The Scheme of Monogenic Generators II:
Local Monogenicity and Twists
Abstract.
This is the second paper in a series of two studying monogenicity of number rings from a moduli-theoretic perspective. By the results of the first paper in this series, a choice of a generator for an -algebra is a point of the scheme . In this paper, we study and relate several notions of local monogenicity that emerge from this perspective. We first consider the conditions under which the extension admits monogenerators locally in the Zariski and finer topologies, recovering a theorem of Pleasants as a special case. We next consider the case in which is étale, where the local structure of étale maps allows us to construct a universal monogenicity space and relate it to an unordered configuration space. Finally, we consider when admits local monogenerators that differ only by the action of some group (usually or ), giving rise to a notion of twisted monogenerators. In particular, we show a number ring has class number one if and only if each twisted monogenerator is in fact a global monogenerator .
2020 Mathematics Subject Classification:
Primary 14D20, Secondary 11R04, 13E151. Introduction
We begin by recalling the essential points of the previous paper of this series[ABHS]. Given an extension of commutative rings with identity (henceforth, rings) , we say that that is monogenic over if there is an element so that . Such an element is called a monogenerator. Similarly, is said to be -genic over if there exists a tuple so that . Such a tuple is a generating -tuple. We are motivated by the case of an extension of number rings .
Such extensions of number rings are finite locally free over a Noetherian base. In fact all we need for our results are maps of schemes that are Zariski locally of this form. We gather these hypotheses into a common “Situation” for convenience.
Situation 1.1.
Let be a finite locally free morphism of schemes of constant degree with locally noetherian, and let be a quasiprojective morphism (almost always or ).
In the preceding paper we prove the following representability result, which implies in particular that if is finite locally free over a Noetherian ring , then there is a scheme that represents the monogenerators for over .
Theorem 1.2 ([ABHS, Proposition 2.3, Corollary 3.8]).
Let be as in Situation 1.1. Then
-
(1)
There exists a smooth, quasiaffine -scheme representing the contravariant functor on -schemes
-
(2)
If , then is an affine -scheme.
We write for the case in which . We call the scheme the scheme of -generators. If , we write instead of and call it the scheme of monogenerators or monogenicity space. If and are affine, we may write or instead.
In the case that , , and , standard universal properties imply that the -points of are in natural bijection with the generating -tuples of over . If we assume further that , we find that the -points of are in bijection with generating -tuples for over .
By analogy with the affine case, we therefore say that the -points of are generating -tuples and the -points of are monogenerators for . Such a morphism is monogenic if a monogenerator exists.
1.1. Equations in local coordinates
The scheme has a simple description in local coordinates on which we recall so that we may use it in computations. We start by noticing that is naturally a subscheme of another moduli scheme, the Weil Restriction.
Definition 1.3.
Let be as in Situation 1.1. The Weil Restriction of to , denoted , is the scheme (unique up to isomorphism) which represents the contravariant functor
on -schemes.
We abbreviate in a parallel fashion to . It is proven in [JLM+17, Theorem 1.3, Proposition 2.10] that the Weil Restriction exists and is a quasiprojective -scheme. We prove in [ABHS, Proposition 2.3] that the natural map is a quasi-compact open immersion.
In the case that is a finite free Noetherian -algebra with -basis , things are simpler. It is easy to check that
via the isomorphism sending to the unique map over sending to .
Definition 1.4.
Suppose is a finite free Noetherian -algebra with -basis . Let for be the unique elements of so that we have
in the ring . We call the matrix the matrix of coefficients with respect to the basis . Its determinant is the local index form with respect to the basis .
Theorem 1.5.
([ABHS, Theorem 3.1]) With notation as above, is the distinguished open subscheme of cut out by the non-vanishing of the local index form. In particular,
Additionally, we recall from [ABHS] that the local index forms give the complement of in a closed subscheme structure:
Definition 1.6 (Non-monogenerators ).
Let be the ideal sheaf on generated locally by local index forms. We call this the index form ideal. Let be the closed subscheme of cut out by the vanishing of . We call this the scheme of non-monogenerators, since its support is the complement of inside of .
Since is a scheme, it is a sheaf in the fpqc topology on . This invites a local study of monogenicity111‘Monogeneity’ is also common in the literature. of , the subject of this paper.
1.2. Results
We identify and relate several “local” notions of monogenicity. To guide the reader, their relationships are indicated in Figure 1.1.
A sheaf theoretic notion of local monogenicity immediately presents itself.222We remark that Zariski local monogenicity is equivalent to the condition of “local monogenicity” considered by Bhargava et. al. in [ABS20] but strictly weaker than their condition of “no local obstruction to monogenicity.”
Definition 1.7.
Let be a subcanonical Grothendieck topology on schemes, for example the Zariski, Nisnevich, étale, fppf, or fpqc topologies. We say that is -locally -genic if the sheaf is locally non-empty in the topology . I.e., there is a -cover of such that is non-empty for all .
The notions of -local monogenicity are considered in Section 2, and we find that these reduce to just two notions of “local monogenicity.”
Theorem 1.8.
Let be as in Situation 1.1.
-
(1)
(Theorem 2.1) The following are equivalent:
-
(a)
is locally monogenic in the Zariski topology;
-
(b)
is “monogenic at completions,” i.e. for all points of , we have that is monogenic;
-
(c)
is “monogenic at points,” i.e. for each point of with residue field , we have that is monogenic.
-
(a)
-
(2)
(Theorem 2.10) The following are equivalent:
-
(a)
is locally monogenic in the étale topology;
-
(b)
is locally monogenic in the fpqc topology;
-
(c)
is “monogenic at geometric points,” i.e. for all algebraically closed fields and maps , we have that is monogenic.
-
(a)
We then use the structure of finite algebras over fields to classify monogenicity at points. In particular, we recover Pleasants’s characterization [Ple74, Theorems 1 and 2] of monogenicity at completions as a corollary.
Theorem 1.9.
Let be induced by where is a field and is a local Artinian -algebra with residue field and maximal ideal . Then is monogenic if and only if
-
(1)
is monogenic;
-
(2)
; and
-
(3)
If and is inseparable, then
is a non-split extension.
Theorem 1.10.
Suppose is induced by where is a field and is an Artinian -algebra. Write where the are local artinian -algebras with respective residue fields . Then is monogenic if and only if
-
(1)
is monogenic for each ;
-
(2)
for each finite extension of , has fewer points with residue field isomorphic to than .
In Section 3, we consider monogenicity spaces of étale , a salient case since extensions of number rings are generically étale and such étale maps share a common local structure: Finite étale maps are étale-locally isomorphic to the trivial -sheeted cover . The latter has equal to the configuration space of distinct points in . Therefore we may interpret monogenicity spaces as twisted generalizations of configuration spaces, at least when is étale.
Using the fact[Sta20, 04HN] that every finite étale map of degree is pulled back from the morphism of stacks , we then construct a monogenicity space .
Theorem 1.11.
There are isomorphisms
in which the action by is in each case the appropriate restriction of the permutation action on coordinates of , and denotes the “fat diagonal” of , the locus where some pair of coordinates coincide.
In particular, the -points of coincide with the points of the unordered configuration space of points in . This monogenicity space is universal for étale maps in the sense that the monogenicity space of each étale is pulled back from .
We then consider several examples enabled by the structure of in the étale case, among them a connection to braid groups, a construction of the moduli space of genus zero pointed curves from monogenicity spaces, the monogenicity space of a -torsor, and the monogenicity space of an isogeny of elliptic curves.
We remark that the monogenicity space appears to be a scheme theoretic enhancement of the universal spaces of [Han80, Section 3] and [GL69, Section 1] in the category of topological spaces. Meanwhile, the universal monogenicity space for -torsors appears to be a scheme theoretic enhancement of the space considered in [DH84, Theorem 1.3].
A local-to-global sequence is missing for -local monogenicity. Yet there are natural group actions on , in particular by and . In Section 4 we study “twisted” versions of monogenicity in which has local monogenerators that differ from each other by the action of such groups. More precisely:
Definition 1.12 (Twisted Monogenerators).
