Generalizations of quasielliptic curves
-
-
scAbstract. We generalize the notion of quasielliptic curves, which have infinitesimal symmetries and exist only in characteristic two and three, to a hierarchy of regular curves having infinitesimal symmetries, defined in all characteristics and having higher genera. This relies on the study of certain infinitesimal group schemes acting on the affine line and certain compactifications. The group schemes are defined in terms of invertible additive polynomials over rings with nilpotent elements, and the compactification is constructed with the theory of numerical semigroups. The existence of regular twisted forms relies on Brion’s recent theory of equivariant normalization. Furthermore, extending results of Serre from the realm of group cohomology, we describe non-abelian cohomology for semidirect products, to compute in special cases the collection of all twisted forms.
scKeywords. Group schemes, regular curves, quasielliptic curves, numerical semigroups, twisted forms, non-abelian cohomology
sc2020 Mathematics Subject Classification. 14G17, 14L15, 14L30, 14H45, 20M25, 20J06, 14D06
-
-
cSeptember 11, 2023Received by the Editors on April 12, 2023.
Accepted on September 29, 2023.
Mathematisches Institut, Heinrich-Heine-Universität, 40204 Düsseldorf, Germany
sce-mail: [email protected]
Mathematisches Institut, Heinrich-Heine-Universität, 40204 Düsseldorf, Germany
sce-mail: [email protected]
The research was conducted in the framework of the research training group GRK 2240: Algebro-Geometric Methods in Algebra, Arithmetic and Topology.
© by the author(s) This work is licensed under http://creativecommons.org/licenses/by-sa/4.0/
1. Introduction
Let be a ground field of characteristic . The goal of this paper is to generalize, in an equivariant way, the rational cuspidal curve
(1.1) |
from the cases and , when the automorphism group scheme is non-reduced, to a hierarchy of integral curves whose automorphism group schemes are likewise non-reduced. Here the index indicates the characteristic, and gives the “size” of non-reducedness.
Our motivation originates from the Enriques classification of algebraic surfaces over ground fields : This vast body of theorems on the structure of surfaces was extended by Bombieri and Mumford to positive characteristics (see [BM77, BM76]). Their main insight was the introduction and analysis of quasielliptic fibrations, which are morphisms whose generic fiber is a so-called quasielliptic curve, in other words, a twisted form of (1.1) over the function field , with all local rings regular. One knows that such twisted forms exist only over imperfect fields of characteristic , and Queen gave explicit equations for them (see [Que71, Que72]), although of rather extrinsic nature.
We discovered the hierarchy somewhat accidentally, while seeking a deeper and more intrinsic understanding of quasielliptic curves. The curves do not reveal themselves in any direct way; one has to understand them through their automorphism group scheme . Its crucial part consists of certain infinitesimal group schemes of order acting in a canonical way on the affine line . The underlying scheme is , a singleton formed with iterated Frobenius kernels of the additive group, but endowed with a non-commutative group law. Its definition relies on the so-called additive polynomials , or equivalently the elements of the skew polynomial ring , formed over rings with nilpotent elements.
According to Brion’s recent theory of equivariantly normal curves, see [Bri22a], there is a unique compactification to which the action of extends in an optimal way. In general, it is very difficult to unravel the structure of such compactifications, but here we were able to “guess” an explicit description in terms of the numerical semigroups
monoids that each comprise all but finitely many natural numbers. The guesswork was assisted by computer algebra computations with Magma and GAP, performed in a handful of special cases. Our first main result is that the ensuing toric compactification has an intrinsic meaning.
Theorem (See Theorems 5.4 and 9.4).
The -action on the affine line extends to the compactification
and this projective curve is equivariantly normal with respect to the -action.
For , this is precisely the rational cuspidal curve. The second main result unravels the numerical invariants and infinitesimal symmetries of this hierarchy of projective curves.
In this iterated semidirect product, the additive group is normalized by the infinitesimal group scheme , and both are normalized by the multiplicative group . Note that for , this precisely gives back the computation of Bombieri and Mumford [BM76, Proposition 6], and the above should be seen as a natural generalization.
The computation of the genus relies on general results of Delorme [Del76] on numerical semigroups, applied to our p,n. The determination of the automorphism group is based on further surprising properties of the projective curves : The tangent sheaf turns out to be invertible, actually very ample, giving a canonical inclusion , where is the Lie algebra of the automorphism group scheme . From the canonical linearization , we get a matrix representation for , which is crucial to gain control over its structure. Furthermore, is globally a complete intersection, defined inside by the following homogeneous equations of degree :
Also note that the curves are related by a hierarchy of blowing-ups , where the center is the singular point (cf. Lemma 8.3).
Again building on Brion’s theory of equivariantly normal curves, see [Bri22a], we show that our have, over ground fields with “enough” imperfection, twisted forms where all local rings are regular (cf. Theorem 9.4). These have the same structural properties of , except that the singularities get “twisted away.” In turn, the passage from the rational cuspidal curve to quasielliptic curves is generalized to our hierarchy .
The above relies on rather general observations, which form the third main result of this paper.
Theorem (See Section 9).
Let be a geometrically integral curve with the action of a finite group scheme . Suppose that is étale. Then is equivariantly normal if and only if for some field extension , the base-change admits a twisted form that is regular.
Quasielliptic fibrations play a crucial role in the arithmetic of algebraic surfaces of special type, in particular for K3 surfaces and Enriques surfaces (for an example, see [Sch23b]). We expect that twists over function fields of our play a similar role for surfaces of general type.
By the general theory of non-abelian cohomology and twisted forms, one may view the collection of isomorphism classes of twisted forms over as non-abelian cohomology . For our curves , we determined the automorphism group scheme. This raises the question of how to compute non-abelian cohomology for semidirect products in general. We establish effective techniques to do so and are able apply them at least in the case . Our fourth main result is as follows.
Theorem (See Theorem 11.7).
For , the non-abelian cohomology is
where the union runs over , with and .
Perhaps this is the first explicit determination of via a purely non-abelian cohomological computation of for some relevant hierarchy of schemes . A crucial step in this is the determination of particular twisted forms in the form given by Russell [Rus70], a technique likely to be of independent interest.
Let us quote Bombieri and Mumford [BM76, p. 198]: “The study of special low characteristics can be one of two types: amusing or tedious. It all depends on whether the peculiarities encountered are felt to be meaningful variations of the general picture […] or are felt instead to be accidental and random, due for instance to numerological interactions […].” We think that our results amply show that what Bombieri and Mumford have uncovered for is indeed far from accidental, and belong to a structural hierarchy that indeed can be understood from general principles.
The paper is organized as follows: In Section 2 we develop the theory of additive polynomials over rings that contain nilpotents, study the resulting groups of units, and introduce . The ensuing actions on polynomial rings are discussed in Section 3. Building on these preparations, we give in Section 4 a scheme-theoretic reinterpretation and determine the Lie algebra and the upper and lower central series for the infinitesimal group scheme . In Section 5 we examine the equivariant compactifications of the affine line and introduce our numerical semigroup p,n and the ensuing curve , which turns out to be equivariantly normal. We determine the numerical invariants and deduce several crucial geometric consequences in Section 6. In Section 7 we show that our curve can also be seen as a global complete intersection . In Section 8 its automorphism group scheme is determined. Section 9 contains general results on the relation between equivariant normality and the existence of twisted forms that are regular, which is then applied to our curves . Section 10 is devoted to twisting and the computation of non-abelian cohomology for semidirect products. We apply this in Section 11 to describe the collection of all twisted forms of in the cases and .
Acknowledgments
We heartily thank the two referees for their thorough reading and many valuable suggestions, which helped to improve the paper.
