiction.
The above diagram chasing in fact gives us
Next we determine classes in that are in . In particular, we show that the subgroup defined earlier is contained in . We do this by checking whether the classes in are unramified with respect to all divisorial valuations of (and not just those that come from prime divisors on ). Consider a class , viewed as an element in . Denote by the image of in . We aim to show that is unramified on if is in . Using the definition of , it is sufficient to check this for valuations whose centers on has codimension at least 1. In the following, we use to denote the local ring of in .
-
Case 1:
The center of on is not contained in the discriminant locus: In this case, for any surface passing through the center of , we have
Then [ABvBP20, Proposition 2.1] tells us is in the image of . Hence is also unramified with respect to in this case.
-
Case 2:
The center of on is contained in the discriminant locus, but not in the intersection of two or more components: Now the center is contained in for a unique . Recall that the component of is or . If , by an argument same as Case 1, is in the image of . Similarly, if , is in the image of . Finally, if , is in the image of . Notice that
So in all three conditions, we have is unramified with respect to in this case.
-
Case 3:
The center on is a curve that is an irreducible component of : In this case, we again check the possible values of and in . We have the following cases:
-
Case 3(a):
If , then the argument in Case 2 above gives us that at least one of or lies in the image of . So we are done in this situation.
-
Case 3(b):
or : By symmetry, we can assume . Notice that
by the exactness of the second row in the diagram. Then we must have
This means that a rational function representing the class
has a zero or a pole of order divisible by 3 along . Without loss of generality, we may assume that the function associated with is contained in the local ring of in .We call this function . Let be a local parameter for in . Such a local parameter exists as is a Cartier divisor on , which in turn follows since is assumed to be factorial along . Then is a unit, and hence any preimage in is also a unit (See Remark 3.2 below). Call such a preimage , which may be viewed as a rational function in . Assume is a local parameter of in . Consider the symbol algebra . Let be a surface containing . By lemma 2.1, we have
On the other hand, by construction, so we have
Also if is a surface passing through other than and .(In fact, by our assumption, such an is not in the discriminant locus and so this agrees with this conclusion.) Hence [ABvBP20, Proposition 2.1] tells us is in the image of . Hence
It then suffices to show that
By assumption, is trivial, hence so is as the center of is .
-
Case 3(c):
or : By symmetry, we can assume . Now the proof is essentially same as Case 3(b), which shows that
It follows that and so is unramified along .
-
Case 3(d):
: Assume . Then applying Case to the class , we see that this case is also proved.
-
Case 3(a):
-
Case 4:
The center of on is a point , here is as in case 3, and are the only surfaces among the that pass through . As we have seen in the discussion of Case 3, we can reduce to the case when . Hence we again have . In fact, this is true for any curve that contains and is contained in . Choose a function representing the class . Then let be all irreducible curves through that are either a zero or a pole for the function . Pick local equations of in , and consider the following rational function on :
Since is assumed to be factorial, in particular, normal at , the above rational function is a unit locally around . Hence it can be lifted to a unit in . Then we can repeat the rest of the proof as in Case 3(b). (Notice that every element in is a cube since is algebraically closed, so the last step of Case 3(b) is automatically true.)
-
Case 5:
The center of on is a point that lies on exactly three distinct surfaces : We consider the possible values of . If , then one of or is unramified. By symmetry and up to subtraction by or , the only remaining cases are , , and . Notice that , and that the case is equivalent to the case . Hence, we only need to consider the case , 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 is a unit, then any preimage in is also a unit. In fact, we have
As is a unit in , there exist a such that . Hence there exist , such that
Notice that is contained in the maximal ideal of , so is a unit in . Hence any preimage is also a unit in .
We prove an immediate corollary in which we weaken the hypothesis about factoriality when . 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 . We continue with the same hypothesis as in the theorem except the following change: we replace the requirement by the following:
(3’) is factorial at every point of .
Then is nontrivial and hence is not stably rational.
Proof.
In this case, the Brauer class in whose representative is can be lifted to a nontrivial unramified Brauer class in ∎
4. Example
In this section, we will construct a Brauer-Severi surface bundle over that is stably non-rational. We use Corollary 3.3 for this purpose.
Example 4.1.
Consider the following two surfaces in :
In the following, we use to denote the equation defines , separately. We start by checking that both and are irreducible and reduced:
-
•
is irreducible and reduced. This follows directly from the fact that the singular locus of has dimension 0. In fact, only singular at 12 isolated points:
Here is a primitive roots of unity. If is not reduced, then the singular locus would have dimension 2. If is not irreducible, the singular locus would have dimension at least 1 by Bèzout theorem.
