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

Generalizations of quasielliptic curves

Cesar Hilario Mathematisches Institut, Heinrich-Heine-Universität, 40204 Düsseldorf, Germany [email protected]  and  Stefan Schröer Mathematisches Institut, Heinrich-Heine-Universität, 40204 Düsseldorf, Germany [email protected]
    • 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 KK be a ground field of characteristic p>0p>0. The goal of this paper is to generalize, in an equivariant way, the rational cuspidal curve

(1.1) X=SpecK[T2,T3]SpecK[T1]X=\operatorname{Spec}K[T^{2},T^{3}]\cup\operatorname{Spec}K[T^{-1}]

from the cases p=2p=2 and p=3p=3, when the automorphism group scheme is non-reduced, to a hierarchy of integral curves Xp,nX_{p,n} whose automorphism group schemes are likewise non-reduced. Here the index p>0p>0 indicates the characteristic, and pn(n+1)/2p^{n(n+1)/2} gives the “size” of non-reducedness.

Our motivation originates from the Enriques classification of algebraic surfaces over ground fields k=kalgk=k^{\text{\rm alg}}: This vast body of theorems on the structure of surfaces SS 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 f:SBf\colon S\rightarrow B whose generic fiber Y=f1(η)Y=f^{-1}(\eta) is a so-called quasielliptic curve, in other words, a twisted form of (1.1) over the function field K=k(B)K=k(B), with all local rings 𝒪Y,y\mathscr{O}_{Y,y} regular. One knows that such twisted forms exist only over imperfect fields of characteristic p3p\leq 3, and Queen gave explicit equations for them (see [Que71, Que72]), although of rather extrinsic nature.

We discovered the hierarchy X=Xp,nX=X_{p,n} 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 AutX/K\operatorname{Aut}_{X/K}. Its crucial part consists of certain infinitesimal group schemes UnU_{n} of order pn(n+1)/2p^{n(n+1)/2} acting in a canonical way on the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}]. The underlying scheme is αpn×αpn1××αp\alpha_{p^{n}}\times\alpha_{p^{n-1}}\times\ldots\times\alpha_{p}, 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 i=0nλiTpi\sumop\displaylimits_{i=0}^{n}\lambda_{i}T^{-p^{i}}, or equivalently the elements of the skew polynomial ring R[F;σ]R[F;\sigma], formed over rings with nilpotent elements.

According to Brion’s recent theory of equivariantly normal curves, see [Bri22a], there is a unique compactification A1Xp,n\mathbb{A}^{1}\subset X_{p,n} to which the action of UnU_{n} 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

=p,npn,pnpn1,,pnp0N,{}_{p,n}=\left\langle p^{n},p^{n}-p^{n-1},\ldots,p^{n}-p^{0}\right\rangle\subset\mathbb{N},

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 UnU_{n}-action on the affine line extends to the compactification

Xp,n=SpecK[Tp,n]SpecK[T1],X_{p,n}=\operatorname{Spec}K\left[T^{{}_{p,n}}\right]\cup\operatorname{Spec}K\left[T^{-1}\right],

and this projective curve is equivariantly normal with respect to the UnU_{n}-action.

For 3pn43\leq p^{n}\leq 4, this is precisely the rational cuspidal curve. The second main result unravels the numerical invariants and infinitesimal symmetries of this hierarchy of projective curves.

Theorem (See Section 6 and Theorem 8.1).

The curves X=Xp,nX=X_{p,n} have

h1(𝒪X)=12(npn+1(n+2)pn+2)andAutX/K=GaUnGm.h^{1}(\mathscr{O}_{X})=\tfrac{1}{2}(np^{n+1}-(n+2)p^{n}+2)\quad\text{and}\quad\operatorname{Aut}_{X/K}=\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}.

In this iterated semidirect product, the additive group Ga\mathbb{G}_{a} is normalized by the infinitesimal group scheme UnU_{n}, and both are normalized by the multiplicative group Gm\mathbb{G}_{m}. Note that for 3pn43\leq p^{n}\leq 4, 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 X=Xp,nX=X_{p,n}: The tangent sheaf =X/KHom¯(,X/K1𝒪X){}_{X/K}=\underline{\operatorname{Hom}}({}^{1}_{X/K},\mathscr{O}_{X}) turns out to be invertible, actually very ample, giving a canonical inclusion XP(g)=Pn+1X\subset\mathbb{P}(\mathfrak{g})=\mathbb{P}^{n+1}, where g=H0(X,)X/K\mathfrak{g}=H^{0}(X,{}_{X/K}) is the Lie algebra of the automorphism group scheme G=AutX/KG=\operatorname{Aut}_{X/K}. From the canonical linearization 𝒪X(1)=X/K\mathscr{O}_{X}(1)={}_{X/K}, we get a matrix representation for GG, which is crucial to gain control over its structure. Furthermore, XX is globally a complete intersection, defined inside Pn+1\mathbb{P}^{n+1} by the following nn homogeneous equations of degree pp:

Un1pVp1Z=0andUjpVp1Uj+1=0(0jn2).U_{n-1}^{p}-V^{p-1}Z=0\quad\text{and}\quad U_{j}^{p}-V^{p-1}U_{j+1}=0\quad(0\leq j\leq n-2).

Also note that the curves are related by a hierarchy of blowing-ups Xp,n1=BlZ(Xp,n)X_{p,n-1}=\operatorname{Bl}_{Z}(X_{p,n}), 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 X=Xp,nX=X_{p,n} have, over ground fields KK with “enough” imperfection, twisted forms YY where all local rings 𝒪Y,y\mathscr{O}_{Y,y} are regular (cf. Theorem 9.4). These have the same structural properties of XX, except that the singularities get “twisted away.” In turn, the passage from the rational cuspidal curve to quasielliptic curves is generalized to our hierarchy X=Xp,nX=X_{p,n}.

The above relies on rather general observations, which form the third main result of this paper.

Theorem (See Section 9).

Let XX be a geometrically integral curve with the action of a finite group scheme GG. Suppose that  Sing(X/K)red\operatorname{Sing}(X/K)_{\operatorname{red}} is étale. Then XX is equivariantly normal if and only if for some field extension KLK\subset L, the base-change XLX\otimes L 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 X=Xp,nX=X_{p,n} 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 Twist(X)\operatorname{Twist}(X) of isomorphism classes of twisted forms over S=Spec(K)S=\operatorname{Spec}(K) as non-abelian cohomology H1(S,AutX/S)H^{1}(S,\operatorname{Aut}_{X/S}). For our curves X=Xp,nX=X_{p,n}, 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 n2n\leq 2. Our fourth main result is as follows.

Theorem (See Theorem 11.7).

For G=GaU2GmG=\mathbb{G}_{a}\rtimes U_{2}\rtimes\mathbb{G}_{m}, the non-abelian cohomology is

H1(S,G)=K/{up2vαvpβpvp2u,vK},H^{1}(S,G)=\bigcupop\displaylimits K/\left\{u^{p^{2}}-v-\alpha v^{p}-\beta^{p}v^{p^{2}}\mid u,v\in K\right\},

where the union runs over (α,β)K/Kp2K/Kp(\alpha,\beta)\in\bigcupop\displaylimits_{K/K^{p^{2}}}K/K^{p}, with αK/Kp2\alpha\in K/K^{p^{2}} and βK/Kp\beta\in K/K^{p}.

Perhaps this is the first explicit determination of Twist(X)\operatorname{Twist}(X) via a purely non-abelian cohomological computation of H1(S,AutX/S)H^{1}(S,\operatorname{Aut}_{X/S}) for some relevant hierarchy of schemes XX. A crucial step in this is the determination of particular twisted forms GaP{}^{P}\!\mathbb{G}_{a} 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 p3p\leq 3 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 Un(R)U_{n}(R). 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 UnU_{n}. In Section 5 we examine the equivariant compactifications of the affine line A1\mathbb{A}^{1} and introduce our numerical semigroup p,n and the ensuing curve Xp,nX_{p,n}, 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 Xp,nPn+1X_{p,n}\subset\mathbb{P}^{n+1}. 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 Xp,nX_{p,n}. 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 Xp,nX_{p,n} in the cases n=1n=1 and n=2n=2.

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 U=UnU=U_{n} in Section 4. Fix some ring RR of characteristic p>0p>0, and let xx be an indeterminate. Recall that polynomials of the form

P(x)=i=0nλixpi=λ0x+λ1xp++λnxpnR[x]P(x)=\sumop\displaylimits_{i=0}^{n}\lambda_{i}x^{p^{i}}=\lambda_{0}x+\lambda_{1}x^{p}+\ldots+\lambda_{n}x^{p^{n}}\in R[x]

are called additive polynomials. Another widespread designation is pp-polynomials. Clearly, the set of all such polynomials is stable under addition P(x)+Q(x)P(x)+Q(x) and substitution P(Q(x))P(Q(x)). 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 P(x)=0P(x)=0 and unit element P(x)=xP(x)=x. Let us call it the ring of additive polynomials.

It can also be seen as the skew polynomial ring R[F;σ]R[F;\sigma], where σ:RR\sigma\colon R\rightarrow R designates the Frobenius map λλp\lambda\mapsto\lambda^{p}. Elements are polynomials in the formal symbol FF, and multiplication is subject to the relations Fλ=λpFF\lambda=\lambda^{p}F. In other words, we have

(2.1) iλiFijμjFj=k(i+j=kλiμjpi)Fk,\sumop\displaylimits_{i}\lambda_{i}F^{i}\cdot\sumop\displaylimits_{j}\mu_{j}F^{j}=\sumop\displaylimits_{k}\left(\sumop\displaylimits_{i+j=k}\lambda_{i}\mu_{j}^{p^{i}}\right)F^{k},

a modification of the usual Cauchy multiplication. The identification of the skew polynomial ring with the ring of additive polynomials is given by λλx\lambda\mapsto\lambda x and FxpF\mapsto x^{p}, so that λiFi\sumop\displaylimits\lambda_{i}F^{i} corresponds to λixpi\sumop\displaylimits\lambda_{i}x^{p^{i}}. 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 R[F;σ]R[F;\sigma] are characterized as in usual polynomial rings.

Proposition 2.1.

An element i=0nλiFi\sumop\displaylimits_{i=0}^{n}\lambda_{i}F^{i} of the skew polynomial ring R[F;σ]R[F;\sigma] is nilpotent if and only if λ0,,λnNil(R)\lambda_{0},\ldots,\lambda_{n}\in\operatorname{Nil}(R).

Proof.

Suppose that all coefficients are nilpotent, say λid=0\lambda_{i}^{d}=0. For each r0r\geq 0, we have

(iλiFi)r=k(i1++ir=kλi1v1λirvr)Fk\left(\sumop\displaylimits_{i}\lambda_{i}F^{i}\right)^{r}=\sumop\displaylimits_{k}\left(\sumop\displaylimits_{i_{1}+\ldots+i_{r}=k}\lambda_{i_{1}}^{v_{1}}\cdots\lambda_{i_{r}}^{v_{r}}\right)F^{k}

for certain exponents v1,,vr1v_{1},\ldots,v_{r}\geq 1 whose precise values are irrelevant in the following reasoning: If r>(n+1)(d1)r>(n+1)(d-1), each tuple 0i1,,irn0\leq i_{1},\ldots,i_{r}\leq n must contain the dd-fold repetition of at least one value 0in0\leq i\leq n. Then the product λi1v1λirvr\lambda_{i_{1}}^{v_{1}}\cdots\lambda_{i_{r}}^{v_{r}} vanishes, and so does the above rr-fold power.

Conversely, suppose that some λsR\lambda_{s}\in R is not nilpotent. Choose a prime pR\mathfrak{p}\subset R not containing λs\lambda_{s}, and set K=κ(p)K=\kappa(\mathfrak{p}). Then the image of i=0nλiFi\sumop\displaylimits_{i=0}^{n}\lambda_{i}F^{i} in the skew polynomial ring K[F;σ]K[F;\sigma] is a non-zero nilpotent element. On the other hand, K[F;σ]K[F;\sigma] 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 P=i=0nλiFiP=\sumop\displaylimits_{i=0}^{n}\lambda_{i}F^{i} of the skew polynomial ring R[F;σ]R[F;\sigma] is invertible if and only if λ0R×\lambda_{0}\in R^{\times} and λ1,,λnNil(R)\lambda_{1},\ldots,\lambda_{n}\in\operatorname{Nil}(R).

Proof.

The condition is sufficient: Set μi=λi/λ0\mu_{i}=-\lambda_{i}/\lambda_{0}. By Proposition 2.1, the element Q=i=1nμiFiQ=\sumop\displaylimits_{i=1}^{n}\mu_{i}F^{i} is nilpotent, say Qr=0Q^{r}=0. Then 1Q1-Q is a unit, with inverse j=0r1Qj\sumop\displaylimits_{j=0}^{r-1}Q^{j}. Thus P=λ0(1Q)P=\lambda_{0}(1-Q) is also a unit.

Conversely, suppose PQ=QP=1PQ=QP=1. From the group law (2.1), one immediately infers that λ0R×\lambda_{0}\in R^{\times}. Seeking a contradiction, we assume that some λsR\lambda_{s}\in R, s1s\geq 1 is not nilpotent. Choose such 1sn1\leq s\leq n maximal. As above, we find some residue field K=κ(p)K=\kappa(\mathfrak{p}) in which λs\lambda_{s} is non-zero. Let d0d\geq 0 be the degree of the image of QQ. From (2.1) one sees that the image of 1=PQ1=PQ has non-zero term in degree s+ds+d, giving a contradiction. ∎

Given a unit of the form P=λiFiP=\sumop\displaylimits\lambda_{i}F^{i}, the inverse P1=μjFjP^{-1}=\sumop\displaylimits\mu_{j}F^{j} can be computed as follows: The condition PP1=1P\cdot P^{-1}=1 means λ0μ0p0=1\lambda_{0}\mu_{0}^{p^{0}}=1 and i+j=kλiμjpi=0\sumop\displaylimits_{i+j=k}\lambda_{i}\mu_{j}^{p^{i}}=0 for k1k\geq 1, which give the recursion formula

(2.2) μ0=λ01andμk=1λ0i=1kλiμkipi(k1).\mu_{0}=\lambda_{0}^{-1}\quad\text{and}\quad\mu_{k}=-\frac{1}{\lambda_{0}}\sumop\displaylimits_{i=1}^{k}\lambda_{i}\mu_{k-i}^{p^{i}}\quad(k\geq 1).

The skew polynomial ring comes with an infinite-dimensional matrix representation R[F;σ]Mat(R)R[F;\sigma]\rightarrow\operatorname{Mat}_{\infty}(R), already determined by the assignments

λ(λλpλp2...)andF(010101......).\lambda\longmapsto\begin{pmatrix}\lambda\\ &\lambda^{p}\\ &&\lambda^{p^{2}}\\ &&&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}\end{pmatrix}\quad\text{and}\quad F\longmapsto\begin{pmatrix}0&1\\ &0&1\\ &&0&1\\ &&&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}\end{pmatrix}.

More explicitly, this homomorphism is given by

(2.3) i=0nλiFi(λsrpr)0rs<=(λ0λ1λ2λ0pλ1pλ2pλ0p2λ1p2λ2p2.........).\sumop\displaylimits_{i=0}^{n}\lambda_{i}F^{i}\longmapsto\left(\lambda^{p^{r}}_{s-r}\right)_{0\leq r\leq s<\infty}=\begin{pmatrix}\lambda_{0}&\lambda_{1}&\lambda_{2}&\cdots\\ &\lambda_{0}^{p}&\lambda_{1}^{p}&\lambda_{2}^{p}&\cdots\\ &&\lambda_{0}^{p^{2}}&\lambda_{1}^{p^{2}}&\lambda_{2}^{p^{2}}&\cdots\\ &&&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}&\mathinner{\mkern 1.0mu\raise 7.0pt\vbox{\kern 7.0pt\hbox{$.$}}\mkern 2.0mu\raise 4.0pt\hbox{$.$}\mkern 2.0mu\raise 1.0pt\hbox{$.$}\mkern 1.0mu}\end{pmatrix}.

Obviously, the map is injective and takes values in the row-finite upper triangular matrices. Note that for each d0d\geq 0, the top left submatrix indexed by 0r,sd10\leq r,s\leq d-1 yields a subrepresentation R[F;σ]Matd(R)R[F;\sigma]\rightarrow\operatorname{Mat}_{d}(R).

We now examine the unit group R[F;σ]×R[F;\sigma]^{\times} in more detail, for the time being as an abstract group. It comes with matrix representations R[F;σ]×GLd(R)R[F;\sigma]^{\times}\rightarrow\operatorname{GL}_{d}(R), d0d\geq 0. Note that this factors over the group of invertible upper triangular matrices Td(R)GLd(R)T_{d}(R)\subset\operatorname{GL}_{d}(R). Write Fild\operatorname{Fil}^{d} for the kernels. Clearly Fil0=R[F;σ]×\operatorname{Fil}^{0}=R[F;\sigma]^{\times}, whereas

Fild={1+i=dnλiFind and λiNil(R)}(d1).\operatorname{Fil}^{d}=\left\{1+\sumop\displaylimits_{i=d}^{n}\lambda_{i}F^{i}\mid\text{$n\geq d$ and $\lambda_{i}\in\operatorname{Nil}(R)$}\right\}\quad(d\geq 1).

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 Fild\operatorname{Fil}^{d} on R[F;σ]×R[F;\sigma]^{\times} has quotients

Fil0/Fil1=R×andFild/Fild+1=Nil(R)(d1).\operatorname{Fil}^{0}/\operatorname{Fil}^{1}=R^{\times}\quad\text{and}\quad\operatorname{Fil}^{d}/\operatorname{Fil}^{d+1}=\operatorname{Nil}(R)\quad(d\geq 1).

Moreover, we have the commutator formula [Fil1,Fild]Fild+1[\operatorname{Fil}^{1},\operatorname{Fil}^{d}]\subset\operatorname{Fil}^{d+1} for all d1d\geq 1.

Proof.

The first assertion is an immediate consequence of the group law (2.1). The commutator formula follows from a corresponding commutator formula for the unitriangular group UTd(R)GLd(R)\operatorname{UT}_{d}(R)\subset\operatorname{GL}_{d}(R) comprising upper triangular matrices with the unit element on the diagonal (cf. [KM79, Chapter 6, Example 16.1.2]). ∎

The multiplicative character R[F;σ]×GL1(R)=R×R[F;\sigma]^{\times}\rightarrow\operatorname{GL}_{1}(R)=R^{\times} given by λiFiλ0\sumop\displaylimits\lambda_{i}F^{i}\mapsto\lambda_{0} comes with a canonical splitting μμF0\mu\mapsto\mu F^{0}, so we get a semidirect product R[F;σ]×=Fil1R×R[F;\sigma]^{\times}=\operatorname{Fil}^{1}\rtimes R^{\times}. The ensuing conjugacy action of R×R^{\times} is given by

μ(λiFi)μ1=μ1piλiFi.\mu\left(\sumop\displaylimits\lambda_{i}F^{i}\right)\mu^{-1}=\sumop\displaylimits\mu^{1-p^{i}}\lambda_{i}F^{i}.