2. Invertible additive polynomials
In this section we gather purely algebraic facts that go into the definition of our infinitesimal group scheme in Section 4. Fix some ring of characteristic , and let be an indeterminate. Recall that polynomials of the form
are called additive polynomials. Another widespread designation is -polynomials. Clearly, the set of all such polynomials is stable under addition and substitution . These two composition laws enjoy the distributive property. In fact, the additive polynomials form an associative ring with respect to these laws, with zero element and unit element . Let us call it the ring of additive polynomials.
It can also be seen as the skew polynomial ring , where designates the Frobenius map . Elements are polynomials in the formal symbol , and multiplication is subject to the relations . In other words, we have
(2.1) |
a modification of the usual Cauchy multiplication. The identification of the skew polynomial ring with the ring of additive polynomials is given by and , so that corresponds to . For psychological reasons, we strongly prefer to make computations in the skew polynomial ring. In the next section, when it comes to actions on the affine line, we shall turn back to the ring of additive polynomials.
Over ground fields, the ring of additive polynomials was introduced and studied by Ore [Ore33]. A discussion from the perspective of skew polynomial rings was given by Jacobson [Jac43, Chapter 3]. More recent presentations appear in [Gos96, Chapter 1] and [GW04, Chapter 2]. For our purposes, however, it will be crucial to allow nilpotent elements. The following two propositions reveal that nilpotent and invertible elements in are characterized as in usual polynomial rings.
Proposition 2.1.
An element of the skew polynomial ring is nilpotent if and only if .
Proof.
Suppose that all coefficients are nilpotent, say . For each , we have
for certain exponents whose precise values are irrelevant in the following reasoning: If , each tuple must contain the -fold repetition of at least one value . Then the product vanishes, and so does the above -fold power.
Conversely, suppose that some is not nilpotent. Choose a prime not containing , and set . Then the image of in the skew polynomial ring is a non-zero nilpotent element. On the other hand, is a domain (this follows from [Jac43, Chapter 3, Section 1, bottom paragraph on p. 29]), giving a contradiction. ∎
This has the following important consequence.
Proposition 2.2.
An element of the skew polynomial ring is invertible if and only if and .
Proof.
The condition is sufficient: Set . By Proposition 2.1, the element is nilpotent, say . Then is a unit, with inverse . Thus is also a unit.
Conversely, suppose . From the group law (2.1), one immediately infers that . Seeking a contradiction, we assume that some , is not nilpotent. Choose such maximal. As above, we find some residue field in which is non-zero. Let be the degree of the image of . From (2.1) one sees that the image of has non-zero term in degree , giving a contradiction. ∎
Given a unit of the form , the inverse can be computed as follows: The condition means and for , which give the recursion formula
(2.2) |
The skew polynomial ring comes with an infinite-dimensional matrix representation , already determined by the assignments
More explicitly, this homomorphism is given by
(2.3) |
Obviously, the map is injective and takes values in the row-finite upper triangular matrices. Note that for each , the top left submatrix indexed by yields a subrepresentation .
We now examine the unit group in more detail, for the time being as an abstract group. It comes with matrix representations , . Note that this factors over the group of invertible upper triangular matrices . Write for the kernels. Clearly , whereas
These form a descending chain of normal subgroups, in other words, a normal series. Clearly, their intersection contains only the unit element.
Proposition 2.3.
The normal series on has quotients
Moreover, we have the commutator formula for all .
Proof.
The multiplicative character given by comes with a canonical splitting , so we get a semidirect product . The ensuing conjugacy action of is given by
For our applications, it will be important to consider certain smaller subgroups inside the unit group, and the following is crucial throughout.
Proposition 2.4.
For each integer , the set
is a subgroup inside the unit group , which is normalized by .
Proof.
Clearly the set contains the unit element. Suppose that and belong to , and write the product as , with coefficients . For , each summand vanishes: If , we already have , and if , we get and thus . For , we have , which vanishes because . Thus .
Next consider the inverse element . The recursion formula (2.2) gives and for . For , each summand vanishes because . For , we have , where the second factor vanishes. Thus .
Finally, for each , we have , which clearly belongs to . So the latter is normalized by . ∎
3. Actions on polynomial rings
We keep the set-up as in the previous section. Obviously, the multiplicative monoid of additive polynomials acts on the polynomial ring via substitution of the indeterminate, in other words by , and one easily checks that this is an action from the right. In turn, we have a group action from the right, given by
(3.1) |
Note that the action of the multiplicative group via is a special case of this. Furthermore, we have the translation action of the additive group , defined by
(3.2) |
Obviously, these actions are faithful, and we arrive at inclusions of and into the opposite automorphism group of .
Proposition 3.1.
Inside the opposite automorphism group of , the group is normalized by , and the intersection is trivial.
Proof.
Suppose that we have elements
For the first assertion, it suffices to check that . This indeed holds because one computes with . It remains to verify the assertion on the intersection. Suppose as automorphisms of ; in other words, . Comparing coefficients at the constant terms gives ; hence the intersection is trivial. ∎
In turn, we get an inclusion of into the opposite automorphism group of . Later, we seek to extend part of this action to certain subrings of in a compatible way. The following observation will be useful: Let be the multiplicative system of all monic polynomials. The resulting localization is denoted by . Since monic polynomials are regular elements from the polynomial ring, the localization map is injective, and we get an inclusion .
Proposition 3.2.
The action from the right of the group on the polynomial ring uniquely extends to .
Proof.
Uniqueness immediately follows from the universal property of localizations. To see existence, consider the larger multiplicative system comprising the polynomials of the form with monic, nilpotent, and . Obviously, this system is stable with respect to the actions (3.1) and (3.2), and we thus get an induced action on . On the other hand, the inclusion gives a canonical map . It remains to verify that every as above becomes invertible in . Indeed, in the factorization , the second factor also is a unit because is invertible and is nilpotent. ∎
4. Scheme-theoretic reinterpretation
Fix a ground field of characteristic . In this section we take a more geometric point of view and reinterpret and extend the results of the preceding sections in terms of schemes and group schemes. We now regard
as a group-valued functor on the category of -algebras . Clearly, the natural transformation
(4.1) |
is an isomorphism of set-valued functors, with group laws ignored. In turn, is a finite group scheme with coordinate ring and order . It contains but one point and is thus an infinitesimal group scheme.
One immediately sees that the restriction of (4.1) to respects the group laws and gives an inclusion of group schemes . Furthermore, for every , we have canonical inclusions of group schemes.
Recall that each scheme over our ground field comes with a relative Frobenius map , given in functorial terms by . Here denotes the abelian group , viewed as an -algebra via the absolute Frobenius map , and . Note that as an -algebra. Hence and thus , provided that arises as base-change from the prime field . For our group scheme , the relative Frobenius map takes the form
Proposition 4.1.
The image of is the subgroup scheme , and its kernel is given by . In particular, we have an identification of restricted Lie algebras .
Proof.
Obviously, the Frobenius map factors over the subgroup scheme . The resulting map is indeed an epimorphism because any -valued point of arises from the -valued point of , for the fppf extension .
An -valued point belongs to the kernel of the Frobenius map if and only if , and hence . The last assertion follows because , and for any group scheme, the inclusion of the Frobenius kernel induces a bijection on Lie algebras. ∎
In turn, the relative Frobenius map yields an extension
(4.2) |
By induction on , we infer that the finite group scheme admits a composition series with quotients isomorphic to . In particular, is unipotent. Since all Lie brackets are trivial, the adjoint representation of the Lie algebra is trivial, and the adjoint representation of the group scheme factors over the quotient . It is not difficult to determine the latter representation: Since the group is trivial, we have
where denotes an indeterminate subject to . The elements , , form a basis of this -vector space. With , where , and using the relations and , we get
Consequently, sends the basis vector to the linear combination . Summing up, in the restricted Lie algebra all brackets and -powers are zero, and the adjoint representation of the group scheme is given by .
As described in Section 3, the groups act from the right on the polynomial ring via -linear maps. This is obviously functorial in and thus constitutes an action of the group scheme on the affine line . Note that this is indeed an action from the left. On -valued points, it is given by
where we set for convenience. Of course, we also have the canonical actions of the multiplicative group and the additive group , given via and , respectively. The following generalizes a key observation of Bombieri and Mumford [BM76, Proposition 6].