-
•
is irreducible and reduced. We may rewrite the equation defining as:
where . We may consider the above polynomial as an element in , 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 in , such that
It is evident that an appropriate prime ideal exists if has an irreducible factor with multiplicity 1 and is coprime to . In fact, any irreducible factor of is inherently coprime to . Therefore it suffices to provide a single regular point of to show the existence of such an irreducible factor. Finally, we directly check that is a regular point of . Hence is irreducible.
Now as we have already shown is irreducible, it is sufficient to find a smooth point in to show it is reduced. Indeed, one can easily check that is indeed a smooth point of .
We choose rational triple covers of and defined by:
We claim the triple covers are not trivial: In fact, by Lemma 2.1, the residue of of a valuation centered at the point is . Hence is not trivial. To show is not trivial is equivalent to show is not a cubic in the function field of . And it’s true because the residue of of a valuation centered at the point is . Hence are not trivial.
Consider the corresponding Bloch-Ogus exact sequence:
We have . In fact, it is easy to check that for any curve such that (or ) has a zero or pole along , the order is divided by 3. Hence
can be lifted to a Brauer class . This can be directly checked by Lemma 2.1.
By Theorem 5.5, the cyclic algebra gives out a Brauer-Severi surface bundle . 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)
We already proved that and are reduced.
-
(2)
The behavior of a general fiber over and is given by [Mae97, Thm. 2.1].
-
(3)
The induced triple cover over and are irreducible because , are not trivial.
-
(4)
To show the last requirement in Definition 2.5 is true, we have the following commutative diagram:
Where is defined right after the large diagram in Theorem 3.1. For , is induced by the cubic extension defined by , is induced by the cubic extension defined by , which is equal to the cubic class defined by ([Art82, Thm. 2.1]. ) Note that is injective, and . On the other hand, an easy diagram chasing as in part (5) in proof of Theorem 3.1 shows that contains . This forces to be injective and . Same argument works for .
On the other hand, we list all irreducible components of :
Where
One can easily check they are indeed irreducible using Eisenstein’s criterion by viewing those polynomials as elements in . Notice that passes through only two singular points of : and . It is straightforward to check is indeed a Cartier divisor of , even along these two singular points:
Lemma 4.2.
are Cartier divisors of .
Proof.
By symmetry, it is sufficient to check the behavior of at singular points of . Notice that only passes through two singular points of : and . Let , then we have the local ring:
By expanding the equation defining , we have
Notice is a unit in , hence the ideal defining , which is , is generated by one element . Similarly one can do the calculation for the point . As a result, is a Cartier divisor of .
∎
5. Flatness
In this section, we check the cyclic algebra
indeed gives us a Brauer-Severi surface bundle over 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 be a sheaf of algebra that is torsion free and coherent as an module. We say is an -order in if contains and
Definition 5.2.
For each point , let denote the regular local ring of at . We say a finitely generated algebra is an -order in , if is torsion free and
Remark 5.3.
In this paper, we always assume an order is locally free.
Recall that of [Mae97] describes an -order which we again denote by in the following, we denote its localization at a point by .
Definition 5.4.
Let (respectively, ) be the functor from the category of -schemes (respectively, -schemes) to the category of sets:
where denotes the dual sheaf, denotes the unit group , is the reduced norm and denotes the functor of Grassmannian of -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 , ) and again denote them by , .
Theorem 5.5.
is a Brauer-Severi surface bundle over .
Proof.
According to Definition 5.4, for every closed point in , we have the following commutative diagram of schemes:
We first show is a flat morphism. In order to do so, it suffices to show is flat for all closed points . Indeed, if this is done, the flat locus of would be an open subset of containing all closed points, hence is equal to . Furthermore, by the "Miracle flatness" theorem [Sta23] and the fact that is regular, it suffices to show each is Cohen-Macaulay and each fiber of has the same dimension. We do this by a case-by-case argument for all closed points in :
-
Case 1:
. It is well know that is an Azumaya algebra outside of discriminant locus [Art82]. All fibers of are smooth Brauer-Severi surfaces and furthermore is regular, hence Cohen-Macaulay. By the "Miracle flatness" theorem, is flat in this case.