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 n0n\geq 0, the set

Un(R)={1+i=1nλiFiλipni+1=0 for all 1in}U_{n}(R)=\left\{1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}F^{i}\mid\text{$\lambda_{i}^{p^{n-i+1}}=0$ for all $1\leq i\leq n$}\right\}

is a subgroup inside the unit group R[F;σ]×R[F;\sigma]^{\times}, which is normalized by R×R[F;σ]×R^{\times}\subset R[F;\sigma]^{\times}.

Proof.

Clearly the set contains the unit element. Suppose that P=1+i=1nλiFiP=1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}F^{i} and Q=1+j=1nμjFjQ=1+\sumop\displaylimits_{j=1}^{n}\mu_{j}F^{j} belong to Un(R)U_{n}(R), and write the product as PQ=1+k=1mαkFkPQ=1+\sumop\displaylimits_{k=1}^{m}\alpha_{k}F^{k}, with coefficients αk=i+j=kλiμjpi\alpha_{k}=\sumop\displaylimits_{i+j=k}\lambda_{i}\mu_{j}^{p^{i}}. For kn+1k\geq n+1, each summand λiμjpi\lambda_{i}\mu_{j}^{p^{i}} vanishes: If jn+1j\geq n+1, we already have μj=0\mu_{j}=0, and if jnj\leq n, we get i=kjn+1ji=k-j\geq n+1-j and thus μjpi=0\mu_{j}^{p^{i}}=0. For knk\leq n, we have αkpnk+1=i+j=kλipnk+1μjpnj+1\alpha_{k}^{p^{n-k+1}}=\sumop\displaylimits_{i+j=k}\lambda_{i}^{p^{n-k+1}}\mu_{j}^{p^{n-j+1}}, which vanishes because μjpnj+1=0\mu_{j}^{p^{n-j+1}}=0. Thus PQUn(R)PQ\in U_{n}(R).

Next consider the inverse element P1=j0βjFjP^{-1}=\sumop\displaylimits_{j\geq 0}\beta_{j}F^{j}. The recursion formula (2.2) gives β0=1\beta_{0}=1 and βk=i=1kλiβkipi\beta_{k}=-\sumop\displaylimits_{i=1}^{k}\lambda_{i}\beta_{k-i}^{p^{i}} for k1k\geq 1. For kn+1k\geq n+1, each summand λiβkipi\lambda_{i}\beta_{k-i}^{p^{i}} vanishes because in(ki)+1i\geq n-(k-i)+1. For knk\leq n, we have (λiβkipi)pnk+1=λipnk+1βkipn(ki)+1(\lambda_{i}\beta_{k-i}^{p^{i}})^{p^{n-k+1}}=\lambda_{i}^{p^{n-k+1}}\beta_{k-i}^{p^{n-(k-i)+1}}, where the second factor vanishes. Thus P1Un(R)P^{-1}\in U_{n}(R).

Finally, for each μR×\mu\in R^{\times}, we have μPμ1=1+i=1nμ1piλiFi\mu\cdot P\cdot\mu^{-1}=1+\sumop\displaylimits_{i=1}^{n}\mu^{1-p^{i}}\lambda_{i}F^{i}, which clearly belongs to Un(R)U_{n}(R). So the latter is normalized by R×R^{\times}. ∎

3.  Actions on polynomial rings

We keep the set-up as in the previous section. Obviously, the multiplicative monoid of additive polynomials λixpi\sumop\displaylimits\lambda_{i}x^{p^{i}} acts on the polynomial ring R[x]R[x] via substitution of the indeterminate, in other words by P(x)P(λixpi)P(x)\mapsto P(\sumop\displaylimits\lambda_{i}x^{p^{i}}), and one easily checks that this is an action from the right. In turn, we have a group action R[x]×R[F;σ]×R[x]R[x]\times R[F;\sigma]^{\times}\rightarrow R[x] from the right, given by

(3.1) Q(x)λiFi=Q(λixpi).Q(x)\ast\sumop\displaylimits\lambda_{i}F^{i}=Q\left(\sumop\displaylimits\lambda_{i}x^{p^{i}}\right).

Note that the action of the multiplicative group Gm(R)=R×\mathbb{G}_{m}(R)=R^{\times} via Q(x)λ0=Q(λ0x)Q(x)\ast\lambda_{0}=Q(\lambda_{0}x) is a special case of this. Furthermore, we have the translation action of the additive group Ga(R)=R\mathbb{G}_{a}(R)=R, defined by

(3.2) Q(x)α=Q(x+α).Q(x)\ast\alpha=Q(x+\alpha).

Obviously, these actions are faithful, and we arrive at inclusions of Ga(R)\mathbb{G}_{a}(R) and R[F;σ]×R[F;\sigma]^{\times} into the opposite automorphism group of R[x]R[x].

Proposition 3.1.

Inside the opposite automorphism group of R[x]R[x], the group Ga(R)\mathbb{G}_{a}(R) is normalized by R[F;σ]×R[F;\sigma]^{\times}, and the intersection Ga(R)R[F;σ]×\mathbb{G}_{a}(R)\cap R[F;\sigma]^{\times} is trivial.

Proof.

Suppose that we have elements

αGa(R)andλiFiR[F;σ]×.\alpha\in\mathbb{G}_{a}(R)\quad\text{and}\quad\sumop\displaylimits\lambda_{i}F^{i}\in R[F;\sigma]^{\times}.

For the first assertion, it suffices to check that PGa(R)=Ga(R)PP\cdot\mathbb{G}_{a}(R)=\mathbb{G}_{a}(R)\cdot P. This indeed holds because one computes λi(x+α)pi=(λixpi)+α\sumop\displaylimits\lambda_{i}(x+\alpha)^{p^{i}}=(\sumop\displaylimits\lambda_{i}x^{p^{i}})+\alpha^{\prime} with α=i=0nλiαpi\alpha^{\prime}=\sumop\displaylimits_{i=0}^{n}\lambda_{i}\alpha^{p^{i}}. It remains to verify the assertion on the intersection. Suppose α=P\alpha=P as automorphisms of R[x]R[x]; in other words, x+α=λixpix+\alpha=\sumop\displaylimits\lambda_{i}x^{p^{i}}. Comparing coefficients at the constant terms gives α=0\alpha=0; hence the intersection Ga(R)R[F;σ]×\mathbb{G}_{a}(R)\cap R[F;\sigma]^{\times} is trivial. ∎

In turn, we get an inclusion of Ga(R)R[F;σ]×\mathbb{G}_{a}(R)\rtimes R[F;\sigma]^{\times} into the opposite automorphism group of R[x]R[x]. Later, we seek to extend part of this action to certain subrings of R[x1]R[x^{-1}] in a compatible way. The following observation will be useful: Let SR[x]S\subset R[x] be the multiplicative system of all monic polynomials. The resulting localization is denoted by R(x)=S1R[x]R(x)=S^{-1}R[x]. Since monic polynomials are regular elements from the polynomial ring, the localization map is injective, and we get an inclusion R[x]R(x)R[x]\subset R(x).

Proposition 3.2.

The action from the right of the group Ga(R)R[F;σ]×\mathbb{G}_{a}(R)\rtimes R[F;\sigma]^{\times} on the polynomial ring R[x]R[x] uniquely extends to R(x)R(x).

Proof.

Uniqueness immediately follows from the universal property of localizations. To see existence, consider the larger multiplicative system S~R[x]\tilde{S}\subset R[x] comprising the polynomials of the form λP+Q\lambda P+Q with PP monic, QQ nilpotent, and λR×\lambda\in R^{\times}. Obviously, this system is stable with respect to the actions (3.1) and (3.2), and we thus get an induced action on S~1R[x]\tilde{S}^{-1}R[x]. On the other hand, the inclusion SS~S\subset\tilde{S} gives a canonical map S1R[x]S~1R[x]S^{-1}R[x]\rightarrow\tilde{S}^{-1}R[x]. It remains to verify that every λP+Q\lambda P+Q as above becomes invertible in S1R[x]S^{-1}R[x]. Indeed, in the factorization λP+Q=P/1(λ+Q/P)\lambda P+Q=P/1\cdot(\lambda+Q/P), the second factor also is a unit because λ\lambda is invertible and Q/PQ/P is nilpotent. ∎

4.  Scheme-theoretic reinterpretation

Fix a ground field KK of characteristic p>0p>0. 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

Un(R)=Un,K(R)={1+i=1nλiFiR[T;σ]× λipni+1=0 for 1in}U_{n}(R)=U_{n,K}(R)=\left\{1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}F^{i}\in R[T;\sigma]^{\times}\mid\text{ $\lambda_{i}^{p^{n-i+1}}=0$ for $1\leq i\leq n$}\right\}

as a group-valued functor UnU_{n} on the category of KK-algebras RR. Clearly, the natural transformation

(4.1) αpn×αpn1××αpUn,(λ1,,λn)1+i=1nλiFi\alpha_{p^{n}}\times\alpha_{p^{n-1}}\times\cdots\times\alpha_{p}\longrightarrow U_{n},\quad(\lambda_{1},\ldots,\lambda_{n})\longmapsto 1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}F^{i}

is an isomorphism of set-valued functors, with group laws ignored. In turn, UnU_{n} is a finite group scheme with coordinate ring i=1nK[xi]/(xipni+1)\bigotimesop\displaylimits_{i=1}^{n}K[x_{i}]\,/(x_{i}^{p^{n-i+1}}) and order |Un|=h0(𝒪Un)=pn(n+1)/2|U_{n}|=h^{0}(\mathscr{O}_{U_{n}})=p^{n(n+1)/2}. It contains but one point and is thus an infinitesimal group scheme.

One immediately sees that the restriction of (4.1) to αpn=αp××αp\alpha_{p}^{\oplus n}=\alpha_{p}\times\ldots\times\alpha_{p} respects the group laws and gives an inclusion of group schemes αpnUn\alpha_{p}^{\oplus n}\subset U_{n}. Furthermore, for every mnm\leq n, we have canonical inclusions UmUnU_{m}\subset U_{n} of group schemes.

Recall that each scheme XX over our ground field KK comes with a relative Frobenius map F:XX(p)F\colon X\rightarrow X^{(p)}, given in functorial terms by X(R)FX(RF)=X(p)(R)X(R)\stackrel{{\scriptstyle F}}{{\rightarrow}}X({}_{F}R)=X^{(p)}(R). Here RF{}_{F}R denotes the abelian group RR, viewed as an RR-algebra via the absolute Frobenius map ffpf\mapsto f^{p}, and X(p)=XK(KF)X^{(p)}=X\otimes_{K}({}_{F}K). Note that R=RFR={}_{F}R as an Fp\mathbb{F}_{p}-algebra. Hence X(R)=X(RF)X(R)=X({}_{F}R) and thus X=X(p)X=X^{(p)}, provided that XX arises as base-change from the prime field Fp\mathbb{F}_{p}. For our group scheme UnU_{n}, the relative Frobenius map takes the form

Un(R)Un(RF)=Un(R),1+λiFi1+λipFi.U_{n}(R)\longrightarrow U_{n}({}_{F}R)=U_{n}(R),\quad 1+\sumop\displaylimits\lambda_{i}F^{i}\longmapsto 1+\sumop\displaylimits\lambda_{i}^{p}F^{i}.
Proposition 4.1.

The image of  F:UnUnF\colon U_{n}\rightarrow U_{n} is the subgroup scheme Un1U_{n-1}, and its kernel is given by αpn\alpha_{p}^{\oplus n}. In particular, we have an identification of restricted Lie algebras Lie(Un)=Kn\operatorname{Lie}(U_{n})=K^{n}.

Proof.

Obviously, the Frobenius map factors over the subgroup scheme Un1UnU_{n-1}\subset U_{n}. The resulting map F:UnUn1F\colon U_{n}\rightarrow U_{n-1} is indeed an epimorphism because any RR-valued point 1+μiFi1+\sumop\displaylimits\mu_{i}F^{i} of Un1U_{n-1} arises from the RR^{\prime}-valued point 1+λi1+\sumop\displaylimits\lambda_{i} of UnU_{n}, for the fppf extension R=R[λi]/(λipμi)R^{\prime}=\bigotimesop\displaylimits R[\lambda_{i}]\,/(\lambda_{i}^{p}-\mu_{i}).

An RR-valued point 1+λiFi1+\sumop\displaylimits\lambda_{i}F^{i} belongs to the kernel of the Frobenius map if and only if λip=0\lambda_{i}^{p}=0, and hence Un[F]=αpnU_{n}[F]=\alpha_{p}^{\oplus n}. The last assertion follows because Lie(αpn)=Kn\operatorname{Lie}(\alpha_{p}^{\oplus n})=K^{n}, and for any group scheme, the inclusion of the Frobenius kernel induces a bijection on Lie algebras. ∎

In turn, the relative Frobenius map F:UnUnF\colon U_{n}\rightarrow U_{n} yields an extension

(4.2) 0αpnUnUn11.0\longrightarrow\alpha_{p}^{\oplus n}\longrightarrow U_{n}\longrightarrow U_{n-1}\longrightarrow 1.

By induction on n0n\geq 0, we infer that the finite group scheme UnU_{n} admits a composition series with quotients isomorphic to αp\alpha_{p}. In particular, UnU_{n} is unipotent. Since all Lie brackets are trivial, the adjoint representation ad:ungl(un)\operatorname{ad}\colon\mathfrak{u}_{n}\rightarrow\mathfrak{gl}(\mathfrak{u}_{n}) of the Lie algebra un=Lie(Un)\mathfrak{u}_{n}=\operatorname{Lie}(U_{n}) is trivial, and the adjoint representation Ad:UnGLun/k\operatorname{Ad}\colon U_{n}\rightarrow\operatorname{GL}_{\mathfrak{u}_{n}/k} of the group scheme factors over the quotient Un1U_{n-1}. It is not difficult to determine the latter representation: Since the group Un(K)U_{n}(K) is trivial, we have

Lie(Un)=Un(K[ϵ])={1+ϵr=1nαrFrαrK},\operatorname{Lie}(U_{n})=U_{n}(K[\epsilon])=\left\{1+\epsilon\sumop\displaylimits_{r=1}^{n}\alpha_{r}F^{r}\mid\text{$\alpha_{r}\in K$}\right\},

where ϵ\epsilon denotes an indeterminate subject to ϵ2=0\epsilon^{2}=0. The elements 1+ϵFr1+\epsilon F^{r}, 1rn1\leq r\leq n, form a basis of this KK-vector space. With P=λiFiP=\sumop\displaylimits\lambda_{i}F^{i}, where λ0=1\lambda_{0}=1, and using the relations ϵ2=0\epsilon^{2}=0 and Fϵ=0F\epsilon=0, we get

P1(1+ϵFs)P=1+ϵFsP=1+ϵi=0nsλipsFs+i=i=0ns(1+ϵλipsFs+i).P^{-1}\cdot(1+\epsilon F^{s})\cdot P=1+\epsilon F^{s}P=1+\epsilon\sumop\displaylimits_{i=0}^{n-s}\lambda_{i}^{p^{s}}F^{s+i}=\prodop\displaylimits_{i=0}^{n-s}\left(1+\epsilon\lambda_{i}^{p^{s}}F^{s+i}\right).

Consequently, Ad(P1)\operatorname{Ad}(P^{-1}) sends the basis vector es=1+ϵFse_{s}=1+\epsilon F^{s} to the linear combination r=snλrspser\sumop\displaylimits_{r=s}^{n}\lambda^{p^{s}}_{r-s}e_{r}. Summing up, in the restricted Lie algebra Lie(Un)=Kn\operatorname{Lie}(U_{n})=K^{n} all brackets and pp-powers are zero, and the adjoint representation of the group scheme is given by (λiFi)1(λrsps)nrs1(\sumop\displaylimits\lambda_{i}F^{i})^{-1}\mapsto(\lambda^{p^{s}}_{r-s})_{n\geq r\geq s\geq 1}.

As described in Section 3, the groups Un(R)U_{n}(R) act from the right on the polynomial ring R[x]R[x] via RR-linear maps. This is obviously functorial in RR and thus constitutes an action of the group scheme UnU_{n} on the affine line A1=SpecK[x]\mathbb{A}^{1}=\operatorname{Spec}K[x]. Note that this is indeed an action from the left. On RR-valued points, it is given by

(λ1,,λn)μ=i=0nλiFiμ=i=0nλiμpi,(\lambda_{1},\ldots,\lambda_{n})\ast\mu=\sumop\displaylimits_{i=0}^{n}\lambda_{i}F^{i}\ast\mu=\sumop\displaylimits_{i=0}^{n}\lambda_{i}\mu^{p^{i}},

where we set λ0=1\lambda_{0}=1 for convenience. Of course, we also have the canonical actions of the multiplicative group Gm\mathbb{G}_{m} and the additive group Ga\mathbb{G}_{a}, given via λ0μ=λ0μ\lambda_{0}\ast\mu=\lambda_{0}\mu and αμ=μ+α\alpha\ast\mu=\mu+\alpha, 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 AutA1/K\operatorname{Aut}_{\mathbb{A}^{1}/K}, the group scheme Ga\mathbb{G}_{a} is normalized by UnU_{n}, and both Ga\mathbb{G}_{a} and UnU_{n} are normalized by Gm\mathbb{G}_{m}. Moreover, the intersections

GaUnand(GaUn)Gm\mathbb{G}_{a}\cap U_{n}\quad\text{and}\quad(\mathbb{G}_{a}\rtimes U_{n})\cap\mathbb{G}_{m}

inside the sheaf AutA1/K\operatorname{Aut}_{\mathbb{A}^{1}/K} are trivial.

Proof.

The assertions follow from Propositions 3.1 and 2.4. ∎

We thus have an iterated semidirect product, for simplicity written as

(4.3) GaUnGm=(GaUn)Gm=Ga(UnGm),\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}=(\mathbb{G}_{a}\rtimes U_{n})\rtimes\mathbb{G}_{m}=\mathbb{G}_{a}\rtimes(U_{n}\rtimes\mathbb{G}_{m}),

acting faithfully on the affine line A1=SpecK[x]\mathbb{A}^{1}=\operatorname{Spec}K[x]. In turn, we get an inclusion of restricted Lie algebras

KLie(Un)gl1(K)Lie(AutA1/K)=DerK(K[x]).K\rtimes\operatorname{Lie}(U_{n})\rtimes\mathfrak{gl}_{1}(K)\subset\operatorname{Lie}(\operatorname{Aut}_{\mathbb{A}^{1}/K})=\operatorname{Der}_{K}(K[x]).

The elements on the left-hand side can be seen as tuples (α,λ1,,λn,λ0)(\alpha,\lambda_{1},\ldots,\lambda_{n},\lambda_{0}) and correspond to the KK-derivation αx+i=1nλixpix+λ0xx\alpha\frac{\partial}{\partial x}+\sumop\displaylimits_{i=1}^{n}\lambda_{i}x^{p^{i}}\frac{\partial}{\partial x}+\lambda_{0}x\frac{\partial}{\partial x} of the polynomial ring K[x]K[x]. For example, the derivation λ1xp/xLie(Un)\lambda_{1}x^{p}\partial/\partial x\in\operatorname{Lie}(U_{n}) acts via xx+ϵλ1xpx\mapsto x+\epsilon\lambda_{1}x^{p}, which coincides with the action of the group element 1+ϵλ1FUn(k[ϵ])1+\epsilon\lambda_{1}F\in U_{n}(k[\epsilon]).