Proposition 4.2.
The above actions of the three group schemes on the affine line are faithful. Inside the sheaf , the group scheme is normalized by , and both and are normalized by . Moreover, the intersections
inside the sheaf are trivial.
We thus have an iterated semidirect product, for simplicity written as
(4.3) |
acting faithfully on the affine line . In turn, we get an inclusion of restricted Lie algebras
The elements on the left-hand side can be seen as tuples and correspond to the -derivation of the polynomial ring . For example, the derivation acts via , which coincides with the action of the group element .
The spectrum of the function field comes with a monomorphism
(4.4) |
According to Proposition 3.2, there is a unique action on that makes the above morphism equivariant.
Let us close this section with some observations on central series. Recall that for a group , the lower central series and the upper central series are inductively defined by
The group is nilpotent if for some , or equivalently for some . Then the smallest such integers coincide, and this number is called the nilpotency class of the group. Note that , but usually this inclusion is not an equality. We refer to [Hal59, Chapter 10] or [KM79, Chapter 6] for basic facts on nilpotent groups.
For group schemes of finite type, one has basically the same construction, with sheafification involved. This is straightforward for the higher centers: An belongs to if and only if it commutes with all members of up to elements of , for all flat extensions . The situation is more complicated for the r because their formation involves schematic images and group scheme closure with respect to the commutator maps ; see [SGA3-1, Exposé , Section 8].
Let us unravel this for our : consider the closed subschemes
(4.5) |
defined by .
Proposition 4.3.
The are subgroup schemes, and the series (4.5) coincides with both the upper and the lower central series for the group scheme . The quotients are .
Proof.
With descending induction one easily checks that are subgroup schemes: the surjection given by respects the group law and has kernel . The isomorphism theorem gives the statement on the quotients.
The arguments for the higher centers rely on the following observation: The recursion formula (2.2) for inverses shows that each coefficient actually depends only on . From this one easily infers
(4.6) |
Write for the higher centers of . We show by induction on . The case is trivial. Now suppose and that the inclusion holds for . For each from , we compute
Suppose that belongs to . Then for all in some ring extension , the above two expressions coincide modulo . From the equivalence (4.6), we obtain for . For , we are in position to compare coefficients and infer , and thus .
This completes our induction and establishes for all . For the reverse inclusion, we use our embedding into the group of unitriangular matrices. According to Lemma 4.4 below, the higher center of is given by the matrices that are zero on the secondary diagonals above the main diagonal. The intersection with equals . Consequently, ; thus the form the upper central series.
The arguments for the higher commutator groups rely on some preliminary observations. For elements of the form with any , the geometric series gives
where the congruence means up to terms of order . Note that the last summand is only relevant in the special case . With , the above formula shows that the commutator is congruent to
Most summands cancel, and the upshot is the commutator formula
(4.7) |
Write r for the higher commutator subgroup schemes. According to general properties of nilpotent groups (cf. [KM79, p. 107]), we have . We claim that the canonical projection
(4.8) |
is an epimorphism. We check this by induction on . The case is trivial. Suppose and that the assertion is true for . According to [DG70, Section IV.2, Proposition 1.1], the iterated Frobenius kernels are the only subgroup schemes of . Seeking a contradiction, we assume that the above map factors over . Consider the ring . By our induction hypothesis, there are some faithfully flat extension and some -valued point of the form from r-1. With , the commutator formula (4.7) shows that projects to under (4.8). So vanishes in the ring . On the other hand, both of the appearing monomials belong to the monomial basis for , giving a contradiction. Thus (4.8) is an epimorphism.
We are now ready to prove that the inclusion is an equality. Fix some , and write it as . We check that by descending induction on . The case is trivial. Now assume and that the assertion holds for . By the preceding paragraph, there are some faithfully flat extension and some -valued point from s with . Then belongs to . Using our induction hypothesis, together with the inclusion , we see that , and hence , belongs to , and by descent . ∎
Let us point out that the ring is not free as a module over , which one sees by analyzing the size of the Jordan blocks for multiplication by . Thus it is not always possible to factor a given element of into commutators, even over flat extensions .
In the preceding proof, we have used the following fact.
Lemma 4.4.
The unitriangular matrix group , over any ring , has upper central series given by
(4.9) |
Proof.
Over fields, this appears in [KM79, Example 16.1.2]. The general case is formulated in [Rob93] as Exercise 5.1.13. For the sake of completeness, we sketch an argument, by induction on . The case is trivial. Now suppose and that the assertion is true for . By definition, a unitriangular belongs to if and only if
(4.10) |
for every unitriangular matrix . By the induction hypothesis, each has for , and one easily checks that right multiplication with elements of to a unitriangular matrix leaves the -entries unchanged for . From this the inclusion of (4.9) easily follows. Conversely, suppose that we have some , so (4.10) holds. This means
where is any unitriangular matrix and the sums actually run over . From this one easily infers by induction on that vanishes for . ∎
5. Compactifications and numerical semigroups
We keep the setting of the previous section but now work with a new indeterminate . The iterated semidirect product has as coordinate ring
endowed with a Hopf algebra structure, and acts on the affine line . We now seek to extend this action to certain compactifications, all of which are denormalizations of the projective line . For this, we have to make extensive computations in the first chart, which are much easier to carry out with rather than . Note that by Proposition 3.2 we have an induced action on the spectrum of the function field , and this action takes the form
(5.1) |
Recall that an additive submonoid whose complement is finite is called a numerical semigroup. Equivalently, the induced inclusions of groups is an equality, or for some members . Each numerical semigroup comes with the following invariants: The multiplicity is the smallest non-zero element in . The conductor is the smallest integer with . The genus is the cardinality of the complement , whose members are called gaps. As monoid, is finitely generated, and among all systems of generators, there is a smallest one; its cardinality is called the embedding dimension . For general overviews, we refer to the textbooks [RGS09] and [AGS16].
For each numerical semigroup , the ring defines a compactification
of the affine line , obtained by adding a single rational point . The gluing of the two affine open sets is given by the common localization of the coordinate rings. The normalization is , and the ensuing map is described by the conductor square
(5.2) |
which is both cartesian and cocartesian (for details see [FS20, Appendix A]). The conductor loci and are the closed subschemes whose respective coordinate rings are and . Consider the short exact sequence of sheaves on , where the inclusion is the diagonal map and the surjection is the difference map. It yields
(5.3) |
with the invariants of the numerical semigroup discussed above. Here and denote the multiplicity and the embedding dimension of the local ring, respectively.
Given a subgroup scheme , it is natural to ask whether the resulting -action on the affine line extends to the compactification . If it exists, such an extension is unique because the open set is schematically dense in for any ring .
In the following assertion on the constituents of the iterated semidirect product, we regard the expression as a Laurent polynomial in the indeterminate with coefficients from , and as a formal power series in with coefficients from . In both cases we use the ensuing notion of supports and inside the group of exponents .
Proposition 5.1.
We keep the notation as above. Then the following hold:
-
(i)
The multiplicative group always admits an extension.
-
(ii)
For the infinitesimal group scheme , the extension exists if and only if for each and for the Laurent polynomial , we also have .
-
(iii)
For the additive group , the extension exists if and only if for each and for the formal power series , we have .
Moreover, it suffices to verify these conditions for a set of generators .
Proof.
(i) Recall that -actions on affine schemes correspond to -gradings, according to [SGA3-1, Exposé I, Corollary 4.7.3.1]. The action on is given by . This also defines compatible gradings on , which yields the desired extension of the action of .
(ii) The group scheme is infinitesimal; hence every open set on a -scheme is -stable. It follows that the -action extends if and only if the map induced from (5.1) factors over the subring . This map sends to . Note that the second factor is invertible because its second summand is nilpotent. The monomial with is mapped to . This belongs to the subring if and only if for each , the resulting integer belongs to the numerical semigroup .