A ()-twisted monogenerator for is:
-
(1)
a Zariski open cover for elements ,
-
(2)
a system of “local” monogenerators for over , and
-
(3)
units
such that
-
•
for all , we have
-
•
for all , the “cocycle condition” holds:
Two such systems , are equivalent if, after passing to a common refinement of their respective covers, there is a global unit such that and . Likewise is -twisted monogenic if there is a cover with ’s as above, but with the units in item (3) replaced by pairs such that each is a unit and .
Under certain hypotheses, we show:
-
Proposition 4.14:
is -twisted monogenic if and only if it is -twisted monogenic.
-
Theorem 4.19:
The class number of a number ring is one if and only if all twisted monogenic extensions of number rings are in fact monogenic.
-
Remark 4.12:
There is a local-to-global sequence relating affine equivalence classes of monogenerators with global monogenerators as above.
-
Theorem 4.2:
There are moduli spaces of and -twisted monogenerators analogous to .
-
Theorem 4.17:
There are finitely many twisted monogenerators up to equivalence.
We remark that a -twisted monogenerator is equivalent to an embedding over into a line bundle on . Such embeddings into line bundles were considered for topological spaces in [Lø80].
Section 5 concludes with ample examples of the scheme of monogenerators and the various interactions between the forms of local monogenicity.
To avoid repetition, we invite the reader to consult the first paper in this series for a more detailed survey of the relevant literature.
1.3. Acknowledgements
The second author would like to thank David Smyth and Ari Shnidman for their support and for helpful conversations.
The third author would like to thank Gebhard Martin, the mathoverflow community for [Herb] and [Hera], Tommaso de Fernex, Robert Hines, and Sam Molcho. Tommaso de Fernex looked over a draft and made helpful suggestions about jet spaces. The third author thanks the NSF for providing partial support by the RTG grant #1840190.
The fourth author would like to thank Henri Johnston and Tommy Hofmann for help with computing a particularly devious relative integral basis in Magma.
All four authors would like to thank their graduate advisors Katherine E. Stange (first and fourth authors) and Jonathan Wise (second and third authors). This project grew out of the fourth author trying to explain his thesis to the second author in geometric terms. We are greatful to Richard Hain for making us aware of the literature on monogenicity for Stein spaces and topological spaces. For numerous computations throughout, we were very happy to be able to employ Magma [BCP97] and SageMath [The19]. For a number of examples the [LMF21] was invaluable. Finally, we are thankful for Lily.
2. Local monogenicity
2.1. Zariski-local monogenicity
This section shows Zariski-local monogenicity can be detected over points and completions as spelled out in Remark 2.3. We will make frequent use of the vocabulary and notation of [ABHS, Section 3].
Theorem 2.1.
The following are equivalent:
-
(1)
is Zariski-locally monogenic.
-
(2)
There exists a family of maps such that
-
(a)
the are jointly surjective;
-
(b)
for each point , there is an index and point so that induces an isomorphism ;
-
(c)
is monogenic for all .
-
(a)
-
(3)
is monogenic over points, i.e., is monogenic for each point .
Proof.
(2)(3): Suppose such a cover is given. For each , let be the section through . monogenicity is preserved by pullback on the base, so pulling back along implies (3).
(3)(1): Let be a point with residue field and let be a monogenerator for . We claim that extends to a monogenerator over an open subset containing , from which (1) follows. The claim is Zariski local, so assume and is globally free. The Weil Restriction is then isomorphic to affine space .
We first extend to a section of . The monogenerator entails a point , i.e. elements . Choose arbitrary lifts of . The elements must have a common denominator, so we have for some . Thus our point extends to for some distinguished open neighborhood containing .
Finally, we restrict to a section of . The monogenicity space is an open subscheme of the Weil Restriction , so restricts to a monogenerator where . By hypothesis, , so is the desired extension of . ∎
Remark 2.2.
The same proof shows that is Zariski-locally -genic if and only if its fibers are -genic.
Remark 2.3.
Item (2) of Theorem 2.1 implies the following are also equivalent to Zariski-local monogenicity:
-
(1)
is “monogenic at local rings,” i.e, for each point of , is monogenic.
-
(2)
is “monogenic at completions,” i.e., for each point of , is monogenic, where denotes the completion of with respect to its maximal ideal.
-
(3)
is locally monogenic in the Nisnevich topology as in Definition 1.7.
Corollary 2.4 ([Ser79, Proposition III.6.12]).
Let be an extension of local rings inducing a separable extension of residue fields. Then is monogenic over .
We now recall some ideas in order to compare with related results in the number ring case.
Definition 2.5.
Let be an extension of number rings. Given generating , we write for the index . A non-zero prime of is a common index divisor333Common index divisors are also called essential discriminant divisors and inessential or nonessential discriminant divisors. The shortcomings of the English nomenclature likely come from what Neukirch [Neu99, page 207] calls “the untranslatable German catch phrase […] ‘außerwesentliche Diskriminantenteile.’” Our nomenclature is closer to Fricke’s ‘ständiger Indexteiler.’ for the extension if for every generating .
Common index divisors are exactly the primes whose splitting in cannot be mirrored by irreducible polynomials in ; see [Hen94] and [Ple74].
We recall[ABHS, Remark 3.11] that if is an open affine subscheme of on which is free with basis then is a local index form on .
Restating the property of being monogenic at points in terms of the index form, we obtain a generalization of the notion of having no common index divisors:
Proposition 2.6.
is monogenic over points if and only if for each point of and local index form around , there is a tuple such that is nonzero in .
Proof.
is monogenic precisely when has a point. ∎
Immediately we recover an explicit corollary validating the generalization:
Corollary 2.7.
Suppose is induced by an extension of number rings . Then is Zariski-locally monogenic if and only if there are no common index divisors for .
Example 2.8.
There are extensions of number rings that are locally monogenic but not monogenic.
In [ABS20], Alpöge, Bhargava, and Shnidman say that an extension has no local obstruction to monogenicity if a local index form represents over for all primes or over for all primes . This is a stronger condition than Zariski-local monogenicity, and they show in [ABS20] and [ABS21] that a positive proportion of quartic and cubic fields are not monogenic despite having no local obstruction to monogenicity.
Narkiewicz [Nar04, Page 65] gives the following concrete example of a locally monogenic but not monogenic extension. Let with , square-free, , and . The number ring is not monogenic over despite having no common index divisors. We consider the case in Example 5.3. This also gives an example of an extension which is Zariski locally monogenic but which has a local obstruction to monogenicity.
2.2. Monogenicity over geometric points
Definition 2.9.
Say that over is monogenic over geometric points if, for each morphism where is an algebraically closed field, is monogenic.
While it is a weaker condition than monogenicity over points in general, it is equivalent to some conditions that might seem more natural.
Theorem 2.10.
The following are equivalent:
-
(1)
the local index forms for are nonzero on each fiber of ;
-
(2)
for each point with residue field , there is a finite Galois extension such that is monogenic, and this extension may be chosen to be trivial if is an infinite field;
-
(3)
is monogenic over geometric points;
-
(4)
there is a jointly surjective collection of maps so that is monogenic for each ;
-
(5)
is surjective;
-
(6)
is étale-, smooth-, fppf-, or fpqc-locally monogenic.
If, in addition, is a Cartier divisor in (i.e., the local index forms are non-zero divisors), the above are also equivalent to:
-
(7)
(Definition 1.6) is flat.
To see some of the subtleties one can compare item (1) above with Lemma 2.6 and item (4) above with item (2) of Theorem 2.1.
Proof.
The assertions are Zariski-local, so we may choose local coordinates as usual (, , coordinates ).
(1) (2): Suppose first that is a point with an infinite field. Let be the corresponding prime of and write for the restriction of the local index form modulo . Recall that since is infinite, polynomials in are determined by their values on . Since is nonzero, must be nonzero for some . This shows that is monogenic, so we have (2).
Next suppose is a finite field. Then, since is nonzero, there is a finite field extension (necessarily Galois) of such that there exists with . This shows that is monogenic, so we have (2) again.
(2) (3): Let be an algebraically closed field and a map. Let be the image of , and let be the field extension given by (2). Then pullback along implies that is monogenic.
(4) (1): For each point , choose an index and a point mapping to . Let be an index form around . By pullback, is monogenic, so pulls back to a nonzero function over . Therefore is nonzero over as well.