-
Case 2:
and and . Following ideas from Artin [Art82], we may write as the symbol algebra . That is, over is generated by subject to the relations
Since , it follows that is a unit in . Let , then
is an étale neighborhood of with faithfully flat as it surjects on the underlying topological space. By faithfully flat descent, it suffices to show is flat over . In [Art82], Artin noticed can be viewed as a subalgebra of the 3 by 3 matrices algebra over by setting
And can be embedded into by the following 9 equations with a cyclic permutations in indices:
Here we use to denote the coordinates in . 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 is part of a regular system of parameters of , then sections 4 of [Art82] tells us is indeed regular. If is a singular point of or which doesn’t lie in ), is not regular. However, from the above equations, a direct calculations show that on each standard affine chart (e.g. ), can be defined by 4 equations. Hence has an open cover with each a complete intersection in , furthermore the coordinate ring of each affine chart is again a complete intersection as a -algebra by counting dimensions. Hence is Cohen-Macaulay.
Consider the points (not necessarily closed) . If , the fiber over is the union of three standard Hirzebruch surfaces , meeting transversally, such that any pair of them meet along a fiber of one and the -curve of the other ([Art82, Prop. 3.10]). If , the fiber over is completely determined by , hence is isomorphic to . So, in particular, the closed fiber is the union of three as desired and all fibers have same relative dimension. Again by the "Miracle flatness" theorem, is flat over . As is faithfully flat, is flat in this case.
-
Case 3:
or . We again write . A same calculation as in [Mae97, Prop. (2.2),Lemma (2.3)] shows that each fiber over 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:
-
(a)
has an open affine cover
where and are hypersurfaces in .
-
(b)
is a complete intersection in .
Notice that here do not have to be regular as are not part of a local parameters in the maximal ideal of the local ring at , for some . For example, when . However, we can still conclude that are Cohen-Macaulay since they are complete intersection, hence so is . So we have is flat by the "Miracle flatness" theorem.
-
(a)
As the base field is , 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 is a Brauer-Severi surface bundle over . We denote it by in the following sections of this paper as before. ∎
Now we explain which surfaces in admit an associated Brauer-Severi surface bundle:
Definition 5.6.
Let be a reduced surface in with irreducible components . Then we say admits a nontrivial triple cover étale in codimension 1 if there is nontrivial element in
Where runs over all irreducible curves in , 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 by Bloch-Ogus sequence as discussed in Example 4.1.
So with the proof of Theorem 5.5, we have:
Corollary 5.7.
Let 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 of degree 3. Then there exists a Brauer-Severi surface bundle with discriminant locus associated to . Furthermore, is reduced. If the discriminant locus is good (Definition 2.5, indeed here we only need that part of this definition holds), then 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 is reduced. Indeed, the map is projective, hence closed. Then for any point , there is a point lying in a closed fiber such that specializes to . Since any localization of a reduced ring is again reduced, it suffices to check the local ring is reduced. Further more it suffices to assume 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 is good (Definition 2.5). Let denote the structure morphism . Consider the restricted Brauer-Severi surface bundle:
the base here is clearly irreducible. Since each fiber is a smooth Brauer-Severi surface, of the same dimension, and is projective, hence closed, we conclude that is irreducible. Now if is reducible, then by the argument above, is reducible. Hence there exist an irreducible component of , such that is reducible (note that is connected). But this is a contradiction to part 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
so that for all field extension , induces an isomorphism :
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 be a projective variety over a field . Let be an irreducible subvariety such that the local ring of at the generic point of , denoted by , is a regular local ring. Then there exists a restriction map:
Proof.
Let be an unramified Brauer class (Def 2.2). Notice that by assumption, is a regular local ring with residue field and fraction field . We have the following diagram:
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 is an unramified Brauer class, it is killed by . Hence comes from a class in , which can be further mapped to by the horizontal map. ∎
We use notation in Example 4.1 and Lemma 6.1. Let be the smooth locus of . Let be the Brauer class which has nontrivial residue along , and trivial residues everywhere else. By Lemma 2.1, can be represents by the cyclic algebra
By arguments in Example 4.1, can be lifted to a nontrivial unramified Brauer class
Now we prove that the second hypothesis in [Sch19a, Proposition 26] is true in our case:
Lemma 6.2.
Let be the Brauer-Severi surface bundle in Example 4.1. Let be the smooth locus of . Then there exists a resolution of singularities , such that for each irreducible component of , is trivial.
Proof.
The existence of resolution of singularities of is guaranteed by Hironaka’s theorem [Hir64]. We can further assume without loss of generality that each is a prime divisor of .
Recall that has singular locus of codimension at least 2. So for every irreducible component of , has dimension at most . On the other hand, since that the generic fiber of is smooth, and the generic fiber over each irreducible component of the discriminant locus (namely ) is the union of three standard Hirzebruch surfaces described in Definition 2.4. We know has dimension at most 1 (This follows from that local model of over and is smooth, see the discussion in Theorem 5.5). In other words, each in would dominate a curve or a point in .