The spectrum of the function field K(x)K(x) comes with a monomorphism

(4.4) SpecK(x)SpecK[x]=A1.\operatorname{Spec}K(x)\longrightarrow\operatorname{Spec}K[x]=\mathbb{A}^{1}.

According to Proposition 3.2, there is a unique action on SpecK(x)\operatorname{Spec}K(x) that makes the above morphism equivariant.

Let us close this section with some observations on central series. Recall that for a group GG, the lower central series =rGr{}^{r}={}^{r}G and the upper central series Zs=ZsGZ_{s}=Z_{s}G are inductively defined by

=0G,=r+1[G,]randZ0={e},Zs+1/Zs=Z(G/Zs).{}^{0}=G,\quad{}^{r+1}=[G,{}^{r}]\quad\text{and}\quad Z_{0}=\{e\},\quad Z_{s+1}/Z_{s}=Z(G/Z_{s}).

The group is nilpotent if =r{e}{}^{r}=\{e\} for some r0r\geq 0, or equivalently Zs=GZ_{s}=G for some s0s\geq 0. Then the smallest such integers coincide, and this number nn is called the nilpotency class of the group. Note that nrZr{}_{n-r}\subset Z_{r}, 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 GG of finite type, one has basically the same construction, with sheafification involved. This is straightforward for the higher centers: An xG(R)x\in G(R) belongs to Zs+1(R)Z_{s+1}(R) if and only if it commutes with all members of G(R)G(R^{\prime}) up to elements of Zs(R)Z_{s}(R^{\prime}), for all flat extensions RRR\subset R^{\prime}. The situation is more complicated for the r because their formation involves schematic images and group scheme closure with respect to the commutator maps G×rr+1G\times{}^{r}\rightarrow{}^{r+1}; see [SGA3-1, Exposé VIB\text{VI}_{B}, Section 8].

Let us unravel this for our G=UnG=U_{n}: consider the closed subschemes

(4.5) {1}=G0G1Gn=Un\{1\}=G_{0}\subset G_{1}\subset\ldots\subset G_{n}=U_{n}

defined by Gr(R)={1+i=nr+1nλiFi}G_{r}(R)=\{1+\sumop\displaylimits_{i=n-r+1}^{n}\lambda_{i}F^{i}\}.

Proposition 4.3.

The GrUnG_{r}\subset U_{n} are subgroup schemes, and the series (4.5) coincides with both the upper and the lower central series for the group scheme UnU_{n}. The quotients are Gr+1/Gr=αpr+1G_{r+1}/G_{r}=\alpha_{p^{r+1}}.

Proof.

With descending induction one easily checks that GrGr+1G_{r}\subset G_{r+1} are subgroup schemes: the surjection Gr+1αpr+1G_{r+1}\rightarrow\alpha_{p^{r+1}} given by 1+i=nrnλiFiλnr1+\sumop\displaylimits_{i=n-r}^{n}\lambda_{i}F^{i}\mapsto\lambda_{n-r} respects the group law and has kernel GrG_{r}. 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 γiFi=(βiFi)1\sumop\displaylimits\gamma_{i}F^{i}=(\sumop\displaylimits\beta_{i}F^{i})^{-1} shows that each coefficient γi=γi(β0,,βn)\gamma_{i}=\gamma_{i}(\beta_{0},\ldots,\beta_{n}) actually depends only on β0,,βi\beta_{0},\ldots,\beta_{i}. From this one easily infers

(4.6) (αiFi)(βiFi)1Gr(R)αi=βi for 0inr.\left(\sumop\displaylimits\alpha_{i}F^{i}\right)\cdot\left(\sumop\displaylimits\beta_{i}F^{i}\right)^{-1}\in G_{r}(R)\quad\Longleftrightarrow\quad\text{$\alpha_{i}=\beta_{i}$ for $0\leq i\leq n-r$}.

Write ZrZ_{r} for the higher centers of UnU_{n}. We show ZrGrZ_{r}\subset G_{r} by induction on 0rn0\leq r\leq n. The case r=0r=0 is trivial. Now suppose r1r\geq 1 and that the inclusion holds for r1r-1. For each x=λiFix=\sumop\displaylimits\lambda_{i}F^{i} from Un(R)U_{n}(R), we compute

(1μF)x=xi=0n1μλipFi+1andx(1μF)=xi=0n1λiμpiFi+1.(1-\mu F)\cdot x=x-\sumop\displaylimits_{i=0}^{n-1}\mu\lambda_{i}^{p}F^{i+1}\quad\text{and}\quad x\cdot(1-\mu F)=x-\sumop\displaylimits_{i=0}^{n-1}\lambda_{i}\mu^{p^{i}}F^{i+1}.

Suppose that xx belongs to Zr(R)Z_{r}(R). Then for all μαpn(R)\mu\in\alpha_{p^{n}}(R^{\prime}) in some ring extension RRR\subset R^{\prime}, the above two expressions coincide modulo Zr1Gr1Z_{r-1}\subset G_{r-1}. From the equivalence (4.6), we obtain μλip=λiμpi\mu\lambda_{i}^{p}=\lambda_{i}\mu^{p^{i}} for 0inr0\leq i\leq n-r. For R=R[μ]/(μpn)R^{\prime}=R[\mu]\,/(\mu^{p^{n}}), we are in position to compare coefficients and infer λ1==λnr=0\lambda_{1}=\ldots=\lambda_{n-r}=0, and thus xGr(R)x\in G_{r}(R).

This completes our induction and establishes ZrGrZ_{r}\subset G_{r} for all 0rn0\leq r\leq n. For the reverse inclusion, we use our embedding UnUTn+1U_{n}\subset\operatorname{UT}_{n+1} into the group of unitriangular matrices. According to Lemma 4.4 below, the rthr^{\mathrm{th}} higher center of UTn+1(R)\operatorname{UT}_{n+1}(R) is given by the matrices that are zero on the nrn-r secondary diagonals above the main diagonal. The intersection with Un(R)U_{n}(R) equals Gr(R)G_{r}(R). Consequently, GrZrG_{r}\subset Z_{r}; thus the Gr=ZrG_{r}=Z_{r} form the upper central series.

The arguments for the higher commutator groups rely on some preliminary observations. For elements of the form b=1βFsγFs+1+b=1-\beta F^{s}-\gamma F^{s+1}+\cdots with any s1s\geq 1, the geometric series (1x)1=1+x+x2+(1-x)^{-1}=1+x+x^{2}+\cdots gives

b1=(1βFsγFs+1+)11+βFs+γFs+1+β1+psF2s,b^{-1}=\left(1-\beta F^{s}-\gamma F^{s+1}+\cdots\right)^{-1}\equiv 1+\beta F^{s}+\gamma F^{s+1}+\beta^{1+p^{s}}F^{2s},

where the congruence means up to terms of order s+2s+2. Note that the last summand is only relevant in the special case s=1s=1. With a=1αFa=1-\alpha F, the above formula shows that the commutator aba1b1aba^{-1}b^{-1} is congruent to

(1βFsγFs+1)1+(1αF)(βFsγFs+1)(1+αF)(1+βFs)\displaystyle\left(1-\beta F^{s}-\gamma F^{s+1}\right)^{-1}+(1-\alpha F)\left(-\beta F^{s}-\gamma F^{s+1}\right)(1+\alpha F)\left(1+\beta F^{s}\right)\equiv
1+βFs+γFs+1+β1+psF2sβFs+αβpFs+1βαpsFs+1β1+psF2sγFs+1.\displaystyle 1+\beta F^{s}+\gamma F^{s+1}+\beta^{1+p^{s}}F^{2s}-\beta F^{s}+\alpha\beta^{p}F^{s+1}-\beta\alpha^{p^{s}}F^{s+1}-\beta^{1+p^{s}}F^{2s}-\gamma F^{s+1}.

Most summands cancel, and the upshot is the commutator formula

(4.7) aba1b1=1+(αβpβαps)Fs+1+.aba^{-1}b^{-1}=1+\left(\alpha\beta^{p}-\beta\alpha^{p^{s}}\right)F^{s+1}+\cdots.

Write r for the higher commutator subgroup schemes. According to general properties of nilpotent groups (cf. [KM79, p. 107]), we have rZnr=Gnr{}^{r}\subset Z_{n-r}=G_{n-r}. We claim that the canonical projection

(4.8) =r[Un,]r1Gnr/Gnr1=αpnr{}^{r}=\left[U_{n},{}^{r-1}\right]\longrightarrow G_{n-r}/G_{n-r-1}=\alpha_{p^{n-r}}

is an epimorphism. We check this by induction on r0r\geq 0. The case r=0r=0 is trivial. Suppose r1r\geq 1 and that the assertion is true for r1r-1. According to [DG70, Section IV.2, Proposition 1.1], the iterated Frobenius kernels are the only subgroup schemes of αpnrGa\alpha_{p^{n-r}}\subset\mathbb{G}_{a}. Seeking a contradiction, we assume that the above map factors over αpnr1\alpha_{p^{n-r-1}}. Consider the ring R=K[α,β]/(αpn,βpnr+1)R=K[\alpha,\beta]\,/(\alpha^{p^{n}},\beta^{p^{n-r+1}}). By our induction hypothesis, there are some faithfully flat extension RRR\subset R^{\prime} and some RR^{\prime}-valued point of the form b=1βFr1γFr+b=1-\beta F^{r-1}-\gamma F^{r}+\cdots from r-1. With a=1αFa=1-\alpha F, the commutator formula (4.7) shows that aba1b1aba^{-1}b^{-1} projects to λ=αβpβαpr\lambda=\alpha\beta^{p}-\beta\alpha^{p^{r}} under (4.8). So λpnr1=αpnr1βpnrβpnr1αpn1\lambda^{p^{n-r-1}}=\alpha^{p^{n-r-1}}\beta^{p^{n-r}}-\beta^{p^{n-r-1}}\alpha^{p^{n-1}} vanishes in the ring RR^{\prime}. On the other hand, both of the appearing monomials belong to the monomial basis for RR, giving a contradiction. Thus (4.8) is an epimorphism.

We are now ready to prove that the inclusion sGns{}^{s}\subset G_{n-s} is an equality. Fix some xGns(R)x\in G_{n-s}(R), and write it as x=1+i=s+1nλiFix=1+\sumop\displaylimits_{i=s+1}^{n}\lambda_{i}F^{i}. We check that x(R)sx\in{}^{s}(R) by descending induction on sns\leq n. The case s=ns=n is trivial. Now assume s<ns<n and that the assertion holds for s+1s+1. By the preceding paragraph, there are some faithfully flat extension RRR\subset R^{\prime} and some RR^{\prime}-valued point y=1+i=s+1nμiFiy=1+\sumop\displaylimits_{i=s+1}^{n}\mu_{i}F^{i} from s with μs+1=λs+1\mu_{s+1}=-\lambda_{s+1}. Then xy=1+i=s+2λiFixy=1+\sumop\displaylimits_{i=s+2}\lambda^{\prime}_{i}F^{i} belongs to Gns1(R)G_{n-s-1}(R^{\prime}). Using our induction hypothesis, together with the inclusion s+1s{}^{s+1}\subset{}^{s}, we see that xyxy, and hence xx, belongs to (R)s{}^{s}(R^{\prime}), and by descent x(R)sx\in{}^{s}(R). ∎

Let us point out that the ring K[α,β]/(αpn,βpnr+1)K[\alpha,\beta]\,/(\alpha^{p^{n}},\beta^{p^{n-r+1}}) is not free as a module over K[λ]K[\lambda], which one sees by analyzing the size of the Jordan blocks for multiplication by λ=αβpβαpr\lambda=\alpha\beta^{p}-\beta\alpha^{p^{r}}. Thus it is not always possible to factor a given element of (R)r+1{}^{r+1}(R) into commutators, even over flat extensions RRR\subset R^{\prime}.

In the preceding proof, we have used the following fact.

Lemma 4.4.

The unitriangular matrix group UTn+1(R)\operatorname{UT}_{n+1}(R), over any ring RR, has upper central series given by

(4.9) Zs={E+(ζij)ζij=0 whenever jins}.Z_{s}=\left\{E+\left(\zeta_{ij}\right)\mid\text{$\zeta_{ij}=0$ whenever $j-i\leq n-s$}\right\}.
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 s0s\geq 0. The case s=0s=0 is trivial. Now suppose s1s\geq 1 and that the assertion is true for s1s-1. By definition, a unitriangular E+(αij)E+(\alpha_{ij}) belongs to ZsZ_{s} if and only if

(4.10) (E+(αij))(E+(βij))(E+(βij))(E+(αij))modulo Zs1\left(E+\left(\alpha_{ij}\right)\right)\cdot\left(E+\left(\beta_{ij}\right)\right)\equiv\left(E+\left(\beta_{ij}\right)\right)\cdot\left(E+\left(\alpha_{ij}\right)\right)\quad\text{modulo $Z_{s-1}$}

for every unitriangular matrix E+(βij)E+(\beta_{ij}). By the induction hypothesis, each E+(ζij)Zs1E+(\zeta_{ij})\in Z_{s-1} has ζij=0\zeta_{ij}=0 for jin+1sj-i\leq n+1-s, and one easily checks that right multiplication with elements of Zs1Z_{s-1} to a unitriangular matrix leaves the (i,k)(i,k)-entries unchanged for kin+1sk-i\leq n+1-s. From this the inclusion \supset of (4.9) easily follows. Conversely, suppose that we have some E+(αij)ZsE+(\alpha_{ij})\in Z_{s}, so (4.10) holds. This means

jαijβjk=jβijαjkwhenever kin+1s,\sumop\displaylimits_{j}\alpha_{ij}\beta_{jk}=\sumop\displaylimits_{j}\beta_{ij}\alpha_{jk}\quad\text{whenever $k-i\leq n+1-s$,}

where E+(βij)E+(\beta_{ij}) is any unitriangular matrix and the sums actually run over i<j<ki<j<k. From this one easily infers by induction on r=jir=j-i that αij\alpha_{ij} vanishes for 1rns1\leq r\leq n-s. ∎

5.  Compactifications and numerical semigroups

We keep the setting of the previous section but now work with a new indeterminate T=x1T=x^{-1}. The iterated semidirect product GaUnGm\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m} has as coordinate ring

(𝒪GaUnGm)=K[α,λ1,,λn,λ0±]/(λ1pn,λ2pn1,,λn1p2,λnp),\Gamma\left(\mathscr{O}_{\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}}\right)=K\left[\alpha,\lambda_{1},\ldots,\lambda_{n},\lambda_{0}^{\pm}\right]/\left(\lambda_{1}^{p^{n}},\lambda_{2}^{p^{n-1}},\ldots,\lambda_{n-1}^{p^{2}},\lambda_{n}^{p}\right),

endowed with a Hopf algebra structure, and acts on the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}]. We now seek to extend this action to certain compactifications, all of which are denormalizations of the projective line P1=SpecK[T]SpecK[T1]\mathbb{P}^{1}=\operatorname{Spec}K[T]\cup\operatorname{Spec}K[T^{-1}]. For this, we have to make extensive computations in the first chart, which are much easier to carry out with TT rather than x1x^{-1}. Note that by Proposition 3.2 we have an induced action on the spectrum of the function field K(T)=K(x)K(T)=K(x), and this action takes the form

(5.1) K(T)(𝒪GaUnGm)K(T),T(α+i=0nλiTpi)1.K(T)\longrightarrow\Gamma\left(\mathscr{O}_{\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}}\right)\otimes K(T),\quad T\longmapsto\left(\alpha+\sumop\displaylimits_{i=0}^{n}\lambda_{i}T^{-p^{i}}\right)^{-1}.

Recall that an additive submonoid N\Gamma\subset\mathbb{N} whose complement is finite is called a numerical semigroup. Equivalently, the induced inclusions of groups grpNgrp=Z{}^{\operatorname{grp}}\subset\mathbb{N}^{\operatorname{grp}}=\mathbb{Z} is an equality, or gcd(a1,,ar)=1\gcd(a_{1},\ldots,a_{r})=1 for some members a1,,ara_{1},\ldots,a_{r}\in\Gamma. Each numerical semigroup comes with the following invariants: The multiplicity e1e\geq 1 is the smallest non-zero element in . The conductor is the smallest integer c0c\geq 0 with {c,c+1,}\{c,c+1,\ldots\}\subset\Gamma. The genus g0g\geq 0 is the cardinality of the complement N\Gamma\smallsetminus\mathbb{N}, 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 d1d\geq 1. For general overviews, we refer to the textbooks [RGS09] and [AGS16].

For each numerical semigroup , the ring K[T]=K[Taa]K[T]=K[T^{a}\mid a\in\Gamma] defines a compactification

X=SpecK[T]SpecK[T1]X=\operatorname{Spec}K\left[T\right]\cup\operatorname{Spec}K\left[T^{-1}\right]

of the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}], obtained by adding a single rational point x0Xx_{0}\in X. The gluing of the two affine open sets is given by the common localization K[T±1]K[T^{\pm 1}] of the coordinate rings. The normalization is P1=SpecK[T]SpecK[T1]\mathbb{P}^{1}=\operatorname{Spec}K[T]\cup\operatorname{Spec}K[T^{-1}], and the ensuing map f:P1Xf\colon\mathbb{P}^{1}\rightarrow X is described by the conductor square

(5.2) AP1fBX,\begin{CD}A@>{}>{}>\mathbb{P}^{1}\\ @V{}V{}V@V{}V{f}V\\ B@>{}>{}>X,\end{CD}

which is both cartesian and cocartesian (for details see [FS20, Appendix A]). The conductor loci AP1A\subset\mathbb{P}^{1} and BXB\subset X are the closed subschemes whose respective coordinate rings are K[T]/(Tc)K[T]\,/(T^{c}) and K[T]/(Tc,Tc+1,)K[T]\,/(T^{c},T^{c+1},\ldots). Consider the short exact sequence 0𝒪Xf(𝒪P1)×𝒪Bf(𝒪A)00\rightarrow\mathscr{O}_{X}\rightarrow f_{*}(\mathscr{O}_{\mathbb{P}^{1}})\times\mathscr{O}_{B}\rightarrow f_{*}(\mathscr{O}_{A})\rightarrow 0 of sheaves on XX, where the inclusion is the diagonal map and the surjection is the difference map. It yields

(5.3) h0(𝒪X)=1,h1(𝒪X)=gande(𝒪X,x0)=e,edim(𝒪X,x0)=d,h^{0}\left(\mathscr{O}_{X}\right)=1,\quad h^{1}\left(\mathscr{O}_{X}\right)=g\quad\text{and}\quad e\left(\mathscr{O}_{X,x_{0}}\right)=e,\quad\operatorname{edim}\left(\mathscr{O}_{X,x_{0}}\right)=d,

with the invariants c,g,e,dc,g,e,d of the numerical semigroup discussed above. Here e(𝒪X,x0)e(\mathscr{O}_{X,x_{0}}) and edim(𝒪X,x0)\operatorname{edim}(\mathscr{O}_{X,x_{0}}) denote the multiplicity and the embedding dimension of the local ring, respectively.

Given a subgroup scheme GGaUnGmG\subset\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}, it is natural to ask whether the resulting GG-action on the affine line A1\mathbb{A}^{1} extends to the compactification XX. If it exists, such an extension is unique because the open set A1R\mathbb{A}^{1}\otimes R is schematically dense in XRX\otimes R for any ring RR.