(5) (6): Note that is smooth, since is the composite of an open immersion and an affine bundle. Moreover, the identity function on by definition yields a monogenerator for . Therefore, is smooth-locally monogenic. Since the smooth topology is equivalent to the étale topology, there is an étale cover factoring through . Since is already monogenic, is also monogenic. This étale cover is also a cover in the fppf and fpqc topologies.
Recall that an -module is flat if and only if for each prime of and ideal of lying over , is flat over . Therefore, by the local criterion for flatness[Sta20, 00MK], is flat over if and only if
for all such ideals and . Therefore, is flat if and only if
is injective for all and as above. All of these maps are injective if and only if the maps of -modules
are all injective as varies over the prime ideals of . Since is an integral domain for each , injectivity fails if and only if reduces to in the fiber over some . We conclude that (1) holds if and only if (7) holds. ∎
Corollary 2.11.
If all of the points of have infinite residue fields, then the following are equivalent:
-
(1)
is monogenic over geometric points;
-
(2)
is Zariski-locally monogenic.
Remark 2.12.
The conclusion of Corollary 2.11 fails dramatically if has finite residue fields. For coming from an extension of number rings condition (1) always holds (see Corollary 2.16 below), yet there are extensions that are not locally monogenic. In this sense, monogenicity is more subtle in the arithmetic context than the geometric one. For an example of an extension that is monogenic over geometric points but is not monogenic over points see Example 5.1.
2.3. Monogenicity over points
In light of Theorem 2.1 and Theorem 2.10, it is particularly interesting to characterize monogenicity of in the case that is the spectrum of a field . In this case is the spectrum of an -dimensional -algebra . Such algebras are Artinian rings, and a well-known structure theorem implies that is a direct product of local Artinian rings . We will exploit this to give a complete characterization of Zariski-local monogenicity.
The result in the case that both and are spectra of fields is well-known.
Theorem 2.13 (Theorem of the primitive element).
Let be a finite field extension. Then is monogenic if and only if there are finitely many intermediate subfields .
In particular, a finite separable extension of fields is monogenic.
We next consider the monogenicity of when is a nilpotent thickening of , leaving fixed. A key ingredient is a study of the square zero extensions of .
We remark for comparison that the proof below does not consider a nilpotent thickening of the base . In fact, if is a nilpotent closed immersion with , any monogenerator extends to locally in the étale topology. This results from the smoothness of .
Theorem 1.9. Let be induced by where is a field and is a local Artinian -algebra with residue field and maximal ideal . Then is monogenic if and only if
-
(1)
is monogenic;
-
(2)
; and
-
(3)
If and is inseparable, then
is a non-split extension.
Proof.
If the tangent space has dimension greater than 1, then no map can be injective on tangent vectors as is required for a closed immersion.
On the other hand, if the tangent space of has dimension , we have , and the result is true by hypothesis.
Now suppose the tangent space of has dimension . A morphism is a closed immersion if and only if it is universally closed, universally injective, and unramified[Sta20, Tag 04XV].
Choose a closed immersion . Equivalently, write where is the monic minimal polynomial of some element . Since is a universal homeomorphism[Sta20, Tag 054M], any extension of to inherits the properties of being universally injective and universally closed from .
Whether such an extension is ramified can be checked on the level of tangent vectors[Sta20, Tag 0B2G]. It follows that a morphism is a closed immersion if and only if its restriction to the vanishing of is. On the other hand, any map extends to (choose a lift of the image of arbitrarily). Therefore, it suffices to consider the case that is a square zero extension of .
By hypothesis, we have a presentation of as . We conclude with some elementary deformation theory, see for example [Ser06, §1.1]. We have a square zero extension of
By assumption, is also a square zero extension of :
By [Ser06, Proposition 1.1.7], there is a morphism of -algebras inducing the identity on . Since as a module, either restricts to an isomorphism or else the zero map. In the former case, the composite is a surjection, and we are done. In the latter case, is the pushout of the extension along , so is the split extension .
If is separable, then the extension is itself split [Ser06, Proposition B.1, Theorem 1.1.10], i.e. there an isomorphism . Composing with gives the required monogenerator.
If is inseparable and , we will show that is not monogenic. Any generator for over must also be a generator for over the maximal separable subextension of , so we may assume that is purely inseparable. Moreover, any generator for over must reduce modulo to a generator of . Since is purely inseparable, the minimal polynomial of satisfies . Note for some . Since is assumed to be a monogenerator, there is a polynomial such that . Reducing, , so for some . Then
a contradiction. We conclude that in this case is not monogenic. ∎
Remark 2.14.
In the case that is perfect, the first and third conditions hold automatically. If is regular of dimension 1 the second condition is trivial.
Theorem 1.10. Suppose is induced by where is a field and is an Artinian -algebra. Write where the are local artinian -algebras with respective residue fields . Then is monogenic if and only if
-
(1)
is monogenic for each ;
-
(2)
for each finite extension of , has fewer points with residue field isomorphic to than .
Proof.
Note that a map is a closed immersion if and only if each map is a closed immersion and the closed immersions are disjoint: for all . This is equivalent to the statement that is surjective if and only if is surjective for each and whenever , which follows quickly in turn from the Chinese remainder theorem.
The proof of Theorem 1.9 shows that a closed immersion can be chosen with image any of the points of with residue field . Then the condition on numbers of points is exactly what we need for the images of the s not to overlap without running out of points. (Since topologically, the components are single points.) ∎
Remark 2.15.
Condition (2) is trivial in the case that the residue fields of are infinite, highlighting the relative simplicity of monogenicity in the geometric context.
If is instead induced by an extension of number rings, then Remark 2.14 implies condition (1) is trivial. In particular, an extension of has common index divisors if and only if there is “too much prime splitting” in the sense of condition (2). This recovers the theorem of [Hen94] (also see [Ple74, Cor. to Thm. 3]) that is a common index divisor if and only if there are more primes in above of residue class degree than there are irreducible polynomials of degree in for some positive integer
Corollary 2.16.
If is induced by an extension of number rings , then is monogenic over geometric points.
Proof.
Let be a point of . Let be the ring for the fiber of over . Note that is either of characteristic 0 or finite, so is perfect. Decompose into a direct product of local Artinian algebras . Since is perfect, conditions (1) and (3) of Theorem 1.9 hold for . Condition (2) holds as well since is regular of dimension 1. Therefore is monogenic for each .
Now consider the base change of to the algebraic closure of . Write for the ring of functions of this base change, and write as a product of local Artinian algebras . For each we have that is monogenic. Each is a closed subscheme of exactly one of the s, so by composition, is monogenic for each . This gives us condition (1) of Theorem 1.10 for . Since is infinite, condition (2) holds triviallly. We conclude that is monogenic, as required. ∎
3. Étale maps, configuration spaces, and monogenicity
This section concerns maps that are étale, or unramified. Locally, the monogenicity space becomes a configuration space, classifying arrangements of distinct points on a given topological space. Philosophically, therefore regards as a twisted family of points to be configured in . We are led to interpret as an arithmetic refinement of the configuration space of . In Remark 3.5, we see that an action of the absolute Galois group on the étale fundamental group of has been observed in anabelian geometry. We end Subsection 3.2 with a handful of exotic applications in other areas.
All extensions sit somewhere between the étale case and jet spaces (see [ABHS, Example 2.7, 4.3, 4.16]), between being totally unramified and totally ramified. In Section 3.4, we recall a general construction of the discriminant which cuts out the locus of ramification. Specifically, [Poo06, §6] says that our description in the étale case holds precisely away from the vanishing of the discriminant. The discriminant plays a similar role in the classical case when investigating the monogenicity of an extension defined by a polynomial. We end with some remarks on using stacks to promote a ramified cover of curves to an étale cover of stacky curves as in [Cos06].
3.1. The trivial cover
We start with the simplest case of an étale cover: the trivial cover of by several copies of itself. We work rather concretely and revisit the general situation with more sophistication in the next subsection. Write and for the trivial degree finite étale cover.
Example 3.1 (Monogenicity of a trivial cover).
Let and let be the map induced by the identity on each copy of . Given a commutative diagram
one expects that the map will be a closed immersion if and only if for all and . A computation in coordinates will confirm this.