In the following of this proof, let be the function field of . Denote by the generic point of , by the field of fractions of the regular complete local ring . We give a case by case argument according to the generic point of :
-
Case 1:
. Then simply follows from the fact that does not belongs to the discriminant locus of (see e.g. the proof of [Sch19b, Proposition 5.1(2)]).
-
Case 2:
. Consider the following commutative diagram coming from functoriality:
Because is unramified, by [Pir23, Proposition 2.5], it suffices to check in . Indeed, as is the Brauer-Severi surface associate to the cyclic algebra and is a nonzero cube when in the residue field of . By Cohen’s structure theorem [Coh46, Theorem 15], the residue field embeds into . Hence, after taking base changes to ,
This implies that is also the Brauer-Severi surface associate to , we conclude that in by Amitsur’s theorem [GS06, Theorem 5.4.1].
-
Case 3:
is a closed point, and . Then notice can be represented by the cyclic algebra
Then is nonzero in the residue field of , which is . Hence is a cube in the residue field as is algebraically closed. By Cohen’s structure theorem, the residue field embeds into , hence is also a cube in . This shows that is trivial in . By the same commutative diagram as in case 2, it is clear that in . We then have by [Pir23, Proposition 2.5].
-
Case 4:
is one of ,,. In these cases, by the discussions in Case 4 of the proof of Theorem 3.1, we can choose another appropriate representing algebras of :
It is straight forward to check this is a representing algebra of by applying Lemma 2.1. And is a nontrivial unit in the residue field of , which is . Hence the remaining proof can be done exactly same as in Case 3.
-
Case 5:
. In this case, the proof are the same as in Case 4, by using the following representing algebra of :
-
Case 6:
is one of the generic point of , and (Example 4.1). Note that by the defining equations of and , is always a nontrivial cube in the residue field of . Again as the residue field embeds into , we get in .
This completes the proof. ∎
7. Main result
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 with Example 4.1 as one closed fiber with smooth general fiber. Start with the cyclic algebra from Example 4.1:
We consider two regular surfaces in :
By Lemma A.3, both and are regular surfaces in , and they intersect transversally. Consider the following pencil of cyclic algebras:
We denote
and
by and respectively. By Lemma A.4 and Lemma A.5, when , both and are irreducible surfaces in . Using Lemma 2.1, the induced triple covers are given by
Similar to the discussion in Example 4.1, note that the residue of of the valuation centered at the point is , where satisfies . And the residue of of the valuation centered at the point is , where . Hence both and are irreducible. By Corollary 5.7, for any , there exists an integral Brauer-Severi surface bundle .
By viewing as a simple algebra over and applying the construction of Theorem 5.5 again, we have a Brauer-Severi surface bundle over which can be viewed as a 1 dimensional family of of Brauer-Severi surface bundles over . Let 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 is integral. is irreducible because each closed fiber is an irreducible variety and the morphism is projective, hence closed. Now let be the closed sub-scheme of with the same underlying topological space equipped with the reduced scheme structure, we have the following Cartesian diagram:
On one hand, 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, is a closed immersion as a base change of the closed immersion . Since is reduced as discussed above, we know is an isomorphism.
Finally, by replacing by if necessary, we get a 1 dimensional flat family of Brauer-Severi surface bundles over , with a special fiber (Example 4.1 and Lemma 6.2) and a regular fiber ([Mae97, Theorem 2.1]), hence we are done.
∎
Appendix A calculations
Lemma A.1.
Let , let . Then for any or or or , singular locus of the curve in are precisely the three points:
Proof.
Taking partial derivatives:
Consider a point with lying in the curve. Then , and this forces as . If one of or is 0, then we get the singularities listed in the statement of this lemma. If not, we have , and the above partial derivatives can be simplified to obtain , and hence , which is a contradiction. Similar calculations work when or . So all singularities when some equal to 0 are precisely the three points listed above.
In the following we assume all of , and are nonzero. It is clear that for , we have . The above partial derivatives also tell us
By Euler’s theorem on homogeneous polynomial, we know
Hence we have
which simplifies to
Similarly,
By our assumption here, , so . Hence , so we have
Hence the original partial derivatives simplify to
Viewing this as three linear equations of , we can calculate the determinant of the coefficient matrix as . Hence when is not one of the four cases listed in the statement, the determinant is non-zero, we know is the only solution, which is impossible. ∎
Lemma A.2.