(iii) The action of on extends to the projective line via the assignments and , with . Note that the origin is fixed but does not admit a stable affine open neighborhood. However, the infinitesimal neighborhoods and in particular the conductor locus are stable.
Since is smooth, any -action on uniquely extends to , according to [Bri22b, Theorem 2]. By [Lau19, Lemma 3.5] the -action on the projective line descends to an action on if and only if the action on the conductor locus descends to an action on . The latter simply means that the map
(5.4) |
factors over . Here the map on the right describes the -action on , and the coordinate ring on the left is , where is the conductor of the numerical semigroup. As -vector space, this is generated by the residue classes of , . The map (5.1) sends to , so the monomial is mapped to . The class of the latter belongs to if and only if for all , we have . ∎
Note that in the expansions of and , some multinomial coefficients appear, and the above conditions involve their congruence properties modulo the prime number . Also note that one may view as a non-normal torus embedding with respect to the one-dimensional torus .
The passage from the constituents to the semidirect product is immediate, thanks to the following observation.
Lemma 5.2.
Suppose that for each constituent of the iterated semidirect product , the action on extends to . Then the whole -action extends to .
Proof.
This is a general fact: all relations between the -valued points of the constituents stemming from the semidirect product structures hold on and thus also on because the former is schematically dense in the latter. ∎
We now introduce a particular that is generated by numbers.
Definition 5.3.
We write for the numerical semigroup generated by
(5.5) |
This is indeed a numerical semigroup because . Its multiplicity is given by
because in the first case, the number is smallest among the generators. Note that is equivalent to , whereas means .
We came up with the above generators by determining for a handful of special cases the largest numerical semigroup for which the group scheme action extends and then guessing the general pattern. The computations were made with the computer algebra systems Magma [BCP97] and GAP [GAP]. One of the main insights of this paper is that the resulting compactifications
lead to the desired generalizations of the quasielliptic curves. Indeed, in the special cases , we get , and the ensuing coordinate rings become . We now verify that the action of the iterated semidirect product extends to this compactification.
Theorem 5.4.
The action of the group scheme on the affine line extends to the compactification .
Proof.
It suffices to extend the action for the three constituents of the iterated semidirect product, by Lemma 5.2, and for this we use Proposition 5.1: The case is immediate. Now suppose , and fix one of the generators listed in (5.5). We have to understand the expression
In the case , the above simplifies to , by the multinomial theorem and . Thus , and obviously . In the case with , we get
Its support equals the set , in light of the defining relations . Obviously, and belong to p,n. Thus the action of extends.
It remains to treat the case . Again we fix one of the generators and now have to examine the formal power series with coefficients from the polynomial ring . For , this becomes
The support is contained in . Clearly, and belong to p,n. The argument for is likewise, and even simpler. Thus the action of extends. ∎
Set and . With respect to the infinitesimal group scheme , all open sets in are stable, and the action on the affine open set is given by the ring homomorphism
with exponents . The orbit map corresponding to the rational point is given by the homomorphism that is implicitly described by
Note that one has to determine the product before substituting because the second factor usually contains terms of negative degree. The computation for the generators (5.5) of our numerical semigroup is immediate: and for . Now recall that the inertia group scheme in is defined by the largest quotient of in which becomes the zero map. Setting , we get the following.
Proposition 5.5.
Inside , the inertia group scheme with respect to the rational point is defined by the equations for .
This inertia group scheme coincides with the canonical inclusion of , which is also the image of the relative Frobenius map, and we thus obtain a -stable closed subscheme . A priori, this is an effective Weil divisor supported by , of degree . The following observation will be crucial in what follows.
Proposition 5.6.
The Weil divisor is an effective Cartier divisor.
Proof.
The closed subscheme lies in the affine open set and corresponds to the ideal . This ideal contains the monomial , and we claim that the inclusion is an equality. In other words, we have to verify that the resulting map
is injective. We computed above that its image is the subring generated by the powers for , which is a -algebra of degree . So it suffices to verify that the -algebra has degree at most . This algebra is generated by the classes of with . From the relation
in the numerical semigroup , we infer a factorization in the ring , and hence . Thus has degree at most . ∎
6. The complete intersection property
We keep the notation as in the preceding section and continue to study the algebra of the numerical semigroup and also the geometry of the compactification of the affine line defined by the coordinate ring .
Recall that any numerical semigroup given by a set of generators and ensuing surjection is called a complete intersection if the congruence is generated by elements. According to [Her70, Corollary 1.13], this is equivalent to the condition that the complete local ring is a complete intersection in the sense of commutative algebra; in other words, for some and some regular sequence , here necessarily with .
Proposition 6.1.
Our numerical semigroup p,n is a complete intersection, and its conductor and genus are given by the formulas
Moreover, is the smallest generating set provided ; for the prime and , one has to omit .
Proof.
First note that for and , the relation shows that the generator does not belong to the smallest generating set.
We now proceed, for general , by induction on . For , we have , and all assertions are obvious. Now suppose and that the assertion holds for . Consider the sets of numbers
Both generate respective numerical semigroups 1 and 2, and the induction hypothesis applies to the former. The numbers and are relatively prime, with and . Furthermore, . According to [Del76, Proposition 10], the monoid is a complete intersection, and the conductor is given by the formula
(6.1) |
Here is the conductor of 2, and is the conductor of 1, which we know by our induction hypothesis. Inserting the latter into (6.1), we get the desired formula for . Every complete intersection semigroup is symmetric (cf. [RGS09, Corollary 9.12]), which simply means that the conductor is twice the genus, and the formula for follows.
Now suppose . By induction, the are the smallest generating sets. The number does not belong to . As explained in [Del76, proof of Proposition 10(ii)], the subset is the smallest generating set. For , one argues likewise, with omitted. ∎
We see that the embedding dimension for the numerical semigroup p,n and the local ring is given by the formula
Let us also record the following geometric consequences.
Corollary 6.2.
The curve has invariants
Moreover, the dualizing sheaf is invertible, of degree .
Proof.
We actually can derive an explicit description for the ring in terms of generators and relations. Write and for the generators of . They give rise to a surjection of monoids and an ensuing congruence . The generators satisfy the obvious relations
(6.2) |
which may be interpreted as members of the congruence . To translate this into commutative algebra, let be indeterminates corresponding to the generators , and consider the surjection
given by and . The map respects the gradings specified by , , and .
Proposition 6.3.
The ideal is generated by the polynomials and for , corresponding to the obvious relations (6.2).
Proof.
This is an application of an observation of Delorme [Del76, Lemma 8]. Recall that our numerical semigroup is generated by the elements . Delorme’s observation hinges on two descending sequences
The first sequence comprises partitions of the generating set , subject to the following condition: is the partition into singletons, and each is obtained from its precursor by replacing certain members by their union. The second sequence consists of homogeneous polynomials in the indeterminates , taking the form for some monic monomials and , each involving only indeterminates indexed by and , respectively. In loc. cit. the sequences are denoted by and , and the pair is called a suite distinguée.
Note that the partitions are fully determined by the sets with . We now define such a partition sequence by setting
Note that this starts with the singletons and . The homogeneous polynomials are declared as
These have for all . One sees
The least common multiple of the above two gcds is given by , which coincides with . Our assertion now follows from [Del76, Lemma 8]. ∎
This has important consequences for Kähler differentials.
Corollary 6.4.
The sheaf is invertible of degree , and the tangent sheaf is invertible of degree .
Proof.
The main task is to compute the module of Kähler differentials for the integral domain . In light of Proposition 6.3, is generated by the differentials and , modulo the relations
(6.3) |
The ring elements and are non-zero because they correspond to monomials in , so and for are torsion. We infer that the map given by the remaining differential is bijective modulo torsion. The latter differential is given by .