We will use the notation of Definition 1.4 and Theorem 1.5 to compute in local coordinates. Working Zariski locally on , we may assume that , , and that are the standard basis vectors for . Let be the corresponding coordinates for , so that
This isomorphism identifies the -point of with the map whose restriction to the th copy of in is .
Next, observe that in
for all . Therefore, the matrix of coefficients is the Vandermonde matrix with th column given by . The index form is then the well-known Vandermonde determinant:
The index form vanishes therefore on the so-called fat diagonal , given by the union of all loci where two coordinates are equal.
It follows that
the complement in of . This space is otherwise known as the space of ordered configurations of points, .
Slightly more abstract reasoning yields a similar result if is any quasi-projective scheme.
Example 3.2 ( for a trivial cover).
Let be a quasiprojective map, and let be the map induced by the identity on each copy of . Observe that if is an -scheme, we have natural identifications
so we may identify with the -fold fiber product of over .
For each , we can construct a subscheme of consisting of the points whose th and coordinates are equal. We let the fat diagonal be the scheme theoretic union of the subschemes . Since is separated, this fat diagonal is a closed subscheme of .
Observe that any morphism of -schemes in will be proper and unramified as is proper and unramified and is separated. In addition, for each point , the induced field extension is an isomorphism, since the same is true of the map .
3.2. The case of étale
Consider the category of schemes over a final scheme equipped with the étale topology. For example, take or . Write for the symmetric group on letters and for the stack on of étale -torsors.
Regard as the subgroup of of permutations fixing the th letter, and let be the map induced by the inclusion. The isomorphism class of the resulting map of classifying spaces is unchanged if is taken as the subgroup fixing some other letter, since resulting inclusion map only differs from this one by conjugation. The morphism is the universal -sheeted cover in the following sense.
Lemma 3.3 ([Cos06, Lemma 2.2.1], [HW21, Lemma 3.2]).
Let be a positive integer. Let be the fibered category over with:
-
(1)
objects the finite étale morphisms of degree ;
-
(2)
arrows the cartesion diagrams
-
(3)
projection to given by .
Then there is an equivalence of fibered categories given by taking a map to the pullback of along .
Recall that pullback squares of schemes
induce identifications
Reduce thereby to the universal -sheeted finite étale cover , .
Each has an affine line obtained via quotienting by the trivial action.
Theorem 1.11. There are isomorphisms
in which the action by is in each case the appropriate restriction of the permutation action on coordinates of .
Proof.
We observe that the -sheeted cover associated to the trivial torsor is the trivial cover . Therefore, by our work in the case of a trivial cover
There is a action on so that is the stack quotient of by . Pulling back to in both ways shows that the action is given by permuting the sheets of . Under the isomorphism of with of Example 3.1, the action is given by permuting the coordinates.
We conclude
in which the action by is in each case the appropriate restriction of the permutation action on coordinates of . ∎
The space is also interpretable as the space of unordered configurations of points:
Observe that the fat diagonal is exactly the locus of where has stabilizers. The coarse moduli space of is by the fundamental theorem of symmetric functions, with the composite
given by the elementary symmetric polynomials [Art11, §16.1-2]. The composite sends a list of roots to the coefficients of the monic polynomial of degree vanishing at those roots, up to sign:
The assignment is plainly invariant under relabeling the by .
The map to the coarse moduli space is an isomorphism precisely over . The image of in is the closed subscheme cut out by the discriminant of the above polynomial
the square of the Vandermonde determinant. The resulting divisor is the pushforward of to the coarse moduli space .
We summarize the above discussion for general targets in the place of :
Theorem 3.4.
Let be a quasiprojective scheme, and let be the stack quotient by the trivial action.
-
•
The Weil Restriction is the stacky symmetric product:
-
•
The space of monogenerators for , is the th unordered configuration space:
-
•
The complementary space of non-monogenerators is the stack quotient by of the “fat diagonal” of points in which are not pairwise distinct:
3.3. Implications
The rest of the section gives sample applications, exotic examples, and directions based on the correspondence with configuration spaces.
Remark 3.5.
Classical work on the analogues of monogenicity in complex geometry, such as [Han79], has recognized that embeddings into -bundles are closely related to the braid group, essentially because the fundamental group of the configuration space of points in is the braid group on strands. In the scheme theoretic setting, our best analogue of the fundamental group is the étale fundamental group.
The computations above imply that
In [Fur12], it is computed that this space has étale fundamental group a semi-direct product
where is the profinite completion of the braid group on strands and is the absolute Galois group of . As discussed in [Fur12], the conjugation action of on extends to an action by the Grothendieck-Teichmüller group . Conjecturally, . Though all varieties over yield actions of the Galois group , we were surprised to rediscover one of its central representations used in number theory.
The following result is well-known, as the square of the Steinitz class is the discriminant, and the discriminant is a unit when is étale. However, we have a pleasant alternative proof in terms of our universal étale cover.
Theorem 3.6.
If is étale, the Steinitz class is 2-torsion in . If has characteristic , the Steinitz class vanishes.
Proof.
It is enough to show that the Steinitz class is 2-torsion for the universal case . Consider the pullback square
The pushforward is trivialized on the étale cover , as
We find that the descent datum for with respect to this cover has gluing on given by permuting the coordinates by over for each . This is represented by a permutation matrix, which has determinant . Therefore the gluing data for is given locally by multiplying by . Since is 2-torsion in and trivial if has characteristic 2, the result follows. ∎
Example 3.7 (Torsors for finite groups).
Let be a finite group. A -torsor is, in particular, a finite étale map of degree admitting the above description. Notice that the action of on induces an action of on .
The map is classified by a map , the stack of -torsors, and we may regard as pulled back from either the monogenicity space of the universal -torsor, , or the monogenicity space of the universal -fold cover . To compare the two, observe that the left regular representation gives an inclusion upon ordering the set . The induced representable map is essentially independent of the ordering since different orderings induce conjugate maps. The classifying map is the composite with the left regular representation. The monogenicity space is where acts on by permuting the basis vectors by the left regular representation.
A similar description locally holds for other finite étale group schemes. For merely finite flat group schemes such as in characteristic , the group action on the monogenicity space of -torsors still holds but the local decomposition and action do not.
Corollary 3.8.
If is a -torsor for a finite group, and either
-
(1)
is odd
-
(2)
is even and has non-cyclic Sylow 2-subgroup
then the Steinitz class of is trivial in
Proof.
Repeating the construction of Theorem 3.6, we see that if the left regular representation of factors through , the Steinitz class is trivial. The conditions given identify precisely when this happens. ∎
The stacks we study arise naturally in log geometry as “Artin fans” [ACMW14].
Example 3.9 (Moduli spaces of curves in genus 0).
Let be the moduli stack of smooth curves of genus 0 (i.e. ) with marked points. The evident [Wes11] isomorphism with a quotient of configuration space gives:
One can always put the first point at and get equivalent descriptions:
where is the group of affine transformations . The stack quotient classifies local affine equivalence classes of monogenerators, as detailed in Section 4.
One can likewise obtain the other moduli spaces of curves by an ad hoc construction. Consider the universal connected, proper, genus- nodal curve, its relative smooth locus , and the monogenicity space
of the trivial cover over the moduli space . The monogenicity stack is naturally isomorphic to the space of nodal, -marked curves . One can also obtain the open substack of stable curves as the universal Deligne-Mumford locus .
3.4. When is a map étale?
We recall from [Poo06, §6] that a map is étale precisely when the discriminant of the algebra does not vanish. We recall from [Poo06] that there is an algebraic moduli stack of finite locally free algebras and the affine scheme of finite type parametrizing such algebras together with a choice of global basis .
Suppose comes from a finite flat algebra with a global basis , corresponding to a map . There is a trace pairing [Sta20, 0BSY] which we can use to define the discriminant:
Changing changes the function by a unit. The function does not descend to , but the vanishing locus does. Writing , for the open complements of the vanishing locus , a map is étale if and only if factors through the open substack [Poo06, Proposition 6.1].
Remark 3.10.
Most finite flat algebras are not étale, nor are they degenerations of étale algebras. B. Poonen shows the moduli of étale algebras inside of all finite flat algebras cannot be dense by computing dimensions [Poo06, Remark 6.11]. The closure is nevertheless an irreducible component.