Using notations in Example 4.1, singular locus of consists of three lines:
Each exactly passes through 5 singular points of , they are as follows:
Proof.
Let , and let . Then any singular point of satisfies the one of the two systems of equations:
(A.1) |
or
(A.2) |
In case A, clearly points in are solutions of this system of equations. Assume . Then multiply the second equation of A by , multiply the third equation of A by , multiply the last equation of A by and add the resulting equations. By Euler’s theorem on homogeneous polynomial, we have
and the partial derivatives can be simplified as
Hence by Lemma A.1, the set of all singularities in this case is identified to .
In case A, it is easy to check if some , then either we are reduced to case A or we obtain that all of them are 0. So we again assume , and use the same trick as the previous case. By Euler’s theorem on homogeneous polynomial, we have
If , then , we are reduced to the case A. So the only new possibility is . Then the partial derivatives are exactly the partial derivatives for the curve . Again by Lemma A.1, the set of singularities of are . ∎
Lemma A.3.
Let
Then and are regular surface in . And furthermore and intersect transversally.
Proof.
is clearly a regular surface in . Taking partial derivatives of defining equation of , the singular points are defined by the equations:
This gives . Note that if one of or is 0, so is the other. Assume and , we get . Then again , a contradiction. Hence is also a regular surface in .
To check intersects transversally, we prove by contradiction: Assume there is a point , such that there exists a nonzero complex number , with
Where the left side of each equations is the partial derivatives of defining equations of , and the right hand side is times the partial derivatives of defining equations of . We split into several cases:
-
Case 1:
or . In this case, we clearly have , hence and . So the partial derivatives with respect and tells us:
As , we also have
This forces , which makes this case impossible.
-
Case 2:
and . Then the partial derivatives with respect to and shows that
(note that the denominator is nonzero, otherwise by the partial derivatives with respect to and , one of or is 0. This contradicts to the assumption.) This can be simplified to
Since , we assume without lose of generality. Let , we see the above relation shows that is a root of the following polynomial:
Now we have three subcases:
-
Sub-case 1:
and . Then the defining polynomial of tells us
This contradicts to the relation: .
-
Sub-case 2:
and . In this case, a similar argument as in Case 1 shows that
So we may assume , where satisfies . Plug these information into defining equations of and , we get:
Note that , and hence the above two equations shows that and . Taking ratio of the above two equations, we have:
Compare the above relation with , we have:
Hence we get:
This is simplified to
That is
Since satisfies , we list all roots of this polynomial:
By taking norm of both sides of for each value of above, we see all four possible values of are impossible. (Indeed, one can check the norm of the left hand side has two possible estimated values: or , while the norm of the right hand side has two possible estimated values: or .)
-
Sub-case 3:
One of or is , and the other is nonzero. A similar discussion as in Sub-case 2 gives us
Again by taking norms of both sides, we see this is also impossible.
-
Sub-case 1:
This completes the calculation. ∎
Lemma A.4.
Proof.
We can view as a polynomial in :
Hence by the Eisenstein’s criterion, it suffices to show the curve in defined by the constant term
has a regular point. It is clear that is such a point. ∎
Lemma A.5.
Proof.
View as a polynomial of . As any factor of a homogeneous polynomial is also homogeneous, it suffices to show the constant term with respect to is itself irreducible. That is we need to show that
is irreducible. View the above polynomial as a polynomial of , using Eisenstein criterion with the prime factor , we get the conclusion. ∎
Lemma A.6.
Proof.
As stated in the proof of Corollary 5.7, it suffices to check that over any closed point , and any point lying in the fiber over , the local ring is reduced.
In the proof of Theorem 5.5, we provide open affine covers for each local model . 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 . Then its coordinate ring is
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 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
for some . Here is a primitive third root of unity. Recall that , where is a coordinate for some standard affine chart in , and is the multiplicative set with the maximal ideal associates to . Hence:
We check that this algebra is reduced using Serre’s criterion. Namely, we verify whether our ring satisfies and [Sta24]. Note that do not have common factors in the polynomial ring . Hence form a regular sequence, and so We have that is a complete intersection. In particular, this affine chart is Cohen-Macaulay. This shows is also Cohen-Macaulay and hence satisfies Serre’s condition . On the other hand, one easily checks that and 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 and already satisfies this property. Hence the Jacobian matrix of 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 containing one of the following ideals: ,, or . Hence singular set of has codimension at least 3, which remains true by passing to the localization with respect to as lives in the maximal ideal of . Thus is regular in codimension 0, namely . 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 -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 cycle. Invent. Math., 201(1):207–237, 2015.