Let be the quotient of by its torsion subsheaf, and consider the affine open covering with and . We have trivializations and , given by and . On the overlap these become and , which are related by the cocycle . This gives . The assertion for the dual sheaf is immediate. ∎
7. The projective model
We keep the set-up of the previous section and now describe a projective model for our curve . First note that the obvious relations (6.2) for our monoid can be replaced by
(7.1) |
by using the first of these relations. Now write , and consider the closed subscheme defined by the homogeneous equations
(7.2) |
First observe that is covered by because it contains only the point on the hyperplane given by . On these two charts, we see that
constitute an isomorphism , which we regard as an identification.
Proposition 7.1.
The homogeneous polynomials (7.2) form a regular sequence in the polynomial ring, the curve has degree , and
In particular, X/k is very ample, and with exponent .
Proof.
Let be the ideal generated by the homogeneous polynomials (7.2) inside the -dimensional Cohen–Macaulay ring . Since the scheme is one-dimensional, we must have . It follows from [Sta18, Tag 02JN] that the polynomials in question form a regular sequence. The assertion on the dualizing sheaf immediately follows from and the adjunction formula.
The intersection of with the hyperplane given by is a singleton, with generators and relations for in the homogeneous coordinate ring. Thus .
It remains to verify the statement on the tangent sheaf. As described in the last paragraph of the proof for Corollary 6.4, the invertible sheaf X/K is given by the cocycle with respect to the open covering and . The latter correspond to the open sets and . On the union of these open sets, the invertible sheaf is defined by the cocycle . This becomes after restricting to , and thus . ∎
Recall that a square root for the dualizing sheaf is called a theta characteristic or spin structure (cf. [Ati71] and [Mum71]). In our situation, the curve comes with what one might call an -fold theta characteristic or spin structure.
Another highly relevant consequence: the very ample sheaf has an intrinsic meaning, and becomes the Lie algebra for the automorphism group scheme . To exploit this, we check that the closed embedding is defined by the complete linear system.
Proposition 7.2.
The restriction map is bijective. In particular, .
Proof.
Since the defining polynomials (7.2) have degree , the homogeneous ideal for contains no linear terms. It follows that the map in question is injective. It remains to compute for .
Let us proceed with some general considerations on invertible sheaves on of arbitrary degree . Recall that the conductor loci for the normalization map are given by
where is the conductor for the numerical semigroup . From the cocartesian diagram (5.2), we now obtain an exact sequence
It is not difficult to determine the map in the middle: Making the identification and and , we get
If , the former groups are contained in the latter, and becomes their intersection, and hence for the set
We have to determine this set for , under the assumption . According to Proposition 6.1, the conductor is , and thus . So our set comprises all with . It clearly contains the generators and also the zero element . It remains to check that for each pair of generators , we have . This is obvious for , so assume and with . Then . ∎
This leads to a matrix interpretation of the full automorphism group scheme : First note that the diagonal action of on , and its effects on graphs, induces the conjugacy action of on itself. Its restriction to the first infinitesimal neighborhood of the diagonal X yields the -linearization of the tangent sheaf X/k, and we infer that the resulting representation on the Lie algebra coincides with the adjoint representation . Its projectivization is injective because X/k is very ample, and it follows that is injective as well. We thus have a canonical inclusion that intersects the center trivially. Write for the resulting subgroup scheme and for the vector subspace generated by the homogeneous polynomials (7.2).
Proposition 7.3.
We keep the notation as above. Then equals the stabilizer group scheme for the vector subspace .
Proof.
We start with some observations on the homogeneous coordinate rings
Both rings are integral, so the kernel of the canonical map is a prime ideal, which equals the radical for the ideal generated by the polynomials (7.2). The ideal becomes prime when localized with respect to any homogeneous because is integral. Since our generators form a regular sequence, this actually holds everywhere, and thus .
Write for the stabilizer group scheme in question. It contains because the polynomials are homogeneous, and its action on stabilizes the curve . Modulo , the induced action on is faithful, according to Proposition 7.2. Thus . Conversely, let be some -valued automorphism of . It induces an action on the homogeneous coordinate rings , and likewise for . These actions are compatible; thus stabilizes . ∎
For , this means that is the stabilizer group scheme for the line generated by in the symmetric power of . In turn, is the inertia group scheme for the rational point corresponding to this line.
8. The automorphism group scheme
Recall that our curves come with an inclusion
(8.1) |
of group schemes. The following is one of the main results of this paper.
Theorem 8.1.
The above inclusion of group schemes is an equality provided .
The cases indeed have to be excluded because then . Also note that for , our curve becomes the rational cuspidal curve, and the assertion was established by Bombieri and Mumford [BM76, Proposition 6]. The proof for the above theorem requires some preparation and will be given stepwise. We start with a simple observation.
Lemma 8.2.
For , the ideal for the closed embedding (8.1) is nilpotent.
Proof.
For this, we may assume that is algebraically closed. Seeking a contradiction, we suppose that there is an automorphism that does not yield a rational point in . The assumption ensures , and fixes this singular point. Hence the induced automorphism on the normalization fixes the point defined by . It thus belongs to the inertia group scheme inside . According to [Bri17, Proposition 2.5.1], the action of any smooth group scheme on lifts to an action on the normalization. Thus belongs to inside , giving a contradiction. ∎
Proof of Theorem 8.1 in the special case .
In this situation we have . According to Proposition 7.2, the tangent sheaf has . One easily computes that the rational vector fields
(8.2) |
are everywhere defined, hence form a basis of . Moreover, the first two basis vectors generate a restricted subalgebra , with trivial bracket and -map, and the last basis vector yields a copy of , giving a semidirect product . Such algebras play a prominent role in [Sch07, KS21, ST23]. The bracket and -map are given by and ; cf. [KS21, Proposition 1.1]. One sees that is the image of the bracket, thus the derived subalgebra. This also holds for -valued points; hence is a subrepresentation for .
As explained in Section 7, our curve may also be regarded as the curve in defined by the homogeneous polynomial , with an identification via , , and . According to Proposition 7.2, the monomials yield a basis for . By Proposition 7.1, the invertible sheaves and X are isomorphic. Computing the order of zeros for the homogeneous polynomials and the vector fields (8.2) on , one sees that each identification sends the former basis to the latter basis, at least up to a diagonal base-change matrix.
Combining this with Proposition 7.3, we have a functorial interpretation of the -valued points of as the group of matrices
subject to the sole condition
(8.3) |
with some multipliers . The zero entries in the matrix stem from the fact that the derived subalgebra is a subrepresentation.
Suppose that has . Comparing coefficients in (8.3), we get , , and , so the matrix takes the form
The group of such matrices contains the diagonal copy of , and becomes our iterated semidirect product inside . Note that the projection admits a splitting, obtained by setting .
Seeking a contradiction, we assume that there is some with . By Lemma 8.2, we must have , and thus . Making a flat extension of , we can assume that there is some with . Setting and and left-multiplying with the resulting matrix , we may assume that both and hold. On the other hand, comparing coefficients in (8.3) immediately yields , giving a contradiction.
This establishes . We already observed that inside and that equals our iterated semidirect product. ∎
To continue inductively, we seek to relate the curves with different indices . First recall a general fact on numerical semigroups with non-zero generators : the blowing-up of the ring with respect to the maximal ideal has coordinate ring for the numerical semigroup ; cf. [BDF97, Proposition I.2.1, and also Equation I.2.4].
For our with , we have , with remaining generators and for . The resulting differences are and . For , we write and in any case see . This reveals the following.
Lemma 8.3.
For , we have , where the center is the singular point endowed with the reduced scheme structure.
Note that for every , we get an inclusion of numerical semigroups inside . The resulting inclusions of coordinate rings define canonical morphisms of compactifications of the affine line .
Proof of Theorem 8.1 in the general case.