What if is not étale? Readers familiar with [Cos06] know one can sometimes endow a ramified map with stack structure and at the ramification to make étale. Then all are -torsors, and not just ramified covers . The ideas in Section 3.2, in particular an analogue of Theorem 3.4, apply in this level of generality. We sketch these ideas over .
Consider
for some . If and is the projective closure of the above affine equation, the projection extends to a finite locally free map . This is in Situation 1.1 so our definitions make sense for it. However is ramified at four points, preventing us from interpreting its monogenicity space using the perspective of this section. Nevertheless, we may observe that the function gives a section of over . The section naturally extends to a section of over all of .
Let . If we work over and endow and with stack structure to obtain a finite étale cover of stacky curves as in [Cos06], the stacky finite étale cover together with the map is parameterized by a representable map to the stack quotient
We can similarly allow and to be nodal families of curves over some base . Maps from nodal curves over entail an -point of the moduli stack of prestable maps to the symmetric product. As in [ABHS, Proposition 2.3], there is an open substack for which the map from the coarse space to is a closed immersion. The stack splits into components indexed by the ramification profiles of the cover of coarse spaces .
There are some subtleties in characteristic —one cannot treat all ramification as a torsor because some ramification is a -torsor in characteristic . The formalism of tuning stacks [ESZ21] is a substitute in arbitrary characteristic.
4. Twisted monogenicity
The Hasse local-to-global principle is the idea that “local” solutions to a polynomial equation over all the -adic fields and the real field can piece together to a single “global” solution over . We ask the same for monogenicity: given local monogenerators, say over completions or local in the Zariski or étale topologies, do they piece together to a single global monogenerator?
The Hasse principle fails for elliptic curves. Let be an elliptic curve over a number field and consider all its places . The Shafarevich-Tate group of an elliptic curve sits in an exact sequence
Elements of are genus-one curves with rational points over each completion that do not have a point over . Similarly, we want sequences of cohomology groups to control when local monogenerators do or do not come from a global monogenerator.
For such a sequence, one needs to know how a pair of local monogenerators can differ. One would like a group or sheaf of groups transitively acting on the set of local monogenerators so that cohomology groups can record the struggle to patch local monogenerators together into a global monogenerator.
Suppose is an algebra extension inducing and are both monogenerators. Then
so each monogenerator is a polynomial in the other:
We can think of the as transition functions or endomorphisms of the affine line . Even though , it is doubtful that or even that are automorphisms of .
One might attempt to find a group containing all possible polynomials . We would then have a homomorphism (of non-commutative monoids) where is some sub-monoid of , the monoid of endomorphisms of (equivalently, the monoid of one-variable polynomials under composition). Even if we only insist that contains , , and , we find that the images of and coincide in since both compose with to the same polynomial. This is not acceptable as and act in distinct ways on monogenerators.
Instead of working with the group of all possible polynomial transition functions as above, we require our transition functions to lie in a group acting on . Two particularly natural options for present themeselves, namely the group sheaves:
Affine transformations are essentially polynomials of degree one under composition. These act on monogenerators:
Definition 4.1 (Twisted Monogenerators).
A ()-twisted monogenerator for is:
-
(1)
a Zariski open cover for elements ,
-
(2)
a system of “local” monogenerators for over , and
-
(3)
units
such that
-
•
for all , we have
-
•
for all , the “cocycle condition” holds:
Two such systems , are equivalent if they differ by further refining the cover or global units : , .
Likewise is -twisted monogenic if there is a cover with ’s as above, but with units (3) replaced by pairs such that each is a unit and .
The elements may or may not come from a single global monogenerator . Nevertheless, the transition functions or define an affine bundle on with global section induced by the ’s. We say is “twisted monogenic” to mean there exists a -twisted monogenerator and similarly say “-twisted” monogenic. Both are clearly Zariski-locally monogenic.
Compare with Cartier divisors:
twisted monogenerator | Cartier divisor |
---|---|
global monogenerator | rational function |
/ action | differing by units. |
We recall the notions of “multiply monogenic orders” and “affine equivalence” in the literature. Two monogenerators are said to be “affine equivalent” if there are , such that . In other words, affine equivalence classes are elements of the quotient . Under certain hypotheses in Remark 4.12, -twisted monogenicity is parameterized by the sheaf quotient . There is almost an “exact sequence”
that dictates whether a twisted monogenerator comes from an affine equivalence class of global monogenerators.
We warm up with a classical approach to -quotients, namely taking . Then we study -twisted monogenerators before finally introducing -twisted monogenerators for arbitrary groups .
There is a moduli space for each notion of twisted monogenicity. We use these moduli spaces now and defer the proof until Theorem 4.26:
Theorem 4.2 (=Theorem 4.26).
Let act on on the left in the natural way, inducing a left action on . The stack quotients and represent - and -twisted monogenerators up to equivalence, respectively.
4.1. -Twisted monogenerators and Proj of the Weil Restriction
Writing and , a twisted monogenerator amounts to a Zariski cover , a system of closed embeddings over , and elements such that
Equivalently, a twisted monogenerator is a line bundle on defined by the above cocycle and a global embedding over . Twisted monogenerators with respect to covers covers are identified if they differ by global units on a common refinement of the covers , i.e., if the corresponding line bundles are isomorphic in a way that identifies the closed embeddings .
For number fields with and , one has . If , then is a monogenerator if and only if is. The multiplication action corresponds to the global action on the vector bundle over .
An action of corresponds to a -grading on the sheaf of algebras [Sta20, 0EKJ]. Locally in , and . The action is the diagonal action and corresponds to the total degree of polynomials in .
The associated projective bundle to the vector bundle is given by the relative Proj [Sta20, 01NS]
with the total-degree grading. The ideal cutting out the complement of is graded by [ABHS, Remark 3.10], defining a closed subscheme .
Definition 4.3.
Define the scheme of projective monogenerators
to be the open complement of the closed subscheme cut out by the graded homogeneous ideal .
The reader may define projective polygenerators in the same fashion.
Lemma 4.4.
The vanishing of the irrelevant ideal of is contained inside of the non-monogenicity locus for .
Proof.
Locally, the lemma states that is not a monogenerator. ∎
Remark 4.5.
We relate the Proj construction to stack quotients by according to [Ols16, Example 10.2.8]. The ring is generated by elements of degree one. Locally, and is generated by the degree one elements . Write for the relative spectrum [Sta20, 01LQ]. The map
is therefore a stack quotient or -torsor.
We have a pullback square
of -torsors and a stack quotient .
Theorem 4.2 states that represents twisted monogenerators, and now we know the quotient stack is wondrously a scheme:
Corollary 4.6.
The scheme represents the -twisted monogenerators of Definition 4.1. That is, is a moduli space for twisted monogenerators. The action is free.
Warning 4.7.
Given a monogenerator and a pair , , write
One may try to define a second action
encoding the degree with respect to , but this action does not define a grading as it is almost never multiplicative. For example, take with monogenerator over . Then
In the case that and , we recall that is the jet space of . Here, the action is multiplicative and induces a second action . The two actions of on a jet
on are and . The Proj of with respect to this second action is known as a “Demailly-Semple jet” or a “Green-Griffiths jet” in the literature [Voj04, Definition 6.1]. For certain , there may be a distinguished one-parameter subgroup, i.e., the image of , that results in a second action and allows an analogous construction.
4.2. -Twisted monogenerators and affine equivalence
We enlarge our study to representing spaces of -twisted monogenerators and the related study of affine equivalence classes of ordinary monogenerators. We delay twisting by general sheaves of groups other than and until the next section. For an -scheme , the automorphism sheaf is the subsheaf of automorphisms in .
Remark 4.8.
The automorphism sheaf has a subgroup of affine transformations under composition. These are identified in turn with via
The automorphism sheaf can be much larger for other . For example,
is an automorphism of .
The automorphism sheaf is not the same as , though they have the same points over reduced rings. See [Dup13] for some discussion over nonreduced rings.
Recall that two monogenerators of an -algebra are said to be equivalent if
where and . Likewise, say that two embeddings of into an bundle over are equivalent if there is in such that . The set of monogenerators up to equivalence is then
If is an isomorphism and , the action of is trivial. Otherwise, the -action is often free:
Lemma 4.9.