In the following assertion on the constituents of the iterated semidirect product, we regard the expression P=(1+i=1nλiT1pi)dP=(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}})^{-d} as a Laurent polynomial in the indeterminate TT with coefficients from Fp[λ1,,λn]/(λ1pn,λ2pn1,,λnp)\mathbb{F}_{p}[\lambda_{1},\ldots,\lambda_{n}]\,/(\lambda_{1}^{p^{n}},\lambda_{2}^{p^{n-1}},\ldots,\lambda_{n}^{p}), and Q=(1+αT)dQ=(1+\alpha T)^{-d} as a formal power series in TT with coefficients from Fp[α]\mathbb{F}_{p}[\alpha]. In both cases we use the ensuing notion of supports Supp(P)\operatorname{Supp}(P) and Supp(Q)\operatorname{Supp}(Q) inside the group of exponents Z\mathbb{Z}.

Proposition 5.1.

We keep the notation as above. Then the following hold:

  1. (i)

    The multiplicative group G=GmG=\mathbb{G}_{m} always admits an extension.

  2. (ii)

    For the infinitesimal group scheme G=UnG=U_{n}, the extension exists if and only if for each dd\in\Gamma and sSupp(P)s\in\operatorname{Supp}(P) for the Laurent polynomial P=(1+i=1nλiT1pi)dP=(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}})^{-d}, we also have d+sd+s\in\Gamma.

  3. (iii)

    For the additive group G=GaG=\mathbb{G}_{a}, the extension exists if and only if for each dd\in\Gamma and sSupp(Q)s\in\operatorname{Supp}(Q) for the formal power series Q=(1+αT)dQ=(1+\alpha T)^{-d}, we have d+sd+s\in\Gamma.

Moreover, it suffices to verify these conditions for a set of generators dd\in\Gamma.

Proof.

(i)  Recall that Gm\mathbb{G}_{m}-actions on affine schemes correspond to Z\mathbb{Z}-gradings, according to [SGA3-1, Exposé I, Corollary 4.7.3.1]. The action on SpecK[T1]\operatorname{Spec}K[T^{-1}] is given by deg(Ti)=i\deg(T^{-i})=-i. This also defines compatible gradings on K[T]K[T], which yields the desired extension of the action of G=GmG=\mathbb{G}_{m}.

(ii)  The group scheme G=UnG=U_{n} is infinitesimal; hence every open set on a GG-scheme is GG-stable. It follows that the GG-action extends if and only if the map K[T](𝒪G)K(T)K[T]\rightarrow\Gamma(\mathscr{O}_{G})\otimes K(T) induced from (5.1) factors over the subring (𝒪G)K[T]\Gamma(\mathscr{O}_{G})\otimes K[T]. This map sends T1T^{-1} to T1+λiTpi=T1(1+λiT1pi)T^{-1}+\sumop\displaylimits\lambda_{i}T^{-p^{i}}=T^{-1}(1+\sumop\displaylimits\lambda_{i}T^{1-p^{i}}). Note that the second factor is invertible because its second summand is nilpotent. The monomial TdT^{d} with dd\in\Gamma is mapped to TdP(T)T^{d}P(T). This belongs to the subring R[T]R[T] if and only if for each sSupp(P)s\in\operatorname{Supp}(P), the resulting integer d+sd+s belongs to the numerical semigroup .

(iii)  The action of G=GaG=\mathbb{G}_{a} on A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}] extends to the projective line P1=ProjK[U0,U1]\mathbb{P}^{1}=\operatorname{Proj}K[U_{0},U_{1}] via the assignments U1U1U_{1}\mapsto U_{1} and U0U0+αU1U_{0}\mapsto U_{0}+\alpha U_{1}, with T=U1/U0T=U_{1}/U_{0}. Note that the origin 0P10\in\mathbb{P}^{1} is fixed but does not admit a stable affine open neighborhood. However, the infinitesimal neighborhoods and in particular the conductor locus AP1A\subset\mathbb{P}^{1} are stable.

Since GG is smooth, any GG-action on A1\mathbb{A}^{1} uniquely extends to P1\mathbb{P}^{1}, according to [Bri22b, Theorem 2]. By [Lau19, Lemma 3.5] the GG-action on the projective line descends to an action on XX if and only if the action on the conductor locus AA descends to an action on BB. The latter simply means that the map

(5.4) (B,𝒪B)(A,𝒪A)K[α](A,𝒪A)\Gamma(B,\mathscr{O}_{B})\longrightarrow\Gamma(A,\mathscr{O}_{A})\longrightarrow K[\alpha]\otimes\Gamma(A,\mathscr{O}_{A})

factors over K[α](B,𝒪B)K[\alpha]\otimes\Gamma(B,\mathscr{O}_{B}). Here the map on the right describes the GG-action on AA, and the coordinate ring on the left is (B,𝒪B)=K[T]/(Tc,Tc+1,)\Gamma(B,\mathscr{O}_{B})=K[T]\,/(T^{c},T^{c+1},\ldots), where c0c\geq 0 is the conductor of the numerical semigroup. As KK-vector space, this is generated by the residue classes of TdT^{d}, dd\in\Gamma. The map (5.1) sends T1T^{-1} to T1+α=T1(1+αT)T^{-1}+\alpha=T^{-1}(1+\alpha T), so the monomial TdT^{d} is mapped to TdQ(T)T^{d}Q(T). The class of the latter belongs to K[T]/(Tc,Tc+1,)K[T]\,/(T^{c},T^{c+1},\ldots) if and only if for all sSupp(Q)s\in\operatorname{Supp}(Q), we have d+sd+s\in\Gamma. ∎

Note that in the expansions of P(T)P(T) and Q(T)Q(T), some multinomial coefficients appear, and the above conditions involve their congruence properties modulo the prime number pp. Also note that one may view XX as a non-normal torus embedding with respect to the one-dimensional torus Gm=SpecK[T±1]\mathbb{G}_{m}=\operatorname{Spec}K[T^{\pm 1}].

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 G=GaUnGmG=\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}, the action on A1\mathbb{A}^{1} extends to XX. Then the whole GG-action extends to XX.

Proof.

This is a general fact: all relations between the RR-valued points of the constituents stemming from the semidirect product structures hold on A1\mathbb{A}^{1} and thus also on XX because the former is schematically dense in the latter. ∎

We now introduce a particular that is generated by n+1n+1 numbers.

Definition 5.3.

We write p,nN{}_{p,n}\subset\mathbb{N} for the numerical semigroup generated by

(5.5) pnandpnpj(0jn1).p^{n}\quad\text{and}\quad p^{n}-p^{j}\quad(0\leq j\leq n-1).

This is indeed a numerical semigroup because gcd(pn,pnp0)=1\gcd(p^{n},p^{n}-p^{0})=1. Its multiplicity is given by

ep,n={pn1(p1)if pn3,1elsee_{p,n}=\begin{cases}p^{n-1}(p-1)&\text{if $p^{n}\geq 3$},\\ 1&\text{else}\\ \end{cases}

because in the first case, the number pn1(p1)p^{n-1}(p-1) is smallest among the generators. Note that ep,n=1e_{p,n}=1 is equivalent to pn2p^{n}\leq 2, whereas ep,n=2e_{p,n}=2 means 3pn43\leq p^{n}\leq 4.

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

Xp,n=SpecK[Tp,n]SpecK[T1]X_{p,n}=\operatorname{Spec}K\left[T^{{}_{p,n}}\right]\cup\operatorname{Spec}K\left[T^{-1}\right]

lead to the desired generalizations of the quasielliptic curves. Indeed, in the special cases 3pn43\leq p^{n}\leq 4, we get =p,n2,3{}_{p,n}=\langle 2,3\rangle, and the ensuing coordinate rings become K[T2,T3]K[T^{2},T^{3}]. We now verify that the action of the iterated semidirect product extends to this compactification.

Theorem 5.4.

The action of the group scheme GaUp,nGm\mathbb{G}_{a}\rtimes U_{p,n}\rtimes\mathbb{G}_{m} on the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}] extends to the compactification X=Xp,nX=X_{p,n}.

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 G=GmG=\mathbb{G}_{m} is immediate. Now suppose G=UnG=U_{n}, and fix one of the generators dp,nd\in{}_{p,n} listed in (5.5). We have to understand the expression

P=(1+i=1nλiT1pi)d.P=\left(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}}\right)^{-d}.

In the case d=pnd=p^{n}, the above simplifies to P=11=1P=1^{-1}=1, by the multinomial theorem and λipn=0\lambda_{i}^{p^{n}}=0. Thus Supp(P)={0}\operatorname{Supp}(P)=\{0\}, and obviously d+0p,nd+0\in{}_{p,n}. In the case d=pnpjd=p^{n}-p^{j} with 0jn10\leq j\leq n-1, we get

P=(1+i=1nλiT1pi)pn(1+i=1nλiT1pi)pj=1+i=1nλipjTpjpi+j.P=\left(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}}\right)^{-p^{n}}\left(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}}\right)^{p^{j}}=1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}^{p^{j}}T^{p^{j}-p^{i+j}}.

Its support equals the set {0}{pjpi+j1inj}\{0\}\cup\{p^{j}-p^{i+j}\mid 1\leq i\leq n-j\}, in light of the defining relations λipni+1=0\lambda_{i}^{p^{n-i+1}}=0. Obviously, d+0=pnpjd+0=p^{n}-p^{j} and d+(pjpi+j)=pnpi+jd+(p^{j}-p^{i+j})=p^{n}-p^{i+j} belong to p,n. Thus the action of G=UnG=U_{n} extends.

It remains to treat the case G=GaG=\mathbb{G}_{a}. Again we fix one of the generators dp,nd\in{}_{p,n} and now have to examine the formal power series Q=(1+αT)dQ=(1+\alpha T)^{-d} with coefficients from the polynomial ring Fp[α]\mathbb{F}_{p}[\alpha]. For d=pnpjd=p^{n}-p^{j}, this becomes

Q=(1+αT)pj/(1+αT)pn=(1+αpjTpj)i=0(αT)ipn.Q=(1+\alpha T)^{p^{j}}/(1+\alpha T)^{p^{n}}=\left(1+\alpha^{p^{j}}T^{p^{j}}\right)\sumop\displaylimits_{i=0}^{\infty}(-\alpha T)^{ip^{n}}.

The support is contained in {ipni0}{pj+ipni0}\{ip^{n}\mid i\geq 0\}\cup\{p^{j}+ip^{n}\mid i\geq 0\}. Clearly, d+ipn=(pnpj)+ipnd+ip^{n}=(p^{n}-p^{j})+ip^{n} and d+(pj+ipn)=(i+1)pnd+(p^{j}+ip^{n})=(i+1)p^{n} belong to p,n. The argument for d=pnd=p^{n} is likewise, and even simpler. Thus the action of G=GaG=\mathbb{G}_{a} extends. ∎

Set =p,n\Gamma={}_{p,n} and X=Xp,nX=X_{p,n}. With respect to the infinitesimal group scheme UnU_{n}, all open sets in XX are stable, and the action on the affine open set SpecK[T]\operatorname{Spec}K[T] is given by the ring homomorphism

K[T](𝒪Un)K[T],TdTd(1+i=1nλiTpi)dK[T]\longrightarrow\Gamma\left(\mathscr{O}_{U_{n}}\right)\otimes K\left[T\right],\quad T^{d}\longmapsto T^{d}\left(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{-p^{i}}\right)^{-d}

with exponents dd\in\Gamma. The orbit map x0:UnXx_{0}\colon U_{n}\rightarrow X corresponding to the rational point x0Xx_{0}\in X is given by the homomorphism φ:K[T](𝒪Un)\varphi\colon K[T]\rightarrow\Gamma(\mathscr{O}_{U_{n}}) that is implicitly described by

φ(Td)=Td(1+i=1nλiT1pi)d|T=0.\varphi\left(T^{d}\right)=T^{d}\left(1+\sumop\displaylimits_{i=1}^{n}\lambda_{i}T^{1-p^{i}}\right)^{-d}\bigg{|}_{T=0}.

Note that one has to determine the product before substituting T=0T=0 because the second factor usually contains terms of negative degree. The computation for the generators (5.5) of our numerical semigroup is immediate: φ(Tpn)=0\varphi(T^{p^{n}})=0 and φ(Tpnpj)=λnjpj\varphi(T^{p^{n}-p^{j}})=\lambda_{n-j}^{p^{j}} for 0jn10\leq j\leq n-1. Now recall that the inertia group scheme in UnU_{n} is defined by the largest quotient of (𝒪Un)\Gamma(\mathscr{O}_{U_{n}}) in which φ\varphi becomes the zero map. Setting i=nji=n-j, we get the following.

Proposition 5.5.

Inside Un=SpecK[λ1,,λn]/(λ1pn,λ2pn1,,λnp)U_{n}=\operatorname{Spec}K[\lambda_{1},\ldots,\lambda_{n}]\,/(\lambda_{1}^{p^{n}},\lambda_{2}^{p^{n-1}},\ldots,\lambda_{n}^{p}), the inertia group scheme with respect to the rational point x0Xx_{0}\in X is defined by the equations λipni=0\lambda_{i}^{p^{n-i}}=0 for 1in1\leq i\leq n.

This inertia group scheme coincides with the canonical inclusion of Un1UnU_{n-1}\subset U_{n}, which is also the image of the relative Frobenius map, and we thus obtain a UnU_{n}-stable closed subscheme Un/Un1XU_{n}/U_{n-1}\subset X. A priori, this is an effective Weil divisor supported by x0x_{0}, of degree [Un:Un1]=h0(𝒪Un)/h0(𝒪Un1)=pn[U_{n}:U_{n-1}]=h^{0}(\mathscr{O}_{U_{n}})/h^{0}(\mathscr{O}_{U_{n-1}})=p^{n}. The following observation will be crucial in what follows.

Proposition 5.6.

The Weil divisor Un/Un1XU_{n}/U_{n-1}\subset X is an effective Cartier divisor.

Proof.

The closed subscheme lies in the affine open set SpecK[T]\operatorname{Spec}K[T] and corresponds to the ideal a=Ker(φ)\mathfrak{a}=\operatorname{Ker}(\varphi). This ideal contains the monomial TpnT^{p^{n}}, and we claim that the inclusion (Tpn)a(T^{p^{n}})\subset\mathfrak{a} is an equality. In other words, we have to verify that the resulting map

φ:K[T]/(Tpn)(𝒪G)=K[λ1,,λn]/(λ1pn,λ2pn1,,λnp)\varphi\colon K\left[T\right]/\left(T^{p^{n}}\right)\longrightarrow\Gamma(\mathscr{O}_{G})=K[\lambda_{1},\ldots,\lambda_{n}]\,/\left(\lambda_{1}^{p^{n}},\lambda_{2}^{p^{n-1}},\ldots,\lambda_{n}^{p}\right)

is injective. We computed above that its image is the subring generated by the powers λipni\lambda_{i}^{p^{n-i}} for 1in1\leq i\leq n, which is a KK-algebra of degree pnp^{n}. So it suffices to verify that the KK-algebra K[T]/(Tpn)K[T]\,/(T^{p^{n}}) has degree at most pnp^{n}. This algebra is generated by the classes xjx_{j} of TpnpjT^{p^{n}-p^{j}} with 0jn10\leq j\leq n-1. From the relation

p(pnpj)=(p1)pn+(pnpj+1)p(p^{n}-p^{j})=(p-1)p^{n}+\left(p^{n}-p^{j+1}\right)

in the numerical semigroup , we infer a factorization (Tpnpj)p=(Tpn)p1Tpnpj+1(T^{p^{n}-p^{j}})^{p}=(T^{p^{n}})^{p-1}\cdot T^{p^{n}-p^{j+1}} in the ring K[T]K[T], and hence xjp=0x_{j}^{p}=0. Thus K[T]/(Tpn)K[T]\,/(T^{p^{n}}) has degree at most pnp^{n}. ∎

6.  The complete intersection property

We keep the notation as in the preceding section and continue to study the algebra of the numerical semigroup =p,n\Gamma={}_{p,n} and also the geometry of the compactification X=Xp,nX=X_{p,n} of the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}] defined by the coordinate ring K[T]K[T].

Recall that any numerical semigroup given by a set of d1d\geq 1 generators a1,,ada_{1},\ldots,a_{d} and ensuing surjection Nd\mathbb{N}^{d}\rightarrow\Gamma is called a complete intersection if the congruence R=Nd×NdR=\mathbb{N}^{d}\times\mathbb{N}^{d} is generated by d1d-1 elements. According to [Her70, Corollary 1.13], this is equivalent to the condition that the complete local ring A=K[[T]]A=K[[T]] is a complete intersection in the sense of commutative algebra; in other words, AK[[u1,,ur]]/(f1,,fs)A\simeq K[[u_{1},\ldots,u_{r}]]\,/(f_{1},\ldots,f_{s}) for some r0r\geq 0 and some regular sequence f1,,fsf_{1},\dots,f_{s}, here necessarily with s=r1s=r-1.

Proposition 6.1.

Our numerical semigroup p,n is a complete intersection, and its conductor cp,nc_{p,n} and genus gp,ng_{p,n} are given by the formulas

cp,n=npn+1(n+2)pn+2andgp,n=12cp,n.c_{p,n}=np^{n+1}-(n+2)p^{n}+2\quad\text{and}\quad g_{p,n}=\tfrac{1}{2}c_{p,n}.

Moreover, G={pnpn1,pnpn2,,pn1,pn}G=\{p^{n}-p^{n-1},p^{n}-p^{n-2},\ldots,p^{n}-1,p^{n}\} is the smallest generating set provided p3p\geq 3; for the prime p=2p=2 and n1n\geq 1, one has to omit pnp^{n}.

Proof.

First note that for p=2p=2 and n1n\geq 1, the relation pn=2(pnpn1)p^{n}=2(p^{n}-p^{n-1}) shows that the generator pnp^{n} does not belong to the smallest generating set.

We now proceed, for general p>0p>0, by induction on n0n\geq 0. For n=0n=0, we have =N\Gamma=\mathbb{N}, and all assertions are obvious. Now suppose n1n\geq 1 and that the assertion holds for n1n-1. Consider the sets of numbers

G1={pn1pn2,pn1pn3,,pn11,pn1}andG2={1}.G_{1}=\{p^{n-1}-p^{n-2},p^{n-1}-p^{n-3},\ldots,p^{n-1}-1,p^{n-1}\}\quad\text{and}\quad G_{2}=\{1\}.

Both generate respective numerical semigroups 1 and 2, and the induction hypothesis applies to the former. The numbers a1=pa_{1}=p and a2=pn1a_{2}=p^{n}-1 are relatively prime, with a12a_{1}\in{}_{2} and a2=p(pn1pn2)+(pn11)1a_{2}=p(p^{n-1}-p^{n-2})+(p^{n-1}-1)\in{}_{1}. Furthermore, =a1+1a22\Gamma=a_{1}{}_{1}+a_{2}{}_{2}. According to [Del76, Proposition 10], the monoid is a complete intersection, and the conductor is given by the formula

(6.1) c=a1c1+a2c2+(a11)(a21)=pc1+(p1)(pn2).c=a_{1}c_{1}+a_{2}c_{2}+(a_{1}-1)(a_{2}-1)=pc_{1}+(p-1)(p^{n}-2).