We proceed by induction on . The case was handled above. Now suppose and that the assertion is true for . To simplify notation, set
and let be the inertia group scheme for the singularity . Likewise, we set . By induction, coincides with . Also note that by the very definition of the group schemes, we have a canonical inclusion . Moreover, with Proposition 5.5 we get an inclusion , and actually . By Lemma 8.3 combined with [Mar22, Proposition 2.7], the blowing-up morphism is equivariant with respect to the action of . We thus get inclusions and infer .
The orbit map for the rational point gives an inclusion of closed subschemes inside . This is actually contained in the scheme of singularities , according to [BS22, Proposition 3.1]. Our task is to show that the inclusion is an equality, and for this it suffices to verify that the coordinate rings have the same degree. We already saw that , and it remains to verify . For this, we may assume that is algebraically closed.
According to (6.3) and Proposition 6.3, the scheme of singularities has coordinate ring of the form . In light of [DG70, Section III.3, Theorem 6.1], the homogeneous space has coordinate ring of the form for some and certain exponents . From the canonical surjection
we infer and . Using the relation , we see that must be contained in , and thus actually . It follows that , as desired. ∎
9. Equivariant normality and twisting
We now seek to construct twisted forms of our curves that are regular. Our methods to achieve this apply in many other contexts, and we first give a general discussion about twisted forms, their regularity properties, and Brion’s recent notion of equivariant normality.
Fix a ground field , and let be a scheme. Recall that another scheme is called a twisted form of if we have for some field extension . Such twisted forms may arise as follows: Suppose that a group scheme acts on , and let be a -torsor. Then acts diagonally on , and the quotient
is a twisted form of . Note that the diagonal action is free; hence the quotient exists as an algebraic space. Such quotients are not necessarily schematic (for concrete examples, see [Sch22a]). However, if is finite and is covered by affine open sets that are -stable, the twisted form is indeed a scheme (cf. [DG70, Section III.2, Theorem 3.2]).
Now suppose that we are in positive characteristic . It then may happen that a noetherian scheme with singularities has twisted forms where all singularities are gone. We now describe a fairly general procedure to achieve this, relying on a combination of works of Brion and the second author [Bri22a, BS22, Sch07, Sch22b]. For simplicity, we assume throughout that is a separated scheme of finite type that is geometrically integral. It is normal if all local rings are integrally closed in the common function field . Equivalently, each finite modification is an isomorphism. Here the term modification refers to an integral scheme , together with a proper surjective morphism inducing a bijection on function fields.
In what follows, we suppose that is endowed with the action of a finite group scheme . We now consider only modifications where is a -scheme and is equivariant. For brevity, we call such a datum an equivariant modification or -modification. Examples are given by blowing-ups with respect to -stable centers . Note that for a given modification , there is at most one -action on making equivariant.
One says that is equivariantly normal, or -normal, if every finite equivariant modification is an isomorphism. This extremely useful notion was introduced and studied by Brion in [Bri22a]. He showed that admits a finite equivariant modification that is equivariantly normal (cf. [Bri22a, Proposition 4.2]). It is actually unique up to unique equivariant isomorphism provided that is one-dimensional (cf. [Bri22a, Corollary 4.4]).
From now on, we furthermore assume that our -scheme is one-dimensional. One could also say that is a -curve. Following [FS20, Section 2] we write for the closed subscheme defined by the first Fitting ideal for . This is the set of points where the local ring fails to be geometrically regular, endowed with a canonical scheme structure. Note that with respect to this scheme structure, it must be -stable (cf. [BS22, Proposition 3.1]). The existence of twisted forms that are regular is intimately related to equivariant normality.
Theorem 9.1.
Let be a -torsor. Then the twisted form is regular provided the following three conditions hold:
-
(i)
The curve is -normal.
-
(ii)
The total space of the -torsor is reduced.
-
(iii)
The reduction of the finite scheme is étale.
Proof.
The residue fields for the points are separable, by assumption (iii), and so is their join . This ensures that the base-change remains reduced. Furthermore, the arguments for [Bri22a, Proposition 4.10] show that remains equivariantly normal. Replacing the ground field with , we may assume that comprises only rational points. Let be the inertia subgroup scheme and be the orbit for . Clearly, the subscheme and its complementary open set are -stable. The latter is geometrically regular, and so is the twisted form .
It remains to verify that the integral curve is regular at the points . According to [Bri22a, Theorem 4.13], the orbit is an effective Cartier divisor. In turn, its twist is an effective Cartier divisor on , so it suffices to verify that it is reduced. The latter becomes the quotient of by the diagonal -action, which can be identified with . Its coordinate ring is a subring inside , which can be seen as a ring of invariants, and is reduced by assumption. ∎
Note that condition (iii) holds in particular if all are rational points. The first two conditions can be achieved after ground field extensions.
Proposition 9.2.
Suppose that the curve is -normal. Then there is a field extension such that the following hold:
-
(i)
The base-change is -normal.
-
(ii)
There is a -torsor whose total space is reduced.
Proof.
(ii) Choose a geometrically integral quasi-projective -scheme with generically free action. This could arise from an embedding into a smooth group scheme of finite type or could arise from a projective scheme with , according to [BS22, Proposition 1.7 or Theorem 2.1]. The quotient is an integral quasi-projective scheme, and the quotient map induces a finite extension of the function field by . By construction, the reduced scheme , viewed as an -scheme, is a torsor with respect to the base-change .
In the reverse direction, we have the following result.
Theorem 9.3.
Suppose that there exist a field extension , a subgroup scheme , and a -torsor so that the twisted form is regular. Then the curve is -normal.
Proof.
According to [Bri22a, Proposition 4.10 and Remark 4.3], it suffices to treat the case and . Let be the regular twisted form. According to [ST23, Lemma 3.1], we have a canonical identification of the sheaves of automorphisms, where the term on the right is formed with respect to the conjugacy action of on . Setting , we get a homomorphism , hence a -action on .
Let be a finite closed subscheme that is -stable. According to [Bri22a, Theorem 4.13], we have to check that is Cartier. Its twist defines a finite closed subscheme inside . The latter is regular, so is Cartier. Choose a point , and let be the resulting field extension. The resulting trivialization of defines an isomorphism with . It follows that , and hence , is Cartier. ∎
Recall that a quasielliptic curve is a regular curve that is a twisted form of the rational cuspidal curve
Clearly, is a singleton, containing only the rational point given by . It turns out that quasielliptic curves exist only in characteristic two and three (compare with the discussion after Proposition 11.1). For , the rational vector field satisfies and actually defines a global section , hence corresponds to an action of , which is the Frobenius kernel of the additive group . The orbit is the Cartier divisor defined by . For , the same holds for and the Cartier divisor .
In both cases, we conclude that the rational cuspidal curve is equivariantly normal with respect to (again by [Bri22a, Theorem 4.13]). For this group scheme, torsors with regular total space exist if and only if is imperfect (see for example [ST23, Lemma 7.1]), and then quasielliptic curves exist by Theorem 9.1. Also note that is equivariantly normal with respect to any larger finite subgroup scheme inside the full automorphism group scheme (obvious, see [Bri22a, Remark 4.3]). According to [BM77, Proposition 6], we have an iterated semidirect product for an infinitesimal group scheme . For , it coincides with the copy of described above, whereas for , it has order .
All this generalizes to our hierarchy of curves .
Theorem 9.4.
The curve is equivariantly normal with respect to the finite group scheme and locally of complete intersection. Moreover, if there is a -torsor so that the quotient is reduced, then the twisted form is regular.
Proof.
Note that after some ground field extension , there is a -torsor that is reduced, according to Proposition 9.2, and our curve acquires twisted forms that are regular. However, the construction of relies, via [BS22, Proposition 1.7 and Theorem 2.1], among other things on embeddings of into smooth group schemes , and here we have little control over and .
Also note that the above argument for locally complete intersection relying on equivariant normality is independent of the arguments relying on numerical semigroups in the proof of Proposition 6.1.