The action has trivial stabilizers, for any quasiprojective . If is normal and is not an isomorphism, the action has trivial stabilizers as well.
Proof.
A stabilizer of the action entails a diagram
The fact that is a monomorphism forces to be the identity.
Normality of means is a finite disjoint union of integral schemes [Sta20, 033N]; we assume is integral without loss of generality.
Computing stabilizers of is local, so we may assume is induced by a non-identity finite map of rings where is an integrally closed domain with field of fractions . A stabilizer
implies . If , then and the stabilizing affine transformation is trivial. Otherwise, and . Elements are all integral over . Since is integrally closed, . Hence , a contradiction. ∎
Remark 4.10.
Suppose given transition functions and local monogenerators as in an -twisted monogenerator that may not satisfy the cocycle condition a priori. For normal with as in the lemma, the cocycle condition holds automatically, since acts without stabilizers.
Corollary 4.11.
If is normal, the stack quotient is represented by the ordinary sheaf quotient .
Proof.
If is a free action, the stack quotient coincides with the sheaf quotient . ∎
Remark 4.12.
If is normal, Corollary 4.11 tells us that an -twisted monogenerator is the same as a global section . Equivalence classes of monogenerators are given by the presheaf quotient .
Affine equivalence classes of monogenerators thereby relate to twisted monogenerators in an exact sequence of pointed sets:
As in sheaf cohomology, the second map takes to its torsor of lifts in :
A section of the sheaf quotient
lifts to an affine equivalence class in the presheaf quotient if and only if the induced -torsor is trivial.
The exact sequence is analogous to Cartier divisors. If is an integral scheme with rational function field , the long exact sequence associated to
is analogous to the above.
Remark 4.13.
One can do the same with , or any other group that acts freely. Compare twisted monogenerators with ordinary monogenerators up to -equivalence to obtain a sequence
Freeness of the action is necessary to identify the stack quotient with the ordinary sheaf quotient.
Sometimes, being -twisted monogenic is the same as being -twisted monogenic:
Proposition 4.14.
If is affine, all -torsors on are induced by -torsors:
The corresponding twisted forms of are the same, so we can furthermore identify -twisted monogenerators with -twisted monogenerators.
Proof.
The maps
and |
fit into a short exact sequence
The sheaf is not commutative. Cohomology sets are nevertheless defined for . By Serre Vanishing [Sta20, 01XB] we have for , and therefore is surjective, yielding an identification in all nonzero degrees:
The action is the restriction of that of , factoring
The corresponding twisted forms of are the same. ∎
4.3. Consequences of -twisted monogenicity
We conclude with several consequences of twisted monogenicity and our Theorem 4.19 that shows twisted monogenerators detect class number-one number rings.
The following theorem constrains the line bundles that may be used for twisted monogenicity. This result constrains the possible Steinitz classes of a twisted monogenic extension. This is an effective constraint in geometric situations: see Lemma 5.4. The structure of the set of ideals corresponding to Steinitz classes of number rings is the subject of a variety of open questions. This has traditionally been the domain of class field theory; two notable papers are [McC66] and [Cob10]. For , write . Theorems 1 and 2 of [McC66] imply that if contains a primitive th root of unity, then the Steinitz classes of Galois extensions of degree are precisely the powers in the class group of . Compare this to the following:
Theorem 4.15.
Suppose is -twisted monogenic, with an embedding into a line bundle . Let be the sheaf of sections of . Then
in .
In particular, if an extension of number rings is twisted monogenic, then its Steinitz class is an th power in the class group.
Proof.
Write for the symmetric algebra. Recall that . We have a surjection of -modules
which we claim factors through the projection of -modules . Such a factorization is a local question and local factorizations automatically glue because there is at most one. Locally, we may assume is induced by a ring homorphism and is trivialized. We have a factorization of -modules
due to the existence of a monic polynomial of degree for the image of in [ABHS, Lemma 2.11]. The -modules and are abstractly isomorphic, and any surjective endomorphism of a finitely generated module is an isomorphism [MRB89, Theorem 2.4].
We conclude that globally
Since is invertible, . Taking the determinant,
∎
The literature abounds with finiteness results on equivalence classes of monogenerators, for example:
Theorem 4.16 ([EG17, Theorem 5.4.4]).
Let be an integrally closed integral domain of characteristic zero and finitely generated over . Let be the quotient field of , a finite étale -algebra with , and the integral closure of in . Then there are finitely many equivalence classes of monogenic generators of over .
We have an analogous finiteness result for equivalence classes of -twisted monogenerators:
Corollary 4.17.
Let , , , be as in Theorem 4.16, with induced from . Assume is finitely generated. Then there are finitely many equivalence classes of -twisted monogenerators for .
Proof.
We essentially use the sequence
of Remark 4.12. If this were a short exact sequence of groups, the outer terms being finite would force the middle term to be; our proof is similar in spirit.
Since is quasicompact and there are finitely many elements of the Picard group which are th-roots of the Steinitz class
we can find an affine open cover by finitely many open sets of that simultaneously trivializes all th-roots of the Steinitz class on .
The above sequence of presheaves restricts to the ’s in a commutative diagram
The restriction is zero on the th-roots of the Steinitz class by construction of the ’s. The restriction is injective by the sheaf condition. A diagram chase reveals that the restriction of any section is in the image of . Theorem 4.16 asserts that each set is finite. ∎
Lemma 4.18.
Degree-two extensions are all -twisted monogenic. If is affine, they are also -twisted monogenic.
Proof.
Localize and choose a basis containing 1 to write for some . Given an element so that is also a basis, we may write
Hence are units, and the transition functions come from . By choosing such generators on a cover of , one obtains a twisted monogenerator. Proposition 4.14 further refines our affine bundle to a line bundle. ∎
Theorem 4.19.
A number ring has class number one if and only if all twisted monogenic extensions of are in fact monogenic.
Proof.
If the class number of is one, then all line bundles on are trivial and the equivalence is clear. Mann [Man58] has shown that has quadratic extensions without an integral basis if and only if the class number of is not one: adjoin the square root of , where with non-principal and square-free. By Lemma 4.18, such an extension is necessarily -twisted monogenic. As the monogenicity of quadratic extensions is equivalent to the existence of an integral basis, the result follows. ∎
Remark 4.20.
Given that twisted monogenic extensions and monogenic extensions coincide over , we should ask for an example where we have twisted monogenicity but not monogenicity. All degree 2 extensions of number rings are twisted monogenic as Lemma 4.18 shows. Thus every quadratic extension without an integral basis is twisted monogenic but not monogenic, and [Man58] provides a construction of such extensions. The aim of the following is the very explicit construction of a higher degree example of such an extension. Though we are ultimately unable to prove non-monogenicity in the following example, we hope it gives the reader a concrete sense of the concepts and methods employed in this section.
Example 4.21 (Properly Twisted Monogenic, Not Quadratic).
Let and let , , and be the unique primes of above 3, 5, and 23, respectively. One can compute . Consider , where . On , the local index form with respect to the local basis is . On , we have the local index form with respect to the local basis . We transition via , which is not a global unit, so the extension is twisted monogenic.
To see what is going on more explicitly, we investigate how the transitions affect the local index forms. We have
If and could be chosen to be -integral so that local index form represented a unit of , then would be a global monogenerator. However, -adic valuations tell us is not a monogenerator. One can also apply Dedekind’s index criterion to . Similarly, we have
If and could be chosen to be -integral so that local index form represented a unit of , then would be a global monogenerator. As above, the -adic valuations tell us this cannot be the case. Again, we could also use polynomial-specific methods.
We have shown that is twisted monogenic, but it remains to show that the twisting is non-trivial. We need to show the ideal is not principal. On it can be generated by and on it can be generated by . We transition between these two generators via , exactly as above. Thus our twisted monogenerators correspond to a non-trivial ideal class.
A computer algebra system can compute a -integral basis for :
with index form:
Because is twisted monogenic, there are no common index divisors. Thus we will always find solutions to when we reduce modulo a prime of . We do not expect to be monogenic over ; however, showing that there are no values of such that appears to be rather difficult.