Here c2=0c_{2}=0 is the conductor of 2, and c1=(n1)pn(n+1)pn1+2c_{1}=(n-1)p^{n}-(n+1)p^{n-1}+2 is the conductor of 1, which we know by our induction hypothesis. Inserting the latter into (6.1), we get the desired formula for cp,nc_{p,n}. 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 gp,ng_{p,n} follows.

Now suppose p3p\geq 3. By induction, the GiiG_{i}\subset{}_{i} are the smallest generating sets. The number a1a2=pn+1pa_{1}a_{2}=p^{n+1}-p does not belong to a1G1a2G2=Ga_{1}G_{1}\cup a_{2}G_{2}=G. As explained in [Del76, proof of Proposition 10(ii)], the subset GG\subset\Gamma is the smallest generating set. For p=2p=2, one argues likewise, with pnp^{n} omitted. ∎

We see that the embedding dimension for the numerical semigroup p,n and the local ring 𝒪X,x0\mathscr{O}_{X,x_{0}} is given by the formula

dp,n={n+1if p3 or n=0,nif p=2 and n1.d_{p,n}=\begin{cases}n+1&\text{if $p\geq 3$ or $n=0$},\\ n&\text{if $p=2$ and $n\geq 1$}.\end{cases}

Let us also record the following geometric consequences.

Corollary 6.2.

The curve X=Xn,pX=X_{n,p} has invariants

h0(𝒪X)=1andh1(𝒪X)=12(npn+1(n+2)pn+2).h^{0}(\mathscr{O}_{X})=1\quad\text{and}\quad h^{1}(\mathscr{O}_{X})=\tfrac{1}{2}(np^{n+1}-(n+2)p^{n}+2).

Moreover, the dualizing sheaf ωX\omega_{X} is invertible, of degree pn(npn2)p^{n}(np-n-2).

Proof.

The values for hi(𝒪X)h^{i}(\mathscr{O}_{X}) follow with (5.3) from Proposition 6.1. Being locally of complete intersection, XX must be Gorenstein, and the dualizing sheaf is invertible. Serre duality gives deg(ωX)=2χ(𝒪X)=pn(npn2)\deg(\omega_{X})=-2\chi(\mathscr{O}_{X})=p^{n}(np-n-2). ∎

We actually can derive an explicit description for the ring K[T]K[T] in terms of generators and relations. Write aj=pnpja_{j}=p^{n}-p^{j} and b=pnb=p^{n} for the generators of =p,n\Gamma={}_{p,n}. They give rise to a surjection Nn+1\mathbb{N}^{n+1}\rightarrow\Gamma of monoids and an ensuing congruence R=Nn+1×Nn+1R=\mathbb{N}^{n+1}\times\mathbb{N}^{n+1}. The n+1n+1 generators satisfy the nn obvious relations

(6.2) pan1=(p1)bandpaj=pan1+aj+1(0jn2),p\cdot a_{n-1}=(p-1)\cdot b\quad\text{and}\quad p\cdot a_{j}=p\cdot a_{n-1}+a_{j+1}\quad(0\leq j\leq n-2),

which may be interpreted as members of the congruence RR. To translate this into commutative algebra, let xj,yx_{j},y be indeterminates corresponding to the generators aj,ba_{j},b\in\Gamma, and consider the surjection

φ:K[x1,,xn1,y]K[T]\varphi\colon K[x_{1},\ldots,x_{n-1},y]\longrightarrow K[T]

given by φ(xj)=Taj\varphi(x_{j})=T^{a_{j}} and φ(y)=Tb\varphi(y)=T^{b}. The map respects the gradings specified by deg(xj)=aj\deg(x_{j})=a_{j}, deg(y)=b\deg(y)=b, and deg(T)=1\deg(T)=1.

Proposition 6.3.

The ideal a=Ker(φ)\mathfrak{a}=\operatorname{Ker}(\varphi) is generated by the polynomials xn1pyp1x_{n-1}^{p}-y^{p-1} and xjpxn1pxj+1x_{j}^{p}-x_{n-1}^{p}x_{j+1} for 0jn20\leq j\leq n-2, 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 n+1n+1 elements a0,,an1,ba_{0},\ldots,a_{n-1},b\in\Gamma. Delorme’s observation hinges on two descending sequences

Pn+1,Pn,,P1andZn+1,Zn,,Z2.P_{n+1},P_{n},\ldots,P_{1}\quad\text{and}\quad Z_{n+1},Z_{n},\ldots,Z_{2}.

The first sequence comprises partitions PiP_{i} of the generating set G={a0,,an1,b}G=\{a_{0},\ldots,a_{n-1},b\}, subject to the following condition: Pn+1P_{n+1} is the partition into singletons, and each Pi1P_{i-1} is obtained from its precursor PiP_{i} by replacing certain members Li,LiPiL_{i},L^{\prime}_{i}\in P_{i} by their union. The second sequence consists of homogeneous polynomials ZiZ_{i} in the indeterminates x0,,xn1,yx_{0},\ldots,x_{n-1},y, taking the form Z=HiHiZ=H_{i}-H^{\prime}_{i} for some monic monomials HiH_{i} and HiH^{\prime}_{i}, each involving only indeterminates indexed by LiL_{i} and LiL^{\prime}_{i}, respectively. In loc. cit. the sequences are denoted by 𝒫\mathscr{P} and 𝒵\mathscr{Z}, and the pair (𝒫,𝒵)(\mathscr{P},\mathscr{Z}) is called a suite distinguée.

Note that the partitions PiP_{i} are fully determined by the sets Li,LiGL_{i},L^{\prime}_{i}\subset G with 2in+12\leq i\leq n+1. We now define such a partition sequence by setting

Li={ai1,,an1,b}andLi={ai2}.L_{i}=\{a_{i-1},\ldots,a_{n-1},b\}\quad\text{and}\quad L^{\prime}_{i}=\{a_{i-2}\}.

Note that this starts with the singletons Ln+1={b}L_{n+1}=\{b\} and Ln+1={an1}L^{\prime}_{n+1}=\{a_{n-1}\}. The homogeneous polynomials are declared as

Zn+1=yp1xn1pandZi=xi2pxn1pxi1(2in).Z_{n+1}=y^{p-1}-x_{n-1}^{p}\quad\text{and}\quad Z_{i}=x_{i-2}^{p}-x_{n-1}^{p}x_{i-1}\quad(2\leq i\leq n).

These have deg(Zi)=pn+1pi1\deg(Z_{i})=p^{n+1}-p^{i-1} for all 2in+12\leq i\leq n+1. One sees

gcd(Li)=gcd(pi1,,pn1,pn)=pi1andgcd(Li)=pnpi2.\gcd(L_{i})=\gcd\left(p^{i-1},\ldots,p^{n-1},p^{n}\right)=p^{i-1}\quad\text{and}\quad\gcd\left(L^{\prime}_{i}\right)=p^{n}-p^{i-2}.

The least common multiple of the above two gcds is given by p(pnpi2)p(p^{n}-p^{i-2}), which coincides with deg(Zi)\deg(Z_{i}). Our assertion now follows from [Del76, Lemma 8]. ∎

This has important consequences for Kähler differentials.

Corollary 6.4.

The sheaf /X/K1Torsion{}^{1}_{X/K}/\operatorname{Torsion} is invertible of degree pn-p^{n}, and the tangent sheaf =X/KHom¯(,X/K1𝒪X){}_{X/K}=\underline{\operatorname{Hom}}({}^{1}_{X/K},\mathscr{O}_{X}) is invertible of degree pnp^{n}.

Proof.

The main task is to compute the module of Kähler differentials for the integral domain K[T]K[T]. In light of Proposition 6.3, K[T]/K1{}^{1}_{K[T]/K} is generated by the n+1n+1 differentials dxjdx_{j} and dydy, modulo the nn relations

(6.3) yp2dyandxn1pdxj+1(0jn2).y^{p-2}dy\quad\text{and}\quad x_{n-1}^{p}dx_{j+1}\quad(0\leq j\leq n-2).

The ring elements yy and xn1x_{n-1} are non-zero because they correspond to monomials in K[T]K[T], so dydy and dxj+1dx_{j+1} for 0jn20\leq j\leq n-2 are torsion. We infer that the map K[T]K[T]/K1K[T]\rightarrow{}^{1}_{K[T]/K} given by the remaining differential dx0dx_{0} is bijective modulo torsion. The latter differential is given by dTpn1dT^{p^{n}-1}.

Let 𝒩\mathscr{N} be the quotient of X/K1{}^{1}_{X/K} by its torsion subsheaf, and consider the affine open covering X=U0U1X=U_{0}\cup U_{1} with U0=SpecK[T]U_{0}=\operatorname{Spec}K[T] and U1=SpecK[T1]U_{1}=\operatorname{Spec}K[T^{-1}]. We have trivializations 𝒩|U0\mathscr{N}|U_{0} and 𝒩|U1\mathscr{N}|U_{1}, given by dTpn1dT^{p^{n}-1} and dT1dT^{-1}. On the overlap these become Tpn2dT-T^{p^{n}-2}dT and T2dT-T^{-2}dT, which are related by the cocycle Tpn(U0U1,𝒪X×)T^{p^{n}}\in\Gamma(U_{0}\cap U_{1},\mathscr{O}_{X}^{\times}). This gives deg(𝒩)=pn\deg(\mathscr{N})=-p^{n}. The assertion for the dual sheaf =X/K𝒩{}_{X/K}=\mathscr{N}^{\vee} is immediate. ∎

7.  The projective model

We keep the set-up of the previous section and now describe a projective model for our curve X=Xp,nX=X_{p,n}. First note that the nn obvious relations (6.2) for our monoid =p,n\Gamma={}_{p,n} can be replaced by

(7.1) pan1=(p1)bandpaj=(p1)b+aj+1(0jn2)p\cdot a_{n-1}=(p-1)\cdot b\quad\text{and}\quad p\cdot a_{j}=(p-1)\cdot b+a_{j+1}\quad(0\leq j\leq n-2)

by using the first of these relations. Now write Pn+1=ProjK[U0,,Un1,V,Z]\mathbb{P}^{n+1}=\operatorname{Proj}K[U_{0},\ldots,U_{n-1},V,Z], and consider the closed subscheme C=Cp,nC=C_{p,n} defined by the nn homogeneous equations

(7.2) Un1pVp1Z=0andUjpVp1Uj+1=0(0jn2).U_{n-1}^{p}-V^{p-1}Z=0\quad\text{and}\quad U_{j}^{p}-V^{p-1}U_{j+1}=0\quad(0\leq j\leq n-2).

First observe that CC is covered by D+(Z)D+(V)D_{+}(Z)\cup D_{+}(V) because it contains only the point (0::0:1:0)(0:\ldots:0:1:0) on the hyperplane given by Z=0Z=0. On these two charts, we see that

TpnpjUj/Z,TpnV/Z,andT1U0/VT^{p^{n}-p^{j}}\longmapsto U_{j}/Z,\quad T^{p^{n}}\longmapsto V/Z,\quad\text{and}\quad T^{-1}\longmapsto U_{0}/V

constitute an isomorphism CXC\rightarrow X, which we regard as an identification.

Proposition 7.1.

The homogeneous polynomials (7.2) form a regular sequence in the polynomial ring, the curve XPn+1X\subset\mathbb{P}^{n+1} has degree pnp^{n}, and

ωX=𝒪X(npn2)and=X/k𝒪X(1).\omega_{X}=\mathscr{O}_{X}(np-n-2)\quad\text{and}\quad{}_{X/k}=\mathscr{O}_{X}(1).

In particular, X/k is very ample, and ωX=X/kr\omega_{X}={}_{X/k}^{\otimes r} with exponent r=npn2r=np-n-2.

Proof.

Let a\mathfrak{a} be the ideal generated by the nn homogeneous polynomials (7.2) inside the (n+2)(n+2)-dimensional Cohen–Macaulay ring A=K[U0,,Un1,V,Z]A=K[U_{0},\ldots,U_{n-1},V,Z]. Since the scheme CC is one-dimensional, we must have dim(A/a)=2\dim(A/\mathfrak{a})=2. It follows from [Sta18, Tag 02JN] that the polynomials in question form a regular sequence. The assertion on the dualizing sheaf immediately follows from ωPn+1=𝒪Pn+1(n2)\omega_{\mathbb{P}^{n+1}}=\mathscr{O}_{\mathbb{P}^{n+1}}(-n-2) and the adjunction formula.

The intersection of CPn+1C\subset\mathbb{P}^{n+1} with the hyperplane given by V=0V=0 is a singleton, with generators Uj/ZU_{j}/Z and relations (Uj/Z)p=0(U_{j}/Z)^{p}=0 for 0jn10\leq j\leq{n-1} in the homogeneous coordinate ring. Thus deg(C)=pn\deg(C)=p^{n}.

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 TpnT^{-p^{n}} with respect to the open covering W0=SpecK[T]W_{0}=\operatorname{Spec}K[T] and W1=SpecK[T1]W_{1}=\operatorname{Spec}K[T^{-1}]. The latter correspond to the open sets D+(Z)D_{+}(Z) and D+(V)D_{+}(V). On the union of these open sets, the invertible sheaf 𝒪Pn(1)\mathscr{O}_{\mathbb{P}^{n}}(1) is defined by the cocycle Z/VZ/V. This becomes TpnT^{-p^{n}} after restricting to CC, and thus =X/k𝒪X(1){}_{X/k}=\mathscr{O}_{X}(1). ∎

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 XX comes with what one might call an rr-fold theta characteristic or spin structure.

Another highly relevant consequence: the very ample sheaf 𝒪X(1)=X/K\mathscr{O}_{X}(1)={}_{X/K} has an intrinsic meaning, and g=H0(X,𝒪X(1))\mathfrak{g}=H^{0}(X,\mathscr{O}_{X}(1)) becomes the Lie algebra for the automorphism group scheme G=AutX/KG=\operatorname{Aut}_{X/K}. To exploit this, we check that the closed embedding XPn+1X\subset\mathbb{P}^{n+1} is defined by the complete linear system.

Proposition 7.2.

The restriction map H0(Pn+1,𝒪Pn+1(1))H0(X,𝒪X(1))H^{0}(\mathbb{P}^{n+1},\mathscr{O}_{\mathbb{P}^{n+1}}(1))\rightarrow H^{0}(X,\mathscr{O}_{X}(1)) is bijective. In particular, h0()X/K=n+2h^{0}({}_{X/K})=n+2.

Proof.

Since the defining polynomials (7.2) have degree p2p\geq 2, the homogeneous ideal for XPn+1X\subset\mathbb{P}^{n+1} contains no linear terms. It follows that the map in question is injective. It remains to compute h0()h^{0}(\mathscr{L}) for =X/K\mathscr{L}={}_{X/K}.

Let us proceed with some general considerations on invertible sheaves \mathscr{L} on XX of arbitrary degree m0m\geq 0. Recall that the conductor loci for the normalization map f:P1Xf\colon\mathbb{P}^{1}\rightarrow X are given by

(𝒪A)=K[T]/(Tc)and(𝒪B)=K[T]/(Tc,Tc+1,),\Gamma(\mathscr{O}_{A})=K[T]\,/(T^{c})\quad\text{and}\quad\Gamma(\mathscr{O}_{B})=K[T]\,/(T^{c},T^{c+1},\ldots),

where c0c\geq 0 is the conductor for the numerical semigroup . From the cocartesian diagram (5.2), we now obtain an exact sequence

0H0()(P1)(B)(A)H1()0.0\longrightarrow H^{0}(\mathscr{L})\longrightarrow\Gamma(\mathscr{L}_{\mathbb{P}^{1}})\oplus\Gamma(\mathscr{L}_{B})\longrightarrow\Gamma(\mathscr{L}_{A})\longrightarrow H^{1}(\mathscr{L})\longrightarrow 0.

It is not difficult to determine the map in the middle: Making the identification P1=𝒪P1(m)\mathscr{L}_{\mathbb{P}^{1}}=\mathscr{O}_{\mathbb{P}^{1}}(m) and B=𝒪B\mathscr{L}_{B}=\mathscr{O}_{B} and A=𝒪A\mathscr{L}_{A}=\mathscr{O}_{A}, we get

(P1)=a=0,,mKTa,(B)=a,ac1KTa,and(A)=a=0,,c1KTa.\Gamma(\mathscr{L}_{\mathbb{P}^{1}})=\bigoplusop\displaylimits_{a=0,\ldots,m}KT^{a},\quad\Gamma(\mathscr{L}_{B})=\bigoplusop\displaylimits_{a\in\Gamma,a\leq c-1}KT^{a},\quad\text{and}\quad\Gamma(\mathscr{L}_{A})=\bigoplusop\displaylimits_{a=0,\ldots,c-1}KT^{a}.

If mc1m\leq c-1, the former groups are contained in the latter, and H0()H^{0}(\mathscr{L}) becomes their intersection, and hence h0()=Card(S)h^{0}(\mathscr{L})=\operatorname{Card}(S) for the set

S={aam and ac1}={aam}.S=\{a\in\Gamma\mid\text{$a\leq m$ and $a\leq c-1$}\}=\{a\in\Gamma\mid\text{$a\leq m$}\}.

We have to determine this set for m=pnm=p^{n}, under the assumption n(p1)3n(p-1)\geq 3. According to Proposition 6.1, the conductor is c=npn+1(n+2)pn+2c=np^{n+1}-(n+2)p^{n}+2, and thus c/pn=np(n+2)+2/pn>n(p1)21c/p^{n}=np-(n+2)+2/p^{n}>n(p-1)-2\geq 1. So our set SS comprises all aa\in\Gamma with apna\leq p^{n}. It clearly contains the generators pn,pn1,,pnpn1p^{n},p^{n}-1,\ldots,p^{n}-p^{n-1} and also the zero element a=0a=0. It remains to check that for each pair of generators aba\leq b, we have a+bpna+b\geq p^{n}. This is obvious for b=pnb=p^{n}, so assume a=pnpia=p^{n}-p^{i} and b=pnpjb=p^{n}-p^{j} with 0ij<n0\leq i\leq j<n. Then a+bpn=pnpipjpn2pjpnpj+10a+b-p^{n}=p^{n}-p^{i}-p^{j}\geq p^{n}-2p^{j}\geq p^{n}-p^{j+1}\geq 0. ∎

This leads to a matrix interpretation of the full automorphism group scheme G=AutX/KG=\operatorname{Aut}_{X/K}: First note that the diagonal action of GG on X×XX\times X, and its effects on graphs, induces the conjugacy action of GG on itself. Its restriction to the first infinitesimal neighborhood of the diagonal X yields the GG-linearization of the tangent sheaf X/k, and we infer that the resulting representation on the Lie algebra g=H0(X,)X/K\mathfrak{g}=H^{0}(X,{}_{X/K}) coincides with the adjoint representation GGL(g)G\rightarrow\operatorname{GL}(\mathfrak{g}). Its projectivization GPGL(g)G\rightarrow\operatorname{PGL}(\mathfrak{g}) is injective because X/k is very ample, and it follows that GGL(g)G\rightarrow\operatorname{GL}(\mathfrak{g}) is injective as well. We thus have a canonical inclusion GGL(g)G\subset\operatorname{GL}(\mathfrak{g}) that intersects the center GmGL(g)\mathbb{G}_{m}\subset\operatorname{GL}(\mathfrak{g}) trivially. Write GGm=G×GmG\cdot\mathbb{G}_{m}=G\times\mathbb{G}_{m} for the resulting subgroup scheme and apSymp(g)\mathfrak{a}_{p}\subset\operatorname{Sym}^{p}(\mathfrak{g}) for the vector subspace generated by the homogeneous polynomials (7.2).