10. Non-abelian cohomology and semidirect products
In this section we review the general notions of torsors and twisting, which will be used in the next section to understand the twisted forms of our curves . The material is well known, but it is not easy to find suitable references that are general enough for our purposes, yet not burdened by over-abstraction. Throughout, we are guided by [Ser94, Section I.5] and [Gir71, Section III.2].
Let be the topos of sheaves on some site , having a final object . For any group-valued object , we write for the group of global sections and for the set of isomorphism classes of -torsors . The latter is an object endowed with a -action that is locally isomorphic to with the translation action. Another widespread term is principal homogeneous spaces. For commutative, our coincides with the sheaf cohomology groups. In general, however, is merely a set, containing the class of the trivial torsor as a distinguished element.
An object is called a twisted form of an object if the two are locally isomorphic. If has a -action and is a -torsor, we get such a twisted form by forming the quotient with respect to the diagonal action . Note that this could also be written as . For , the above construction gives an identification between the non-abelian cohomology and the set of isomorphism classes of twisted forms of the object . In any case, the -action on induces a conjugacy action on , and we have ; compare for example with [ST23, Lemma 3.1].
Now suppose as sheaves of sets without group laws, and consider the homomorphism given by . One easily checks that the map is equivariant with respect to factorwise conjugation and conjugation with inner automorphisms on . In turn, we get an induced homomorphism
Note that the equation stems from the identification . The above endows each -torsor with the additional structure of a -torsor and a -torsor. Furthermore, is the automorphism group object of as a -torsor, and is the automorphism group object of as a -torsor. In turn, we get what we like to call the torsor translation map
(10.1) |
where the quotient on the right is formed with respect to the action and the -action on stems from the action on the first factor . The map (10.1) is bijective but does not respect the distinguished points: rather, it sends to .
Now suppose that we have a short exact sequence
(10.2) |
of group objects. Then the group acts from the right on the set in the following way: For each global section , the fiber with respect to the surjection carries compatible -torsor structures from both sides, coming from the group law in . We now define . The stabilizer group at each torsor class is the subgroup of global sections where the set of global sections is non-empty.
Let us write for the quotient of the action. Using the distinguished point in , the orbit map yields . The latter serves as a connecting map and yields a six-term sequence of sets
The maps on the right come from extension of structure groups. Here all arrows preserve the distinguished points, and in degree zero the above is an exact sequence of groups.
The group object acts on itself and its quotient via conjugacy. On the normal subgroup , we have an induced action. Twisting with respect to the -actions gives another exact sequence
(10.3) |
which also yields a six-term sequence. Note that in general there is no map relating and because the -action on usually fails to be inner.
We now choose for each -torsor whose class belongs to the image of the mapping some -torsor with . As in [Ser94, Section 5.5, Corollary 2], one has the following.
Theorem 10.1.
The first cohomology of can be written as a disjoint union
running over all from the image of . The inclusions are obtained by composing the induced maps with the torsor translation maps given by (10.1).
We are interested in cases where the above simplifies. Recall that is the canonical epimorphism. Let us call a morphism a set-theoretical section if . The point here is that does not have to preserve the group laws. The resulting given by is an isomorphism of objects that does not necessarily respect the group laws. The latter is determined by the two-cocycle defined by . We like to indicate this situation by writing
and say that the extension is set-theoretically split. Note that this always holds in the category of groups but often fails in the category of group schemes (compare with [Sch23a, around Theorem 8.5]). Also note that this property is not necessarily preserved in twisted extensions (10.3). If respects the group laws, the above becomes a semidirect product , where is given by conjugacy, and the extension is called split.
Corollary 10.2.
Proof.
Set and . If is set-theoretically split, all fibers over are trivial torsors, and hence the action of on is trivial. The same holds, of course, if the group itself is trivial. So the theorem implies the first assertion.
If the projection admits a section that respects the group structure, the induced map on cohomology is right inverse to , so the latter is surjective. ∎
In particular, for satisfying the assumptions of the corollary, we get a disjoint union
running over all . It is convenient to regard its elements as “pairs” with and . If is a singleton, the choice of isomorphisms indeed identifies the above with the product , independently of the . Note that this carries a canonical group structure if and are commutative, which may happen without being commutative.
11. Description of the set of twisted forms
Let be a ground field of characteristic , and set . Using the general results of the previous section, we seek to compute the first non-abelian cohomology for the iterated semidirect products and thereby the set of isomorphism classes of twisted forms for . We are able to do so for .
Throughout, we work over the site , endowed with the fppf topology. Let us start with some general useful facts.
Proposition 11.1.
The following hold for group schemes of finite type:
-
(i)
If for some group scheme of finite type, then the canonical map is bijective.
-
(ii)
In the special case , the set is a singleton.
-
(iii)
If is infinitesimal, the group is trivial.
Proof.
Note that for , the automorphism group scheme for the rational cuspidal curve is given by , so this curve has no twisted forms besides itself. This purely cohomological argument shows again that quasielliptic curves are confined to characteristic . The following observation will also be useful.
Lemma 11.2.
Let and be twisted forms of . If they are isomorphic as schemes, they are isomorphic as group schemes.
Proof.
Let be an isomorphism of schemes. Composing with a translation, we may assume . To verify that respects the group law, we may assume , and this reduces to the case . The induced map on the coordinate ring is given by for some . We have because is non-constant, and because respects the origin. So for each , we have , which respects the group laws. ∎
For each pair of additive polynomials with (in other words, not both polynomials vanish), the resulting homomorphism given by is an epimorphism. The short exact sequence
(11.1) |
defines a unipotent group scheme , and the resulting long exact sequence yields
(11.2) |
By Russell’s theorem [Rus70, Theorem 2.1], every twisted form of the additive group is isomorphic to with for some , and separable. The following sheds further light on this.
Proposition 11.3.
The unipotent group scheme is a twisted form of if and only if inside the euclidean domain .
Proof.
It suffices to treat the case that is algebraically closed. Set . First suppose that there are with . These yield a section for , defined via , which does not have to preserve the group laws. It induces an identification of schemes. In turn, the coordinate ring has the property . According to Zariski cancellation (see [AHE72, Corollary 2.8]), the underlying scheme is isomorphic to the affine line . By Lazard’s theorem (see [DG70, Section IV.4, Theorem 4.1]), we must have as group schemes.
Conversely, suppose . Since is algebraically closed, we have and for some exponents and some finite subgroups . Their intersection is non-zero, by the assumption on the gcd. Consequently, we can write and for some additive polynomial of the form , with some non-zero . In turn, the projection factors over the morphism . So the kernel is disconnected or non-reduced. ∎
Proposition 11.4.
For each group scheme of the form , where is any infinitesimal group scheme of finite type, we have a canonical identification
for certain additive polynomials with .
Proof.
Roughly speaking, to understand this cohomology of , one has to understand the cohomology of and the dependence of the additive polynomials on the class . We now seek to unravel this with . For , the term in the middle becomes . The short exact sequence yields an identification for every . The dependence on the additive polynomials can be described as follows.
Proposition 11.5.
For the -torsor , the additive polynomials and give the twisted form .
Proof.
Set . By definition, the coordinate ring for is the subring of elements that are invariant under and , for all group elements and all rings . Clearly, and are such invariants, satisfying the relation . Its partial derivatives generate the unit ideal, and we conclude that the subring generated by and is regular and one-dimensional and can be identified with the residue class ring . The composite extension has degree , and the outer steps have degree . Consequently, . The assertion now follows from Lemma 11.2. ∎
To tackle the case , we use the short exact sequence
where the inclusion on the left is and the surjection on the right . Given , the finite scheme defined by
(11.3) |
carries a -action via the formula . One easily checks that this indeed takes values in , that it satisfies the axioms for group actions, and that the action is free and transitive. The induced -torsor is obtained from as a quotient by , in other words, by the action of . This yields . In turn, we get the description . It remains to express the twisted form in terms of additive polynomials.