A clever way to get around this issue would be to show that the different of was non-principal. This would preclude monogenicity by prohibiting an integral basis all together. Unfortunately, one can compute that the different is principal, so the extension does have a relative integral basis.
Remark 4.22.
One can perform the same construction of Example 4.21 with radical cubic number rings other than . Specifically, take any radical cubic where , , and are distinct primes with , , and neither nor principal. The ideas behind this construction can be taken further by making appropriate modifications.
4.4. Twisting in general
Throughout this section, fix notation as in Situation 1.1 and work in the category of schemes over equipped with the étale topology. In particular, we allow to be any quasiprojective -scheme.
Definition 4.1 readily generalizes. Replace by any étale sheaf of groups with a left action . A -twisted monogenerator for (into ) is an étale cover , closed embeddings , and elements such that
Say two -twisted monogenerators are equivalent if, after passing to a common refinement of the associated covers, there is a global section so that for all . Equivalently, the ’s glue to a global closed embedding into a twisted form of the same way the -twisted monogenerators give embeddings into a line bundle.
The twisted form arises from transition functions in , meaning there is a -torsor such that is the contracted product:
We have already seen the variant , . Other interesting cases include , , an ellliptic curve acting on itself , etc.
Remark 4.23.
Usually, contracted products are defined for a left action and a right action by quotienting by the antidiagonal action
defined by
We instead take two left actions and quotient by the diagonal action of . The literature often turns left actions into right actions anyway, as in [Bre06, Remark 1.7].
Group sheaves beget stacks classifying -torsors on -schemes with universal -torsor .
Twisted forms of are equivalent to torsors for ([Poo17, Theorem 4.5.2]), as follows: Given a twisted form , we obtain the torsor of local isomorphisms. Given a -torsor , we define a twisted form via contracted product:
The stack is thereby a moduli space for twisted forms of with universal family . An action lets one turn a -torsor into a twisted form
classified by the map .
The automorphism sheaf acts on the scheme via postcomposition with the embeddings , yielding a map of sheaves
Similarly, the automorphism sheaf acts on on the right via precomposition:
The induced map sends a twisted form of to the twisted form of given by looking at closed embeddings into . We package these twisted forms into a universal version: .
Definition 4.24 ().
Let be the -stack whose -points are given by
The map is representable by schemes, since pullbacks are twisted forms of itself:
The universal torsor over is , but the universal twisted form is obtained by the contracted product with over :
One can exhibit as an open substack of the Weil Restriction of as in [ABHS, Proposition 2.3]. There is a universal closed embedding over into the universal twisted form of as in the definition of :
The universal case is concise to describe but unwieldy because need not be finite, smooth, or well-behaved in any sense. We simplify by specifying our twisted form to get a scheme or by specifying the structure group .
Fixing the structure group requires for the specified sheaf of groups and some -torsor . These -twisted forms are parameterized by the pullback
The fibers of over maps are again twisted forms of . If and is a -twisted form, we may refer to the existence of sections of by saying is -twisted -genic, etc.
Remark 4.25.
Trivializing is not the same as trivializing the torsor that induces unless the group is itself. For example, take the trivial action .
Theorem 4.26.
The stack of twisted monogenerators is isomorphic to over . More generally, for any sheaf of groups , we have an isomorphism over .
Proof.
Address the second, more general assertion and let be an -scheme. Write , , etc. A -point of is a -torsor and a solid diagram
with a closed immersion. Form by pullback: is a left -torsor with an equivariant map to with the diagonal action. The map entails a pair of equivariant maps and . The map over forces . These data form an equivariant map over , or . Reverse the process to finish the proof. ∎
To see this theorem in practice, we have the following example.
Definition 4.27.
The group sheaf of affine transformations is the set of functions
where and , under composition. Note .
Example 4.28.
Let be the spectrum of a quadratic order such as , and take . The space of -generators is according to [ABHS, Proposition 4.5]. Take the quotient by the groups of affine transformations:
These quotients represent twisted monogenerators according to Theorem 4.26. The corresponding -torsors were classically identified with Azumaya algebras and Severi-Brauer varieties (see the exposition in [Kol16]) or twisted forms of . These yield classes in the Brauer group via the connecting homomorphism from
The same holds locally for any degree-two extension with integral using [ABHS, Proposition 4.5].
If is an abelian variety over a number field , let be a -torsor inducing a twisted form of . Given a twisted monogenerator , one can try to promote to a global monogenerator by trivializing and thus .
Suppose one is given trivializations of over the completions at each place. Whether these glue to a global trivialization of over and thus a monogenerator is governed by the Shafarevich-Tate group .
Given a -twisted monogenerator with local trivializations, the Shafarevich-Tate group obstructs lifts of to a global monogenerator the same way classes of line bundles in Pic obstruct -twisted monogenerators from being global monogenerators. Theorem 4.19 showed a converse—nontrivial elements of Pic imply twisted monogenerators that are not global monogenerators.
Question 4.29.
Is the same true for ? Does every element of the Shafarevich-Tate group arise this way?
The Shafarevich-Tate group approach is useless for or because of Hilbert’s Theorem 90 [Sta20, 03P8]:
The same goes for any “special” group with étale and Zariski cohomology identified. The strategy may work better for or elliptic curves .
Remark 4.30.
This section defined -twisted monogenerators using covers in the étale topology, whereas Definition 4.1 used the Zariski topology. For or , either topology gives the same notion of twisted monogenerators. Observe that has the same Zariski and étale cohomology by Hilbert’s Theorem 90. The same is true for by [Sta20, 03P2] and so also .
5. Examples of the scheme of monogenerators
We conclude with several examples to further illustrate the interaction of the various forms of monogenicity considered in this paper. We will make frequent reference to computation of the index form using the techniques of the previous paper in this series. Some of these examples were already considered in the previous paper, but are revisited in order to add some commentary on their relationship to notions of local monogenicity.
5.1. Orders in number rings
Example 5.1 (Dedekind’s Non-Monogenic Cubic Field).
Let denote a root of the polynomial and consider the field extension over . When Dedekind constructed this example [Ded78] it was the first example of a non-monogenic extension of number rings. Indeed two generators are necessary to generate : take and , for example. In fact, is a -basis for . The matrix of coefficients with respect to the basis is
Taking its determinant, the index form associated to this basis is
Were the extension monogenic, we would be able to find so that the index form above is equal to .
In fact, is not even locally monogenic. By Lemma 2.6, we may check by reducing at primes. Over the prime the index form reduces to
and iterating through the four possible values of shows that the index form always to reduces to 0. That is, is a common index divisor.
Dedekind showed that is non-monogenic, not by using an index form, but by deriving a contradiction from the -factorization of the ideal 2, which splits into three primes. In our terms, has three points over , all with residue field . Therefore, condition (2) of Theorem 1.10 for monogenicity at the prime fails, so is not monogenic.
With base extension, one can eventually resolve the obstructions presented by common index divisors. The following example illustrates a non-maximal order where we have a slightly different obstruction to monogenicity.
Example 5.2 (An order that fails to be monogenic over geometric points).
Consider the extension of . (This is not the maximal order of .) We recall from [ABHS, Example 4.13] that the index form with respect to the basis is
We can contrast the above examples with the following example where the obstruction to monogenicity is global as opposed to local.
Example 5.3 (A Zariski-locally monogenic, but not twisted-monogenic extension).
Here we take a closer look at one member of the family in Example 2.8. Let , . The ring of integers is not monogenic over . Let , . We recall from [ABHS, Example 4.14] that is a -basis for and the associated index form is . Thus, for a given choice of , the primes which divide the value are precisely the primes at which will fail to generate the extension.
The values obtained by this index form are modulo . Since 5 is a unit in , we do have local monogenerators over . Similarly, we have local monogenerators over . Together, and form an open cover, and we see that this extension is Zariski-locally monogenic. However, as we can see by reducing modulo 7, the index form cannot be equal to . Therefore there are no global monogenerators and is not monogenic: in the language of [ABS20], has a local obstruction to monogenicity, despite being locally monogenic.
5.2. Maps of curves
One benefit of our more geometric notion of monogenicity is it allows us readily ask questions about monogenicity in classical geometric situations with the same language that we use in the arithmetic context. Our next examples concern the case that is a finite map of algebraic curves, which is essentially never monogenic. On the other hand, we find explicit examples of -twisted monogenic . Theorem 4.15 constrains the possible line bundles that we may use to show -twisted monogenicity. We make this precise in the lemma below.