Proposition 7.3.

We keep the notation as above. Then GGmGL(g)G\cdot\mathbb{G}_{m}\subset\operatorname{GL}(\mathfrak{g}) equals the stabilizer group scheme for the vector subspace apSymp(g)\mathfrak{a}_{p}\subset\operatorname{Sym}^{p}(\mathfrak{g}).

Proof.

We start with some observations on the homogeneous coordinate rings

(𝒪Pn+1)=Sym(g)and(𝒪X)=n0(X,𝒪X(n)).{}_{\bullet}(\mathscr{O}_{\mathbb{P}^{n+1}})=\operatorname{Sym}^{\bullet}(\mathfrak{g})\quad\text{and}\quad{}_{\bullet}(\mathscr{O}_{X})=\bigoplusop\displaylimits_{n\geq 0}\Gamma(X,\mathscr{O}_{X}(n)).

Both rings are integral, so the kernel p\mathfrak{p} of the canonical map (𝒪Pn+1)(𝒪X){}_{\bullet}(\mathscr{O}_{\mathbb{P}^{n+1}})\rightarrow{}_{\bullet}(\mathscr{O}_{X}) is a prime ideal, which equals the radical for the ideal a\mathfrak{a} generated by the polynomials (7.2). The ideal a\mathfrak{a} becomes prime when localized with respect to any homogeneous fSym+(g)f\in\operatorname{Sym}^{+}(\mathfrak{g}) because XX is integral. Since our generators form a regular sequence, this actually holds everywhere, and thus a=p\mathfrak{a}=\mathfrak{p}.

Write IGL(g)I\subset\operatorname{GL}(\mathfrak{g}) for the stabilizer group scheme in question. It contains Gm\mathbb{G}_{m} because the polynomials are homogeneous, and its action on Pn+1\mathbb{P}^{n+1} stabilizes the curve XX. Modulo Gm\mathbb{G}_{m}, the induced action on XX is faithful, according to Proposition 7.2. Thus IGGmI\subset G\cdot\mathbb{G}_{m}. Conversely, let fG(R)f\in G(R) be some RR-valued automorphism of XX. It induces an action on the homogeneous coordinate rings (𝒪Pn+1R)=(𝒪Pn+1)R{}_{\bullet}(\mathscr{O}_{\mathbb{P}^{n+1}}\otimes R)={}_{\bullet}(\mathscr{O}_{\mathbb{P}^{n+1}})\otimes R, and likewise for (𝒪X){}_{\bullet}(\mathscr{O}_{X}). These actions are compatible; thus ff stabilizes ppR=apR\mathfrak{p}_{p}\otimes R=\mathfrak{a}_{p}\otimes R. ∎

For n=1n=1, this means that GGmGL3G\cdot\mathbb{G}_{m}\subset\operatorname{GL}_{3} is the stabilizer group scheme for the line generated by UpVp1ZU^{p}-V^{p-1}Z in the pthp^{\mathrm{th}} symmetric power of g=KUKVKZ\mathfrak{g}=KU\oplus KV\oplus KZ. In turn, GPGL3G\subset\operatorname{PGL}_{3} is the inertia group scheme for the rational point corresponding to this line.

8.  The automorphism group scheme

Recall that our curves X=Xp,nX=X_{p,n} come with an inclusion

(8.1) GaUnGmAutX/K\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}\subset\operatorname{Aut}_{X/K}

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 pn3p^{n}\geq 3.

The cases pn2p^{n}\leq 2 indeed have to be excluded because then X=P1X=\mathbb{P}^{1}. Also note that for 3pn43\leq p^{n}\leq 4, our curve XX 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 pn3p^{n}\geq 3, the ideal for the closed embedding (8.1) is nilpotent.

Proof.

For this, we may assume that KK is algebraically closed. Seeking a contradiction, we suppose that there is an automorphism φ:XX\varphi\colon X\rightarrow X that does not yield a rational point in GaGm\mathbb{G}_{a}\rtimes\mathbb{G}_{m}. The assumption ensures Sing(X)={x0}\operatorname{Sing}(X)=\{x_{0}\}, and φ\varphi fixes this singular point. Hence the induced automorphism on the normalization P1=SpecK[T]SpecK[T1]\mathbb{P}^{1}=\operatorname{Spec}K[T]\cup\operatorname{Spec}K[T^{-1}] fixes the point defined by T=0T=0. It thus belongs to the inertia group scheme GaGm\mathbb{G}_{a}\rtimes\mathbb{G}_{m} inside PGL2\operatorname{PGL}_{2}. According to [Bri17, Proposition 2.5.1], the action of any smooth group scheme on XX lifts to an action on the normalization. Thus φ\varphi belongs to GaGm\mathbb{G}_{a}\rtimes\mathbb{G}_{m} inside AutX/k\operatorname{Aut}_{X/k}, giving a contradiction. ∎

Proof of Theorem 8.1 in the special case n=1n=1.

In this situation we have X=SpecK[Tp,Tp1]SpecK[T1]X=\operatorname{Spec}K[T^{p},T^{p-1}]\cup\operatorname{Spec}K[T^{-1}]. According to Proposition 7.2, the tangent sheaf has h0()X/k=3h^{0}({}_{X/k})=3. One easily computes that the rational vector fields

(8.2) T2pT,T2T,andTTT^{2-p}\frac{\partial}{\partial T},\quad T^{2}\frac{\partial}{\partial T},\quad\text{and}\quad T\frac{\partial}{\partial T}

are everywhere defined, hence form a basis of g=H0(X,)X/k\mathfrak{g}=H^{0}(X,{}_{X/k}). Moreover, the first two basis vectors generate a restricted subalgebra K2K^{2}, with trivial bracket and pp-map, and the last basis vector yields a copy of gl1(K)\operatorname{\mathfrak{gl}}_{1}(K), giving a semidirect product K2gl1(K)K^{2}\rtimes\operatorname{\mathfrak{gl}}_{1}(K). Such algebras play a prominent role in [Sch07, KS21, ST23]. The bracket and pp-map are given by [(x,λ),(x,λ)]=λxλx[(x,\lambda),(x^{\prime},\lambda^{\prime})]=\lambda x^{\prime}-\lambda^{\prime}x and (x,λ)[p]=(λpx,λp1)(x,\lambda)^{[p]}=(\lambda^{p}x,\lambda^{p-1}); cf. [KS21, Proposition 1.1]. One sees that K2K^{2} is the image of the bracket, thus the derived subalgebra. This also holds for RR-valued points; hence K2K^{2} is a subrepresentation for GG.

As explained in Section 7, our curve XX may also be regarded as the curve in P2=ProjK[U0,V,Z]\mathbb{P}^{2}=\operatorname{Proj}K[U_{0},V,Z] defined by the homogeneous polynomial P=U0pVp1ZP=U_{0}^{p}-V^{p-1}Z, with an identification via Tp1=U0/ZT^{p-1}=U_{0}/Z, Tp=V/ZT^{p}=V/Z, and T1=U0/VT^{-1}=U_{0}/V. According to Proposition 7.2, the monomials Z,V,U0Z,V,U_{0} yield a basis for H0(X,𝒪X(1))H^{0}(X,\mathscr{O}_{X}(1)). By Proposition 7.1, the invertible sheaves 𝒪X(1)\mathscr{O}_{X}(1) and X are isomorphic. Computing the order of zeros for the homogeneous polynomials Z,V,U0Z,V,U_{0} and the vector fields (8.2) on P1\mathbb{P}^{1}, one sees that each identification 𝒪X(1)=X\mathscr{O}_{X}(1)={}_{X} 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 RR-valued points of H=GGmH=G\cdot\mathbb{G}_{m} as the group of matrices

A=(abcdef00g)GL3(R)A=\begin{pmatrix}a&b&c\\ d&e&f\\ 0&0&g\end{pmatrix}\in\operatorname{GL}_{3}(R)

subject to the sole condition

(8.3) P(aU0+dV,bU0+eV,cU0+fV+gZ)=λP(U0,V,Z),P(aU_{0}+dV,bU_{0}+eV,cU_{0}+fV+gZ)=\lambda\cdot P(U_{0},V,Z),

with some multipliers λR×\lambda\in R^{\times}. The zero entries in the matrix AA stem from the fact that the derived subalgebra [g,g]g[\mathfrak{g},\mathfrak{g}]\subset\mathfrak{g} is a subrepresentation.

Suppose that AH(R)A\in H(R) has b=0b=0. Comparing coefficients in (8.3), we get aep1=gpae^{p-1}=g^{p}, cp=0c^{p}=0, and de=fpde=f^{p}, so the matrix takes the form

A=(gpe1p0cfpe1ef00g)with cp=0.A=\begin{pmatrix}g^{p}e^{1-p}&0&c\\ f^{p}e^{-1}&e&f\\ 0&0&g\end{pmatrix}\quad\text{with $c^{p}=0$.}

The group H0HH_{0}\subset H of such matrices contains the diagonal copy of Gm\mathbb{G}_{m}, and H0/GmH_{0}/\mathbb{G}_{m} becomes our iterated semidirect product GaαpGm\mathbb{G}_{a}\rtimes\alpha_{p}\rtimes\mathbb{G}_{m} inside AutX/kPGL3\operatorname{Aut}_{X/k}\subset\operatorname{PGL}_{3}. Note that the projection H0H0/GmH_{0}\rightarrow H_{0}/\mathbb{G}_{m} admits a splitting, obtained by setting g=1g=1.

Seeking a contradiction, we assume that there is some AH(R)A\in H(R) with b0b\neq 0. By Lemma 8.2, we must have bNil(R)b\in\operatorname{Nil}(R), and thus a,eR×a,e\in R^{\times}. Making a flat extension of RR, we can assume that there is some fRf^{\prime}\in R with fp=d/af^{\prime p}=-d/a. Setting a=e=g=1a^{\prime}=e^{\prime}=g^{\prime}=1 and b=c=0b^{\prime}=c^{\prime}=0 and left-multiplying with the resulting matrix AH(R)A^{\prime}\in H(R), we may assume that both b0b\neq 0 and d=0d=0 hold. On the other hand, comparing coefficients in (8.3) immediately yields b=0b=0, giving a contradiction.

This establishes H0=HH_{0}=H. We already observed that H/Gm=AutX/KH/\mathbb{G}_{m}=\operatorname{Aut}_{X/K} inside PGL3\operatorname{PGL}_{3} and that H0/GmH_{0}/\mathbb{G}_{m} equals our iterated semidirect product. ∎

To continue inductively, we seek to relate the curves Xp,nX_{p,n} with different indices nn. First recall a general fact on numerical semigroups =b,a2,,ar\Gamma=\langle b,a_{2},\ldots,a_{r}\rangle with non-zero generators b<a2<<arb<a_{2}<\ldots<a_{r}: the blowing-up Blm(A)\operatorname{Bl}{m}(A) of the ring A=K[T]A=K[T] with respect to the maximal ideal m=(Tb,Ta2,,Tar)\mathfrak{m}=(T^{b},T^{a_{2}},\ldots,T^{a_{r}}) has coordinate ring A=K[T]A^{\prime}=K[T^{{}^{\prime}}] for the numerical semigroup =b,a2b,,arb{}^{\prime}=\langle b,a_{2}-b,\ldots,a_{r}-b\rangle; cf. [BDF97, Proposition I.2.1, and also Equation I.2.4].

For our =p,n\Gamma={}_{p,n} with n1n\geq 1, we have b=pnpn1b=p^{n}-p^{n-1}, with remaining generators pnp^{n} and pnpjp^{n}-p^{j} for 0jn20\leq j\leq n-2. The resulting differences are pnb=pn1p^{n}-b=p^{n-1} and (pnpj)b=pn1pj(p^{n}-p^{j})-b=p^{n-1}-p^{j}. For n2n\geq 2, we write b=p(pn1pn2)b=p\cdot(p^{n-1}-p^{n-2}) and in any case see =p,n1{}^{\prime}={}_{p,n-1}. This reveals the following.

Lemma 8.3.

For pn3p^{n}\geq 3, we have Xp,n1=BlZ(Xp,n)X_{p,n-1}=\operatorname{Bl}_{Z}(X_{p,n}), where the center ZZ is the singular point x0Xp,nx_{0}\in X_{p,n} endowed with the reduced scheme structure.

Note that for every mnm\leq n, we get an inclusion p,np,m{}_{p,n}\subset{}_{p,m} of numerical semigroups inside N\mathbb{N}. The resulting inclusions of coordinate rings K[Tp,n]K[Tp,m]K[T^{{}_{p,n}}]\subset K[T^{{}_{p,m}}] define canonical morphisms Xp,mXp,nX_{p,m}\rightarrow X_{p,n} of compactifications of the affine line A1=SpecK[T1]\mathbb{A}^{1}=\operatorname{Spec}K[T^{-1}].

Proof of Theorem 8.1 in the general case.

We proceed by induction on n1n\geq 1. The case n=1n=1 was handled above. Now suppose n2n\geq 2 and that the assertion is true for n1n-1. To simplify notation, set

X=Xp,n,G=AutX/K,andH=GaUnGm,X=X_{p,n},\quad G=\operatorname{Aut}_{X/K},\quad\text{and}\quad H=\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m},

and let IGI\subset G be the inertia group scheme for the singularity x0Xx_{0}\in X. Likewise, we set X=Xp,n1X^{\prime}=X_{p,n-1}. By induction, H=GaUn1GmH^{\prime}=\mathbb{G}_{a}\rtimes U_{n-1}\rtimes\mathbb{G}_{m} coincides with G=AutX/KG^{\prime}=\operatorname{Aut}_{X^{\prime}/K}. Also note that by the very definition of the group schemes, we have a canonical inclusion HHH^{\prime}\subset H. Moreover, with Proposition 5.5 we get an inclusion HIH^{\prime}\subset I, and actually H=IHH^{\prime}=I\cap H. By Lemma 8.3 combined with [Mar22, Proposition 2.7], the blowing-up morphism f:XXf\colon X^{\prime}\rightarrow X is equivariant with respect to the action of II. We thus get inclusions HIG=HH^{\prime}\subset I\subset G^{\prime}=H^{\prime} and infer H=IH^{\prime}=I.

The orbit map for the rational point x0Xx_{0}\in X gives an inclusion H/HG/IH/H^{\prime}\subset G/I of closed subschemes inside XX. This is actually contained in the scheme of singularities Z=Sing(X/K)Z=\operatorname{Sing}(X/K), 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 h0(𝒪H/H)=pnh^{0}(\mathscr{O}_{H/H^{\prime}})=p^{n}, and it remains to verify h0(𝒪G/I)pnh^{0}(\mathscr{O}_{G/I})\leq p^{n}. For this, we may assume that KK is algebraically closed.

According to (6.3) and Proposition 6.3, the scheme of singularities Z=Sing(X/K)Z=\operatorname{Sing}(X/K) has coordinate ring of the form A=K[x0,,xn1,y]/(x0p,,xn1p,yp2)A=K[x_{0},\ldots,x_{n-1},y]\,/(x_{0}^{p},\ldots,x_{n-1}^{p},y^{p-2}). In light of [DG70, Section III.3, Theorem 6.1], the homogeneous space G/IG/I has coordinate ring of the form B=K[u1,,ur]/(u1pν1,,urpνr)B=K[u_{1},\ldots,u_{r}]\,/(u_{1}^{p^{\nu_{1}}},\ldots,u_{r}^{p^{\nu_{r}}}) for some r0r\geq 0 and certain exponents νi1\nu_{i}\geq 1. From the canonical surjection

φ:A=(𝒪Z)(𝒪G/I)=B,\varphi\colon A=\Gamma(\mathscr{O}_{Z})\longrightarrow\Gamma(\mathscr{O}_{G/I})=B,

we infer νi=1\nu_{i}=1 and rn+1r\leq n+1. Using the relation yp2=0y^{p-2}=0, we see that φ(y)mB\varphi(y)\in\mathfrak{m}_{B} must be contained in mB2\mathfrak{m}_{B}^{2}, and thus actually rnr\leq n. It follows that h0(𝒪G/I)pnh^{0}(\mathscr{O}_{G/I})\leq p^{n}, as desired. ∎

9.  Equivariant normality and twisting

We now seek to construct twisted forms of our curves Xp,nX_{p,n} 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 KK, and let XX be a scheme. Recall that another scheme YY is called a twisted form of XX if we have YLXLY\otimes L\simeq X\otimes L for some field extension KLK\subset L. Such twisted forms may arise as follows: Suppose that a group scheme GG acts on XX, and let PP be a GG-torsor. Then GG acts diagonally on P×XP\times X, and the quotient

XP=G\(P×X){}^{P}\!X=G\backslash(P\times X)

is a twisted form of XX. 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 GG is finite and XX is covered by affine open sets that are GG-stable, the twisted form is indeed a scheme (cf. [DG70, Section III.2, Theorem 3.2]).

Now suppose that we are in positive characteristic p>0p>0. 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 XX is a separated scheme of finite type that is geometrically integral. It is normal if all local rings 𝒪X,a\mathscr{O}_{X,a} are integrally closed in the common function field F=Frac(𝒪X,a)F=\operatorname{Frac}(\mathscr{O}_{X,a}). Equivalently, each finite modification f:XXf\colon X^{\prime}\rightarrow X is an isomorphism. Here the term modification refers to an integral scheme XX^{\prime}, together with a proper surjective morphism f:XXf\colon X^{\prime}\rightarrow X inducing a bijection on function fields.

In what follows, we suppose that XX is endowed with the action of a finite group scheme GG. We now consider only modifications f:XXf\colon X^{\prime}\rightarrow X where XX^{\prime} is a GG-scheme and ff is equivariant. For brevity, we call such a datum an equivariant modification or GG-modification. Examples are given by blowing-ups X=BlZ(X)X^{\prime}=\operatorname{Bl}_{Z}(X) with respect to GG-stable centers ZXZ\subset X. Note that for a given modification XX^{\prime}, there is at most one GG-action on XX^{\prime} making f:XXf\colon X^{\prime}\rightarrow X equivariant.

One says that XX is equivariantly normal, or GG-normal, if every finite equivariant modification f:XXf\colon X^{\prime}\rightarrow X is an isomorphism. This extremely useful notion was introduced and studied by Brion in [Bri22a]. He showed that XX admits a finite equivariant modification X~\tilde{X} that is equivariantly normal (cf. [Bri22a, Proposition 4.2]). It is actually unique up to unique equivariant isomorphism provided that XX is one-dimensional (cf. [Bri22a, Corollary 4.4]).

From now on, we furthermore assume that our GG-scheme XX is one-dimensional. One could also say that XX is a GG-curve. Following [FS20, Section 2] we write Sing(X/K)\operatorname{Sing}(X/K) for the closed subscheme defined by the first Fitting ideal for X/K1{}^{1}_{X/K}. This is the set of points aXa\in X where the local ring 𝒪X,a\mathscr{O}_{X,a} fails to be geometrically regular, endowed with a canonical scheme structure. Note that with respect to this scheme structure, it must be GG-stable (cf. [BS22, Proposition 3.1]). The existence of twisted forms that are regular is intimately related to equivariant normality.