Proposition 11.6.
For the -torsor as above, the additive polynomials
give the twisted form .
Proof.
Set . The coordinate ring for is the subring of elements that are invariant under
for all group elements and all rings . One easily checks that and are invariant and that these invariants satisfy the relation . The argument concludes as in the preceding proof. ∎
Note that the invariant can be found by starting with the non-invariant and successively adding monomials to cancel non-invariance. Collecting all the above, we have determined the non-abelian cohomology for in the cases .
Theorem 11.7.
With the above notation, we have
where the union runs over all , and
where the union runs over with and .
Note that for , this gives back, in an intrinsic fashion, Queen’s descriptions for quasielliptic curves (cf. [Que71, Que72]).
Also note that for , the group is trivial, provided that is separably closed or . It follows that the Frobenius map
is trivial, in the sense that it sends every class to the distinguished class. In particular, every twisted form of is untwisted by Frobenius pullback and becomes a rational curve (compare with [HS22]). Likewise, for , the group is trivial if is separably closed or , . Now the map is trivial, and every twisted form of gets untwisted by the second Frobenius pullback.
With the notation from the theorem, let be a -torsor and be the ensuing class, for . Write for the twisted form of corresponding to .
Proposition 11.8.
With the above notation, the curve is regular provided that does not belong to .
Proof.
The -torsor is induced from some torsor with respect to , according to Proposition 11.1. By construction, the class corresponds to the quotient , and the latter has coordinate ring , where we also write for the scalar rather than the class. The coordinate ring is reduced, in light of our assumption. According to Theorem 9.4, the curve is regular. ∎
It should be possible to extend the above results to all . For this, one has to find an inductive description for the -torsors, analogous to (11.3). The main problem is to cope with the non-commutativity involved in the torsors.
References
- [AHE72] S. Abhyankar, W. Heinzer, and P. Eakin, On the uniqueness of the coefficient ring in a polynomial ring, J. Algebra 23 (1972), 310–342.
- [AGS16] A. Assi, P. García-Sánchez, Numerical semigroups and applications, Springer, Cham, 2016.
- [Ati71] M. Atiyah, Riemann surfaces and spin structures, Ann. Sci. École Norm. Sup. 4 (1971), 47–62.
- [BDF97] V. Barucci, D. Dobbs, and M. Fontana, Maximality properties in numerical semigroups and applications to one-dimensional analytically irreducible local domains, Mem. Amer. Math. Soc. 125 (1997), no. 598.
- [BM76] E. Bombieri and D. Mumford, Enriques’ classification of surfaces in char. , III, Invent. Math. 35 (1976), 197–232.
- [BM77] by same author, Enriques’ classification of surfaces in char. , II, in: Complex analysis and algebraic geometry (W. Baily and T. Shioda, eds), pp. 23–42, Cambridge Univ. Press, London, 1977.
- [BCP97] W. Bosma, J. Cannon, and C. Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), 235–265.
- [Bri17] M. Brion, Some structure theorems for algebraic groups, in: Algebraic groups: structure and actions (M. Can, ed.), pp. 53–126. Amer. Math. Soc., Providence, RI, 2017.
- [Bri22a] by same author, Actions of finite group schemes on curves, preprint arXiv:2207.08209 (2022).
- [Bri22b] by same author, On models of algebraic group actions, Proc. Indian Acad. Sci. Math. Sci. 132 (2022), Paper No. 61.
- [BS22] M. Brion and S. Schröer, The inverse Galois problem for connected algebraic groups, preprint arXiv:2205.08117 (2022).
- [Del76] C. Delorme, Sous-monoïdes d’intersection compléte de N, Ann. Sci. École Norm. Sup. 9 (1976), 145–154.
- [DG70] M. Demazure and P. Gabriel, Groupes algébriques, Masson, Paris, 1970.
- [SGA3-1] M. Demazure and A. Grothendieck (eds), Séminaire de Géométrie algébrique du Bois-Marie, 1962–1964 (SGA3), Schémas en groupes, Tome 1, Lecture Notes in Math., vol. 151, Springer-Verlag, Berlin-New York, 1970.
- [FS20] A. Fanelli and S. Schröer, Del Pezzo surfaces and Mori fiber spaces in positive characteristic, Trans. Amer. Math. Soc. 373 (2020), 1775–1843.
- [GAP] The GAP Group, GAP – Groups, Algorithms, and Programming, Version 4.12.1, 2022, https://www.gap-system.org.
- [Gir71] J. Giraud, Cohomologie non abélienne, Springer, Berlin, 1971.
- [GW04] K. Goodearl and R. Warfield, An introduction to noncommutative Noetherian rings, 2nd edn, Cambridge Univ. Press, Cambridge, 2004.
- [Gos96] D. Goss, Basic structures of function field arithmetic, Springer, Berlin, 1996.
- [Hal59] M. Hall, The theory of groups, Macmillan, New York, 1959.
- [Her70] J. Herzog, Generators and relations of abelian semigroups and semigroup rings, Manuscripta Math. 3 (1970), 175–193.
- [HS22] C. Hilario and K.-O. Stöhr, On regular but non-smooth integral curves, preprint arXiv:2211.16962 (2022).
- [Jac43] N. Jacobson, The theory of rings, Amer. Math. Soc., New York, 1943.
- [KM79] M. Kargapolov and J. Merzljakov, Fundamentals of the theory of groups, Springer, New York-Berlin, 1979.
- [KS21] S. Kondō and S. Schröer, Kummer surfaces associated with group schemes, Manuscripta Math. 166 (2021), 323–342.
- [Lau19] B. Laurent, Almost homogeneous curves over an arbitrary field, Transform. Groups 24 (2019), 845–886.
- [Mar22] G. Martin, Infinitesimal automorphisms of algebraic varieties and vector fields on elliptic surfaces, Algebra Number Theory 16 (2022), 1655–1704.
- [Mil80] J. Milne, Étale cohomology, Princeton Univ. Press, Princeton, 1980.
- [Mum71] D. Mumford, Theta characteristics of an algebraic curve, Ann. Sci. École Norm. Sup. 4 (1971), 181–192.
- [Ore33] O. Ore, On a special class of polynomials, Trans. Amer. Math. Soc. 35 (1933), 559–584.
- [Que71] C. Queen, Non-conservative function fields of genus one I, Arch. Math. 22 (1971), 612–623.
- [Que72] by same author, Non-conservative function fields of genus one II, Arch. Math. 23 (1972), 30–37.
- [Rob93] D. Robinson, A course in the theory of groups, Springer, New York, 1993.
- [RGS09] J. Rosales and P. García-Sánchez, Numerical semigroups, Springer, New York, 2009.
- [Rus70] P. Russell, Forms of the affine line and its additive group, Pacific J. Math. 32 (1970), 527–539.
- [Sch07] S. Schröer, Kummer surfaces for the selfproduct of the cuspidal rational curve, J. Algebraic Geom. 16 (2007), 305–346.
- [Sch22a] by same author, Algebraic spaces that become schematic after ground field extension, Math. Nachr. 295 (2022), 1008–1012
- [Sch22b] by same author, The structure of regular genus-one curves over imperfect fields, preprint arXiv:2211.04073 (2022).
- [Sch23a] by same author, Albanese maps for open algebraic spaces, Int. Math. Res. Not. IMRN, published online on September 9, 2023, to appear in print.
- [Sch23b] by same author, There is no Enriques surface over the integers, Ann. of Math. 197 (2023), 1–63.
- [ST23] S. Schröer and N. Tziolas, The structure of Frobenius kernels for automorphism group schemes, Algebra Number Theory 17 (2023), 1637-–1680.
- [Ser94] J.-P. Serre, Cohomologie galoisienne, 5th edn, Lecture Notes in Math., vol. 5, Springer, Berlin, 1994.
- [Sta18] The Stacks project authors, Stacks project, https://stacks.math.columbia.edu, 2018.