Theorem 9.1.

Let PP be a GG-torsor. Then the twisted form Y=XPY={}^{P}\!X is regular provided the following three conditions hold:

  1. (i)

    The curve XX is GG-normal.

  2. (ii)

    The total space of the GG-torsor PP is reduced.

  3. (iii)

    The reduction of the finite scheme Sing(X/K)\operatorname{Sing}(X/K) is étale.

Proof.

The residue fields κ(a)\kappa(a) for the points aSing(X/K)a\in\operatorname{Sing}(X/K) are separable, by assumption (iii), and so is their join LL. This ensures that the base-change PLP_{L} remains reduced. Furthermore, the arguments for [Bri22a, Proposition 4.10] show that XLX_{L} remains equivariantly normal. Replacing the ground field with LL, we may assume that Sing(X/K)={a1,,ar}\operatorname{Sing}(X/K)=\{a_{1},\ldots,a_{r}\} comprises only rational points. Let HiGH_{i}\subset G be the inertia subgroup scheme and Zi=Gai=G/HiZ_{i}=G\cdot a_{i}=G/H_{i} be the orbit for aiXa_{i}\in X. Clearly, the subscheme Z1ZrZ_{1}\cup\ldots\cup Z_{r} and its complementary open set UU are GG-stable. The latter is geometrically regular, and so is the twisted form UP{}^{P}\!U.

It remains to verify that the integral curve Y=XPY={}^{P}\!X is regular at the points bZiPb\in{}^{P}\!Z_{i}. According to [Bri22a, Theorem 4.13], the orbit ZiXZ_{i}\subset X is an effective Cartier divisor. In turn, its twist ZiPXP{}^{P}\!Z_{i}\subset{}^{P}\!X is an effective Cartier divisor on XP{}^{P}\!X, so it suffices to verify that it is reduced. The latter becomes the quotient of P×G/HiP\times G/H_{i} by the diagonal GG-action, which can be identified with Hi\PH_{i}\backslash P. Its coordinate ring is a subring inside (P,𝒪P)\Gamma(P,\mathscr{O}_{P}), which can be seen as a ring of invariants, and (P,𝒪P)\Gamma(P,\mathscr{O}_{P}) is reduced by assumption. ∎

Note that condition (iii) holds in particular if all aSing(X/K)a\in\operatorname{Sing}(X/K) are rational points. The first two conditions can be achieved after ground field extensions.

Proposition 9.2.

Suppose that the curve XX is GG-normal. Then there is a field extension KLK\subset L such that the following hold:

  1. (i)

    The base-change XLX_{L} is GLG_{L}-normal.

  2. (ii)

    There is a GLG_{L}-torsor PP whose total space is reduced.

Proof.

(ii)  Choose a geometrically integral quasi-projective GG-scheme UU with generically free action. This could arise from an embedding GHG\subset H into a smooth group scheme of finite type or could arise from a projective scheme XX with G=AutX/K0G=\operatorname{Aut}^{0}_{X/K}, according to [BS22, Proposition 1.7 or Theorem 2.1]. The quotient V=U/GV=U/G is an integral quasi-projective scheme, and the quotient map f:UVf\colon U\rightarrow V induces a finite extension of the function field L=k(V)L=k(V) by E=k(U)E=k(U). By construction, the reduced scheme P=Spec(E)P=\operatorname{Spec}(E), viewed as an LL-scheme, is a torsor with respect to the base-change GL=GKLG_{L}=G\otimes_{K}L.

(i)  The above extension KEK\subset E is separable because UU is geometrically reduced. In turn, the subextension LL is also separable. The arguments for [Bri22a, Proposition 4.10] show that XLX_{L} remains equivariantly normal. ∎

In the reverse direction, we have the following result.

Theorem 9.3.

Suppose that there exist a field extension KLK\subset L, a subgroup scheme HGLH\subset G_{L}, and a HH-torsor PP so that the twisted form (XL)P{}^{P}\!(X_{L}) is regular. Then the curve XX is GG-normal.

Proof.

According to [Bri22a, Proposition 4.10 and Remark 4.3], it suffices to treat the case L=KL=K and H=GH=G. Let X=XPX^{\prime}={}^{P}\!X be the regular twisted form. According to [ST23, Lemma 3.1], we have a canonical identification AutX/K=AutX/KP\operatorname{Aut}_{X^{\prime}/K}={}^{P}\!\operatorname{Aut}_{X/K} of the sheaves of automorphisms, where the term on the right is formed with respect to the conjugacy action of GG on AutX/K\operatorname{Aut}_{X/K}. Setting G=GPG^{\prime}={}^{P}G, we get a homomorphism GAutX/KG^{\prime}\rightarrow\operatorname{Aut}_{X^{\prime}/K}, hence a GG^{\prime}-action on XX^{\prime}.

Let ZXZ\subset X be a finite closed subscheme that is GG-stable. According to [Bri22a, Theorem 4.13], we have to check that ZZ is Cartier. Its twist Z=ZPZ^{\prime}={}^{P}\!Z defines a finite closed subscheme inside XX^{\prime}. The latter is regular, so ZZ^{\prime} is Cartier. Choose a point pPp\in P, and let E=κ(p)E=\kappa(p) be the resulting field extension. The resulting trivialization of PEP\otimes E defines an isomorphism g:XEXEg\colon X\otimes E\rightarrow X^{\prime}\otimes E with g(ZE)=ZEg(Z\otimes E)=Z^{\prime}\otimes E. It follows that ZEZ\otimes E, and hence ZZ, is Cartier. ∎

Recall that a quasielliptic curve is a regular curve YY that is a twisted form of the rational cuspidal curve

X=SpecK[T2,T3]SpecK[T1].X=\operatorname{Spec}K\left[T^{2},T^{3}\right]\cup\operatorname{Spec}K\left[T^{-1}\right].

Clearly, Sing(X/K)\operatorname{Sing}(X/K) is a singleton, containing only the rational point x0x_{0} given by T2=T3=0T^{2}=T^{3}=0. It turns out that quasielliptic curves exist only in characteristic two and three (compare with the discussion after Proposition 11.1). For p=2p=2, the rational vector field D=T2/TD=T^{-2}\partial/\partial T satisfies D[p]=0D^{[p]}=0 and actually defines a global section DH0(X,)X/KD\in H^{0}(X,{}_{X/K}), hence corresponds to an action of G=αpG=\alpha_{p}, which is the Frobenius kernel of the additive group Ga\mathbb{G}_{a}. The orbit Gx0G\cdot x_{0} is the Cartier divisor defined by T2=0T^{2}=0. For p=3p=3, the same holds for D=/TD=\partial/\partial T and the Cartier divisor T3=0T^{3}=0.

In both cases, we conclude that the rational cuspidal curve is equivariantly normal with respect to G=αpG=\alpha_{p} (again by [Bri22a, Theorem 4.13]). For this group scheme, torsors with regular total space exist if and only if KK is imperfect (see for example [ST23, Lemma 7.1]), and then quasielliptic curves exist by Theorem 9.1. Also note that XX 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 AutX/K=GaUGm\operatorname{Aut}_{X/K}=\mathbb{G}_{a}\rtimes U\rtimes\mathbb{G}_{m} for an infinitesimal group scheme UU. For p=3p=3, it coincides with the copy of αp\alpha_{p} described above, whereas for p=2p=2, it has order |U|=8|U|=8.

All this generalizes to our hierarchy of curves Xp,nX_{p,n}.

Theorem 9.4.

The curve X=Xp,nX=X_{p,n} is equivariantly normal with respect to the finite group scheme UnU_{n} and locally of complete intersection. Moreover, if there is a UnU_{n}-torsor PP so that the quotient P¯=Un1\P\bar{P}=U_{n-1}\backslash P is reduced, then the twisted form Y=XPY={}^{P}\!X is regular.

Proof.

The first assertion follows from [Bri22a, Theorem 4.13 and Corollary 4.18]. If PP itself is reduced, the assertion on the twisted form Y=XPY={}^{P}\!X directly follows from Theorem 9.1. Its proof actually shows our slightly stronger claim because the singularity x0Xx_{0}\in X is rational, with orbit Un/Un1U_{n}/U_{n-1}. ∎

Note that after some ground field extension KLK\subset L, there is a UnU_{n}-torsor PP that is reduced, according to Proposition 9.2, and our curve X=Xp,nX=X_{p,n} acquires twisted forms that are regular. However, the construction of LL relies, via [BS22, Proposition 1.7 and Theorem 2.1], among other things on embeddings of UnU_{n} into smooth group schemes HH, and here we have little control over dim(H)\dim(H) and trdeg(L)\operatorname{trdeg}(L).

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 Xp,nX_{p,n}. 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 𝒫=Sh(𝒞)\mathscr{P}={\operatorname{Sh}}(\mathcal{C}) be the topos of sheaves on some site 𝒞\mathcal{C}, having a final object SS. For any group-valued object G𝒫G\in\mathscr{P}, we write H0(S,G)H^{0}(S,G) for the group of global sections and H1(S,G)H^{1}(S,G) for the set of isomorphism classes of GG-torsors PP. The latter is an object endowed with a GG-action that is locally isomorphic to P0=GP_{0}=G with the translation action. Another widespread term is principal homogeneous spaces. For GG commutative, our H1(S,G)H^{1}(S,G) coincides with the sheaf cohomology groups. In general, however, H1(S,G)H^{1}(S,G) is merely a set, containing the class of the trivial torsor P0=GP_{0}=G as a distinguished element.

An object X~\tilde{X} is called a twisted form of an object XX if the two are locally isomorphic. If XX has a GG-action and PP is a GG-torsor, we get such a twisted form by forming the quotient X~=XP=PGX=G\(P×X)\tilde{X}={}^{P}\!X=P\wedge^{G}X=G\backslash(P\times X) with respect to the diagonal action σ(p,x)=(σp,σx)\sigma\cdot(p,x)=(\sigma p,\sigma x). Note that this could also be written as (pσ1,σx)(p\sigma^{-1},\sigma x). For G=AutX/SG=\operatorname{Aut}_{X/S}, the above construction gives an identification between the non-abelian cohomology H1(S,G)H^{1}(S,G) and the set Twist(X)\operatorname{Twist}(X) of isomorphism classes of twisted forms X~\tilde{X} of the object XX. In any case, the GG-action on XX induces a conjugacy action on AutX/S\operatorname{Aut}_{X/S}, and we have Aut(PGX)/S=PGAutX/S\operatorname{Aut}_{(P\wedge^{G}X)/S}=P\wedge^{G}\operatorname{Aut}_{X/S}; compare for example with [ST23, Lemma 3.1].

Now suppose X=GX=G as sheaves of sets without group laws, and consider the homomorphism G×GopAutX/SG\times G^{\text{\rm op}}\rightarrow\operatorname{Aut}_{X/S} given by (σ1,σ2)x=σ1xσ21(\sigma_{1},\sigma_{2})\cdot x=\sigma_{1}x\sigma_{2}^{-1}. One easily checks that the map is equivariant with respect to factorwise conjugation η(σ1,σ2)=(ησ1η1,ησ2η1)\eta\cdot(\sigma_{1},\sigma_{2})=(\eta\sigma_{1}\eta^{-1},\eta\sigma_{2}\eta^{-1}) and conjugation with inner automorphisms on AutX/S\operatorname{Aut}_{X/S}. In turn, we get an induced homomorphism

GP×GopPAut(PGX)/S=AutP/S.{}^{P}\!G\times{}^{P}\!G^{\text{\rm op}}\longrightarrow\operatorname{Aut}_{(P\wedge^{G}X)/S}=\operatorname{Aut}_{P/S}.

Note that the equation stems from the identification PGX=PP\wedge^{G}X=P. The above endows each GG-torsor PP with the additional structure of a GP{}^{P}\!G-torsor and a GopP{}^{P}\!G^{\text{\rm op}}-torsor. Furthermore, GopP{}^{P}\!G^{\text{\rm op}} is the automorphism group object of PP as a GG-torsor, and GG is the automorphism group object of PP as a GopP{}^{P}\!G^{\text{\rm op}}-torsor. In turn, we get what we like to call the torsor translation map

(10.1) H1(S,GP)H1(S,G),TPGPT,H^{1}\left(S,{}^{P}\!G\right)\longrightarrow H^{1}(S,G),\quad T\longmapsto P\wedge^{{}^{P}\!G}T,

where the quotient on the right is formed with respect to the action σ~(p,t)=(pσ~1,σ~t)\tilde{\sigma}\cdot(p,t)=(p\tilde{\sigma}^{-1},\tilde{\sigma}t) and the GG-action on PGTP\wedge^{G}T stems from the action on the first factor PP. The map (10.1) is bijective but does not respect the distinguished points: rather, it sends T0=GPT_{0}={}^{P}\!G to P=PGPT0P=P\wedge^{{}^{P}\!G}T_{0}.

Now suppose that we have a short exact sequence

(10.2) 1ABprC11\longrightarrow A\longrightarrow B\stackrel{{\scriptstyle\operatorname{pr}}}{{\longrightarrow}}C\longrightarrow 1

of group objects. Then the group H0(S,C)H^{0}(S,C) acts from the right on the set H1(S,A)H^{1}(S,A) in the following way: For each global section cH0(S,C)c\in H^{0}(S,C), the fiber Bc=pr1(c)B_{c}=\operatorname{pr}^{-1}(c) with respect to the surjection BCB\rightarrow C carries compatible AA-torsor structures from both sides, coming from the group law in BB. We now define c[P]=[BcAP]c\cdot[P]=[B_{c}\wedge^{A}P]. The stabilizer group at each torsor class is the subgroup of global sections cH0(S,C)c\in H^{0}(S,C) where the set of global sections H0(S,Bc)H^{0}(S,B_{c}) is non-empty.

Let us write H1(S,A)/H0(S,C)op=H0(S,C)\H1(S,A)H^{1}(S,A)/H^{0}(S,C)^{\text{\rm op}}=H^{0}(S,C)\backslash H^{1}(S,A) for the quotient of the action. Using the distinguished point in H1(S,A)H^{1}(S,A), the orbit map cBcc\mapsto B_{c} yields H0(S,C)H1(S,A)H^{0}(S,C)\rightarrow H^{1}(S,A). The latter serves as a connecting map and yields a six-term sequence of sets

1H0(S,A)H0(S,B)H0(S,C)H1(S,A)H1(S,B)H1(S,C).1\longrightarrow H^{0}(S,A)\longrightarrow H^{0}(S,B)\longrightarrow H^{0}(S,C)\stackrel{{\scriptstyle\partial}}{{\longrightarrow}}H^{1}(S,A)\longrightarrow H^{1}(S,B)\longrightarrow H^{1}(S,C).

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 BB acts on itself and its quotient C=B/AC=B/A via conjugacy. On the normal subgroup AA, we have an induced action. Twisting with respect to the BB-actions gives another exact sequence

(10.3) 1APBPCP1,1\longrightarrow{}^{P}\!A\longrightarrow{}^{P}\!B\longrightarrow{}^{P}\!C\longrightarrow 1,

which also yields a six-term sequence. Note that in general there is no map relating H1(S,A)H^{1}(S,A) and H1(S,AP)H^{1}(S,{}^{P}\!A) because the BB-action on AA usually fails to be inner.

We now choose for each CC-torsor P¯\bar{P} whose class belongs to the image of the mapping H1(S,B)H1(S,C)H^{1}(S,B)\rightarrow H^{1}(S,C) some BB-torsor PP with CBTP¯C\wedge^{B}T\simeq\bar{P}. As in [Ser94, Section 5.5, Corollary 2], one has the following.

Theorem 10.1.

The first cohomology of BB can be written as a disjoint union

H1(S,B)=H1(S,AP)/H0(S,CP¯)opH^{1}(S,B)=\bigcupop\displaylimits H^{1}\left(S,{}^{P}\!A\right)/H^{0}\left(S,{}^{\bar{P}}\!C\right)^{\text{\rm op}}

running over all [P¯][\bar{P}] from the image of H1(S,B)H1(S,C)H^{1}(S,B)\rightarrow H^{1}(S,C). The inclusions are obtained by composing the induced maps H1(S,AP)H1(S,BP)H^{1}(S,{}^{P}\!A)\rightarrow H^{1}(S,{}^{P}\!B) with the torsor translation maps H1(S,BP)H1(S,B)H^{1}(S,{}^{P}\!B)\rightarrow H^{1}(S,B) given by (10.1).

We are interested in cases where the above simplifies. Recall that pr:BC\operatorname{pr}\colon B\rightarrow C is the canonical epimorphism. Let us call a morphism s:CBs\colon C\rightarrow B a set-theoretical section if prs=idC\operatorname{pr}\circ s={\operatorname{id}}_{C}. The point here is that ss does not have to preserve the group laws. The resulting A×CBA\times C\rightarrow B given by (a,c)as(c)(a,c)\mapsto a\cdot s(c) is an isomorphism of objects that does not necessarily respect the group laws. The latter is determined by the two-cocycle τ:C2A\tau\colon C^{2}\rightarrow A defined by s(cc)=τc,cs(c)s(c)s(cc^{\prime})=\tau_{c,c^{\prime}}\cdot s(c)s(c^{\prime}). We like to indicate this situation by writing

B=A×~C=A×~τCB=A\tilde{\times}C=A\tilde{\times}_{\tau}C

and say that the extension CC 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 ss respects the group laws, the above becomes a semidirect product B=AC=AφCB=A\rtimes C=A\rtimes_{\varphi}C, where φ\varphi is given by conjugacy, and the extension is called split.

Corollary 10.2.

Suppose that for all P¯\bar{P} as above, the extension (10.3) is set-theoretically split or the group H0(S,CP¯)H^{0}(S,{}^{\bar{P}}\!C) is trivial. Then we have a disjoint union

H1(S,B)=H1(S,AP),H^{1}(S,B)=\bigcupop\displaylimits H^{1}\left(S,{}^{P}\!A\right),

running over all [P¯][\bar{P}] from the image of H1(S,B)H1(S,C)H^{1}(S,B)\rightarrow H^{1}(S,C). The latter map is actually surjective provided that the extension (10.2) is split.

Proof.

Set C=CP¯C^{\prime}={}^{\bar{P}}\!C and A=APA^{\prime}={}^{P}\!A. If B=BPB^{\prime}={}^{P}\!B is set-theoretically split, all fibers over cH0(S,C)c^{\prime}\in H^{0}(S,C^{\prime}) are trivial torsors, and hence the action of H0(S,C)H^{0}(S,C^{\prime}) on H1(S,A)H^{1}(S,A^{\prime}) is trivial. The same holds, of course, if the group H0(S,C)H^{0}(S,C^{\prime}) itself is trivial. So the theorem implies the first assertion.

If the projection pr:BC\operatorname{pr}\colon B\rightarrow C admits a section ss that respects the group structure, the induced map on cohomology is right inverse to H1(S,B)H1(S,C)H^{1}(S,B)\rightarrow H^{1}(S,C), so the latter is surjective. ∎

In particular, for B=ACB=A\rtimes C satisfying the assumptions of the corollary, we get a disjoint union

H1(S,B)=H1(S,AP)H^{1}(S,B)=\bigcupop\displaylimits H^{1}\left(S,{}^{P}\!A\right)

running over all [P¯]H1(S,C)[\bar{P}]\in H^{1}(S,C). It is convenient to regard its elements as “pairs” (γ,α)(\gamma,\alpha) with γ=[P]H1(S,C)\gamma=[P]\in H^{1}(S,C) and αH1(S,AP)\alpha\in H^{1}(S,{}^{P}\!A). If H1(S,AutA/S)H^{1}(S,\operatorname{Aut}_{A/S}) is a singleton, the choice of isomorphisms h:APAh\colon{}^{P}\!A\rightarrow A indeed identifies the above with the product H1(S,C)×H1(S,A)H^{1}(S,C)\times H^{1}(S,A), independently of the hh. Note that this carries a canonical group structure if AA and CC are commutative, which may happen without BB being commutative.

11.  Description of the set of twisted forms

Let KK be a ground field of characteristic p>0p>0, and set S=Spec(K)S=\operatorname{Spec}(K). Using the general results of the previous section, we seek to compute the first non-abelian cohomology for the iterated semidirect products GaUnGm=AutXp,n/K\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}=\operatorname{Aut}_{X_{p,n}/K} and thereby the set of isomorphism classes of twisted forms for Xp,nX_{p,n}. We are able to do so for 1n21\leq n\leq 2.

Throughout, we work over the site 𝒞=(Aff/S)\mathcal{C}=(\text{\rm Aff}/S), endowed with the fppf topology. Let us start with some general useful facts.

Proposition 11.1.

The following hold for group schemes GG of finite type:

  1. (i)

    If  G=N×~GmG=N\,\widetilde{\times}\,\mathbb{G}_{m} for some group scheme NN of finite type, then the canonical map H1(S,N)H1(S,G)H^{1}(S,N)\rightarrow H^{1}(S,G) is bijective.

  2. (ii)

    In the special case G=Gar×~GmG=\mathbb{G}_{a}^{\oplus r}\,\widetilde{\times}\,\mathbb{G}_{m}, the set H1(S,G)H^{1}(S,G) is a singleton.

  3. (iii)

    If  GG is infinitesimal, the group H0(S,G)H^{0}(S,G) is trivial.

Proof.

We have H1(S,Gm)=0H^{1}(S,\mathbb{G}_{m})=0 by Hilbert 90, so (i) follows from Corollary 10.2. With N=GarN=\mathbb{G}_{a}^{\oplus r} we use [Mil80, Theorem III.3.7] and Serre’s vanishing theorem to get H1(S,N)=0H^{1}(S,N)=0 and (ii) follows as well. Assertion (iii) is obvious because an infinitesimal group scheme has Gred=SG_{\operatorname{red}}=S. ∎

Note that for p5p\geq 5, the automorphism group scheme for the rational cuspidal curve X=SpecK[T2,T3]SpecK[T1]X=\operatorname{Spec}K[T^{2},T^{3}]\cup\operatorname{Spec}K[T^{-1}] is given by G=GaGmG=\mathbb{G}_{a}\rtimes\mathbb{G}_{m}, so this curve has no twisted forms besides itself. This purely cohomological argument shows again that quasielliptic curves are confined to characteristic p3p\leq 3. The following observation will also be useful.

Lemma 11.2.

Let G1G_{1} and G2G_{2} be twisted forms of  Ga\mathbb{G}_{a}. If they are isomorphic as schemes, they are isomorphic as group schemes.

Proof.

Let f:G1G2f\colon G_{1}\rightarrow G_{2} be an isomorphism of schemes. Composing with a translation, we may assume f(e1)=e2f(e_{1})=e_{2}. To verify that ff respects the group law, we may assume K=KalgK=K^{\text{\rm alg}}, and this reduces to the case G1=G2=GaG_{1}=G_{2}=\mathbb{G}_{a}. The induced map φ:K[T]K[T]\varphi\colon K[T]\rightarrow K[T] on the coordinate ring is given by φ(T)=λT+μ\varphi(T)=\lambda T+\mu for some λ,μK\lambda,\mu\in K. We have λ0\lambda\neq 0 because ff is non-constant, and μ=0\mu=0 because ff respects the origin. So for each αGa(R)\alpha\in\mathbb{G}_{a}(R), we have f(α)=λαf(\alpha)=\lambda\alpha, which respects the group laws. ∎

For each pair ,K[u]\Phi,\Psi\in K[u] of additive polynomials with gcd(,)0\gcd(\Phi,\Psi)\neq 0 (in other words, not both polynomials vanish), the resulting homomorphism (,):Ga2Ga(\Phi,-\Psi)\colon\mathbb{G}_{a}^{\oplus 2}\rightarrow\mathbb{G}_{a} given by (u,v)(u)(v)(u,v)\mapsto\Phi(u)-\Psi(v) is an epimorphism. The short exact sequence

(11.1) 0U,Ga2(,)Ga00\longrightarrow U_{\Phi,\Psi}\longrightarrow\mathbb{G}_{a}^{\oplus 2}\xrightarrow{(\Phi,-\Psi)}\mathbb{G}_{a}\longrightarrow 0

defines a unipotent group scheme U,U_{\Phi,\Psi}, and the resulting long exact sequence yields

(11.2) H1(S,U,)=K/{(u)(v)u,vK}.H^{1}(S,U_{\Phi,\Psi})=K/\left\{\Phi(u)-\Psi(v)\mid u,v\in K\right\}.

By Russell’s theorem [Rus70, Theorem 2.1], every twisted form of the additive group is isomorphic to U,U_{\Phi,\Psi} with (u)=upn\Phi(u)=u^{p^{n}} for some n0n\geq 0, and (u)\Psi(u) separable. The following sheds further light on this.

Proposition 11.3.

The unipotent group scheme U,U_{\Phi,\Psi} is a twisted form of  Ga\mathbb{G}_{a} if and only if gcd(,)=u\gcd(\Phi,\Psi)=u inside the euclidean domain K[u]K[u].

Proof.

It suffices to treat the case that KK is algebraically closed. Set U=U,U=U_{\Phi,\Psi}. First suppose that there are a,bk[u]a,b\in k[u] with ab=ua\Phi-b\Psi=u. These yield a section for (,):Ga2Ga(\Phi,-\Psi)\colon\mathbb{G}_{a}^{\oplus 2}\rightarrow\mathbb{G}_{a}, defined via u(a(u),b(u))u\mapsto(a(u),b(u)), which does not have to preserve the group laws. It induces an identification Ga2=U×Ga\mathbb{G}_{a}^{\oplus 2}=U\times\mathbb{G}_{a} of schemes. In turn, the coordinate ring A=(U,𝒪U)A=\Gamma(U,\mathscr{O}_{U}) has the property A[y]=K[x,y]A[y]=K[x,y]. According to Zariski cancellation (see [AHE72, Corollary 2.8]), the underlying scheme is isomorphic to the affine line A1\mathbb{A}^{1}. By Lazard’s theorem (see [DG70, Section IV.4, Theorem 4.1]), we must have UGaU\simeq\mathbb{G}_{a} as group schemes.

Conversely, suppose gcd(,)u\gcd(\Phi,\Psi)\neq u. Since KK is algebraically closed, we have (u)=ωA(uω)pm\Phi(u)=\prodop\displaylimits_{\omega\in A}(u-\omega)^{p^{m}} and (u)=ωB(uω)pn\Psi(u)=\prodop\displaylimits_{\omega\in B}(u-\omega)^{p^{n}} for some exponents m,n0m,n\geq 0 and some finite subgroups A,BKA,B\subset K. Their intersection is non-zero, by the assumption on the gcd. Consequently, we can write (u)=(h(u))1\Phi(u)={}_{1}(h(u)) and (u)=(h(u))1\Psi(u)={}_{1}(h(u)) for some additive polynomial of the form h(u)=i=0p1(uiω0)h(u)=\prodop\displaylimits_{i=0}^{p-1}(u-i\omega_{0}), with some non-zero ω0K\omega_{0}\in K. In turn, the projection (,):Ga2Ga(\Phi,-\Psi)\colon\mathbb{G}_{a}^{\oplus 2}\rightarrow\mathbb{G}_{a} factors over the morphism h:GaGah\colon\mathbb{G}_{a}\rightarrow\mathbb{G}_{a}. So the kernel U,U_{\Phi,\Psi} is disconnected or non-reduced. ∎

Proposition 11.4.

For each group scheme of the form G=GaQGmG=\mathbb{G}_{a}\rtimes Q\rtimes\mathbb{G}_{m}, where QQ is any infinitesimal group scheme of finite type, we have a canonical identification

H1(S,G)=αH1(S,Q)K/{(u)α(v)αu,vK}H^{1}(S,G)=\bigcupop\displaylimits_{\alpha\in H^{1}(S,Q)}K/\{{}_{\alpha}(u)-{}_{\alpha}(v)\mid u,v\in K\}

for certain additive polynomials ,ααK[u]{}_{\alpha},{}_{\alpha}\in K[u] with gcd(,α)α=u\gcd({}_{\alpha},{}_{\alpha})=u.

Proof.

By Proposition 11.1, the canonical map H1(S,GaQ)H1(S,G)H^{1}(S,\mathbb{G}_{a}\rtimes Q)\rightarrow H^{1}(S,G) is bijective. Since QQ is infinitesimal, we can apply Proposition 10.2, and the assertion follows with (11.2). ∎

Roughly speaking, to understand this cohomology of GG, one has to understand the cohomology of QQ and the dependence of the additive polynomials ,αα{}_{\alpha},{}_{\alpha} on the class α\alpha. We now seek to unravel this with G=GaUnGmG=\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m}. For n=1n=1, the term in the middle becomes U1=αpU_{1}=\alpha_{p}. The short exact sequence 0αpnGaFnGa00\rightarrow\alpha_{p^{n}}\rightarrow\mathbb{G}_{a}\stackrel{{\scriptstyle F^{n}}}{{\rightarrow}}\mathbb{G}_{a}\rightarrow 0 yields an identification H1(S,αpn)=K/KpnH^{1}(S,\alpha_{p^{n}})=K/K^{p^{n}} for every n1n\geq 1. The dependence on the additive polynomials can be described as follows.

Proposition 11.5.

For the αp\alpha_{p}-torsor P=SpecK[y]/(ypα)P=\operatorname{Spec}K[y]\,/(y^{p}-\alpha), the additive polynomials (u)=up\Phi(u)=u^{p} and (v)=vαvp\Psi(v)=v-\alpha v^{p} give the twisted form GaP=U,{}^{P}\!\mathbb{G}_{a}=U_{\Phi,\Psi}.

Proof.

Set B=K[x,y]/(ypα)B=K[x,y]\,/(y^{p}-\alpha). By definition, the coordinate ring for GaP{}^{P}\!\mathbb{G}_{a} is the subring ABA\subset B of elements that are invariant under xx+λxpx\mapsto x+\lambda x^{p} and yy+λy\mapsto y+\lambda, for all group elements λαp(R)\lambda\in\alpha_{p}(R) and all rings RR. Clearly, v=xpv=x^{p} and u=xxpyu=x-x^{p}y are such invariants, satisfying the relation up=vαvpu^{p}=v-\alpha v^{p}. Its partial derivatives generate the unit ideal, and we conclude that the subring AAA^{\prime}\subset A generated by uu and vv is regular and one-dimensional and can be identified with the residue class ring K[u,v]/(upv+αvp)K[u,v]\,/(u^{p}-v+\alpha v^{p}). The composite extension K(xp)Frac(A)Frac(A)Frac(B)K(x^{p})\subset\operatorname{Frac}(A^{\prime})\subset\operatorname{Frac}(A)\subset\operatorname{Frac}(B) has degree p2p^{2}, and the outer steps have degree pp. Consequently, A=AA^{\prime}=A. The assertion now follows from Lemma 11.2. ∎

To tackle the case n=2n=2, we use the short exact sequence

0αpU2αp20,0\longrightarrow\alpha_{p}\longrightarrow U_{2}\longrightarrow\alpha_{p^{2}}\longrightarrow 0,

where the inclusion on the left is λ1+λxp2\lambda\mapsto 1+\lambda x^{p^{2}} and the surjection on the right (1+λ1xp+λ2xp2)λ1(1+\lambda_{1}x^{p}+\lambda_{2}x^{p^{2}})\mapsto\lambda_{1}. Given α,βK\alpha,\beta\in K, the finite scheme P=Pα,βP=P_{\alpha,\beta} defined by

(11.3) P(R)={(y,z)R2yp2=α and zp=β+αyp}P(R)=\left\{(y,z)\in R^{2}\mid\text{$y^{p^{2}}=\alpha$ and $z^{p}=\beta+\alpha y^{p}$}\right\}

carries a U2U_{2}-action via the formula (λ1,λ2)(y,z)=(λ1+y,λ2+z+λ1yp)(\lambda_{1},\lambda_{2})\cdot(y,z)=(\lambda_{1}+y,\lambda_{2}+z+\lambda_{1}y^{p}). One easily checks that this indeed takes values in P(R)P(R), that it satisfies the axioms for group actions, and that the action is free and transitive. The induced αp2\alpha_{p^{2}}-torsor P¯\bar{P} is obtained from PP as a quotient by αp\alpha_{p}, in other words, by the action of λ2\lambda_{2}. This yields P¯(R)={yRyp2=α}\bar{P}(R)=\{y\in R\mid y^{p^{2}}=\alpha\}. In turn, we get the description H1(S,U2)=K/Kp2K/KpH^{1}(S,U_{2})=\bigcupop\displaylimits_{K/K^{p^{2}}}K/K^{p}. It remains to express the twisted form GaP{}^{P}\!\mathbb{G}_{a} in terms of additive polynomials.

Proposition 11.6.

For the U2U_{2}-torsor PP as above, the additive polynomials

(u)=up2and(u)=uαupβpup2\Phi(u)=u^{p^{2}}\quad\text{and}\quad\Psi(u)=u-\alpha u^{p}-\beta^{p}u^{p^{2}}

give the twisted form GaP=U,{}^{P}\!\mathbb{G}_{a}=U_{\Phi,\Psi}.

Proof.

Set B=K[x,y,z]/(yp2α,zpβαyp)B=K[x,y,z]\,/(y^{p^{2}}-\alpha,z^{p}-\beta-\alpha y^{p}). The coordinate ring for GaP{}^{P}\!\mathbb{G}_{a} is the subring ABA\subset B of elements that are invariant under

xx+λ1xp+λ2xp2,yy+λ1,andzz+λ2+λ1ypx\longmapsto x+\lambda_{1}x^{p}+\lambda_{2}x^{p^{2}},\quad y\longmapsto y+\lambda_{1},\quad\text{and}\quad z\longmapsto z+\lambda_{2}+\lambda_{1}y^{p}

for all group elements (λ1,λ2)U2(R)(\lambda_{1},\lambda_{2})\in U_{2}(R) and all rings RR. One easily checks that v=xp2v=x^{p^{2}} and u=xxpyxp2z+xp2yp+1u=x-x^{p}y-x^{p^{2}}z+x^{p^{2}}y^{p+1} are invariant and that these invariants satisfy the relation up2=vαvpβpvp2u^{p^{2}}=v-\alpha v^{p}-\beta^{p}v^{p^{2}}. The argument concludes as in the preceding proof. ∎

Note that the invariant uu can be found by starting with the non-invariant xx and successively adding monomials to cancel non-invariance. Collecting all the above, we have determined the non-abelian cohomology for Gn=GaUnGmG_{n}=\mathbb{G}_{a}\rtimes U_{n}\rtimes\mathbb{G}_{m} in the cases 1n21\leq n\leq 2.

Theorem 11.7.

With the above notation, we have

H1(S,G1)=αK/{upv+αvpu,vK},H^{1}(S,G_{1})=\bigcupop\displaylimits_{\alpha}K/\left\{u^{p}-v+\alpha v^{p}\mid u,v\in K\right\},

where the union runs over all αK/Kp\alpha\in K/K^{p}, and

H1(S,G2)=(α,β)K/{up2vαvpβpvp2u,vK},H^{1}(S,G_{2})=\bigcupop\displaylimits_{(\alpha,\beta)}K/\left\{u^{p^{2}}-v-\alpha v^{p}-\beta^{p}v^{p^{2}}\mid u,v\in K\right\},

where the union runs over (α,β)K/Kp2K/Kp(\alpha,\beta)\in\bigcupop\displaylimits_{K/K^{p^{2}}}K/K^{p} with αK/Kp2\alpha\in K/K^{p^{2}} and βK/Kp\beta\in K/K^{p}.

Note that for 3pn43\leq p^{n}\leq 4, this gives back, in an intrinsic fashion, Queen’s descriptions for quasielliptic curves (cf. [Que71, Que72]).

Also note that for n=1n=1, the group K/{upv+αvpu,vK}K/\{u^{p}-v+\alpha v^{p}\mid u,v\in K\} is trivial, provided that KK is separably closed or αKp\alpha\in K^{p}. It follows that the Frobenius map

H1(S,G1)H1(S,G1),PP(p)=PKFH^{1}(S,G_{1})\longrightarrow H^{1}(S,G_{1}),\quad P\longmapsto P^{(p)}=P\otimes{}_{F}K

is trivial, in the sense that it sends every class to the distinguished class. In particular, every twisted form YY of Xp,1X_{p,1} is untwisted by Frobenius pullback and becomes a rational curve (compare with [HS22]). Likewise, for n=2n=2, the group K/{up2vαvpβpvp2u,vK}K/\{u^{p^{2}}-v-\alpha v^{p}-\beta^{p}v^{p^{2}}\mid u,v\in K\} is trivial if KK is separably closed or αKp2\alpha\in K^{p^{2}}, βKp\beta\in K^{p}. Now the map PP(p2)P\mapsto P^{(p^{2})} is trivial, and every twisted form YY of Xp,2X_{p,2} gets untwisted by the second Frobenius pullback.

With the notation from the theorem, let TT be a GnG_{n}-torsor and αK/Kpn\alpha\in K/K^{p^{n}} be the ensuing class, for 1n21\leq n\leq 2. Write X~\tilde{X} for the twisted form of X=Xp,nX=X_{p,n} corresponding to TT.

Proposition 11.8.

With the above notation, the curve X~\tilde{X} is regular provided that αK/Kpn\alpha\in K/K^{p^{n}} does not belong to Kp/KpnK^{p}/K^{p^{n}}.

Proof.

The GnG_{n}-torsor TT is induced from some torsor PP with respect to GaUn\mathbb{G}_{a}\rtimes U_{n}, according to Proposition 11.1. By construction, the class αK/Kpn\alpha\in K/K^{p^{n}} corresponds to the quotient P¯=(GaUn1)\P\bar{P}=(\mathbb{G}_{a}\rtimes U_{n-1})\backslash P, and the latter has coordinate ring K[T]/(Tpnα)K[T]\,/(T^{p^{n}}-\alpha), where we also write α\alpha for the scalar rather than the class. The coordinate ring is reduced, in light of our assumption. According to Theorem 9.4, the curve X~\tilde{X} is regular. ∎

It should be possible to extend the above results to all n1n\geq 1. For this, one has to find an inductive description for the UnU_{n}-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. pp, III, Invent. Math. 35 (1976), 197–232.
  • [BM77] by same author, Enriques’ classification of surfaces in char. pp, 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.