Lemma 5.4.
Let be a finite map of smooth projective curves of degree and let denote the genus.
-
(1)
is only monogenic if it is the identity map;
-
(2)
If is -twisted monogenic, then is divisible by in . Moreover, if factors through a closed embedding into a line bundle with sheaf of sections , then
Proof.
To see (1), note that a map is determined by a global section of . Since is a proper variety, the global sections of are constant functions. It follows that a map is constant on fibers of . Therefore cannot be an immersion unless has degree 1, i.e., is the identity.
First, we will investigate one of the most basic families of maps of curves.
Example 5.5 (Maps ).
Let be an algebraically closed field and let be a finite map of degree . If , then is trivially monogenic. When , Lemma 5.4 tells us that cannot be monogenic, while Lemma 4.18 tells us that is -twisted monogenic. Lemma 5.4 tells us that for degrees the map is neither monogenic nor -twisted monogenic, although Theorems 2.1 and 1.10 tell us that is Zariski-locally monogenic.
Working with instead of an algebraically closed field, consider the map given by . We will show by direct computation that this map is -twisted monogenic. Write and for the standard affine charts of the target . The map is then given on charts by
and
Let us compute . Over , has -basis . Let be the coordinates of , with universal map
The index form associated to this basis is
Similarly, has -basis , analogous coordinates , and the index form associated to this basis is
An element of (resp. ) is a unit if and only if it is , so
We can see directly that is not monogenic: the condition that a monogenerator on glue with a monogenerator on is that
But this is impossible to satisfy since .
Lemma 5.4 tells us that if is twisted monogenic, the line bundle into which embeds must have degree 1. Let us therefore attempt to embed into the line bundle with sheaf of sections . The sheaf restricts to the trivial line bundle on both and , and a section glues to a section if
Embedding into this line bundle is therefore equivalent to finding a monogenerator on , and a monogenerator on such that
Bearing in mind that on , we find a solution by taking positive signs, , and . Therefore is twisted monogenic.
Lemma 5.4 tells us that we must pass to higher genus to find a -twisted monogenic cover of of degree greater than . Here is an example where the source is an elliptic curve.
Example 5.6 (Twisted monogenic cover of degree 3).
Let be the Fermat elliptic curve . Consider the projection from , i.e., the map defined by . Write and for the standard affine charts of . The map is given on charts by and . The gluing on overlaps is given by on and by on .
We now compute . Note that over , has the -basis . The index form associated to is
Over , has the -basis . The index form associated to this basis is
An element of or is a unit if and only if it is . This implies that
We see that there are no global sections of , since coefficients of cannot match on overlaps.
However, if we twist so that we are considering embeddings of into , then the condition for a monogenerator on to glue with a monogenerator on is that
This is satisfied, for example by taking the positive sign for both generators and . Therefore is twisted monogenic with class .
References
- [ABHS] Sarah Arpin, Sebastian Bozlee, Leo Herr, and Hanson Smith, The scheme of monogenic generators i: Representability.
- [ABS20] Levent Alpöge, Manjul Bhargava, and Ari Shnidman, A positive proportion of cubic fields are not monogenic yet have no local obstruction to being so, 2020.
- [ABS21] by same author, A positive proportion of quartic fields are not monogenic yet have no local obstruction to being so, 2021.
- [ACMW14] Dan Abramovich, Qile Chen, Steffen Marcus, and Jonathan Wise, Boundedness of the space of stable logarithmic maps, Journal of the European Mathematical Society 19 (2014).
- [Art11] M. Artin, Algebra, Pearson Prentice Hall, 2011.
- [BCP97] Wieb Bosma, John Cannon, and Catherine Playoust, The Magma algebra system. I. The user language, J. Symbolic Comput. 24 (1997), no. 3-4, 235–265, Computational algebra and number theory (London, 1993). MR MR1484478
- [Bre06] Lawrence Breen, Notes on 1- and 2-gerbes, arXiv Mathematics e-prints (2006), math/0611317.
- [Cob10] Alessandro Cobbe, Steinitz classes of tamely ramified Galois extensions of algebraic number fields, J. Number Theory 130 (2010), no. 5, 1129–1154. MR 2607305
- [Cos06] Kevin Costello, Higher genus Gromov-Witten invariants as genus zero invariants of symmetric products., Ann. Math. (2) 164 (2006), no. 2, 561–601 (English).
- [Ded78] Richard Dedekind, Über den Zusammenhang zwischen der Theorie der Ideale und der Theorie der höheren Kongruenzen, Gött. Abhandlungen (1878), 1–23.
- [DH84] Tom Duchamp and Richard Hain, Primitive elements in rings of holomorphic functions, Journal für die reine und angewandte Mathematik 346 (1984), 199–220.
- [Dup13] Taylor Dupuy, Automorphisms of the Affine Line over Nonreduced Rings, arXiv e-prints (2013), arXiv:1311.0973.
- [EG17] Jan-Hendrik Evertse and Kálmán Győry, Discriminant equations in Diophantine number theory, New Mathematical Monographs, vol. 32, Cambridge University Press, Cambridge, 2017. MR 3586280
- [ESZ21] Jordan S. Ellenberg, Matthew Satriano, and David Zureick-Brown, Heights on stacks and a generalized Batyrev-Manin-Malle conjecture, arXiv e-prints (2021), arXiv:2106.11340.
- [Fur12] Hidekazu Furusho, Galois action on knots i: Action of the absolute galois group, Quantum Topology 8 (2012).
- [GL69] E. A. Gorin and V. Ja. Lin, Algebraic equations with continuous coefficients and some problems of the algebraic theory of braids, Math. USSR-Sbornik 7 (1969), 569–596.
- [Han79] V. L. Hansen, Polynomial covering spaces and homomorphisms into the braid groups, Pacific J. Math. 81 (1979), 399–410.
- [Han80] by same author, Coverings defined by weierstrass polynomials, J. reine angew. Math. 314 (1980), 29–39.
- [Hen94] K. Hensel, Arithmetische Untersuchungen über die gemeinsamen ausserwesentlichen Discriminantentheiler einer Gattung, J. Reine Angew. Math. 113 (1894), 128–160. MR 1580350
- [Hera] Leo Herr, If the normalization is affine, is it affine? (if quasiaffine), MathOverflow, version: 2020-11-30.
- [Herb] by same author, When is a twisted form coming from a torsor trivial?, MathOverflow, version: 2020-09-06.
- [HW21] Leo Herr and Jonathan Wise, Costello’s pushforward formula: errata and generalization, arXiv e-prints (2021), arXiv:2103.10348.
- [JLM+17] Lena Ji, Shizhang Li, Patrick McFaddin, Drew D. Moore, and Matthew Stevenson, Weil restriction for schemes and beyond, 2017.
- [Kol16] János Kollár, Severi-Brauer varieties; a geometric treatment, arXiv e-prints (2016), arXiv:1606.04368.
- [LMF21] The LMFDB Collaboration, The L-functions and modular forms database, http://www.lmfdb.org, 2021, [Online; accessed 15 July 2021].
- [Lø80] K. Lønsted, On the embedding of coverings into 1-dimensional bundles, J. reine angew. Math. 317 (1980), 88–101.
- [Man58] Henry B. Mann, On integral bases, Proc. Amer. Math. Soc. 9 (1958), 167–172. MR 93502
- [McC66] Leon R. McCulloh, Cyclic extensions without relative integral bases, Proc. Amer. Math. Soc. 17 (1966), 1191–1194. MR 225760
- [MRB89] H. Matsumura, M. Reid, and B. Bollobas, Commutative ring theory, Cambridge Studies in Advanced Mathematics, Cambridge University Press, 1989.
- [Nar04] Władysław Narkiewicz, Elementary and analytic theory of algebraic numbers, third ed., Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2004. MR 2078267
- [Neu99] Jürgen Neukirch, Algebraic number theory, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 322, Springer-Verlag, Berlin, 1999, Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder. MR 1697859
- [Ols16] Martin Olsson, Algebraic spaces and stacks, American Mathematical Society Colloquium Publications, vol. 62, American Mathematical Society, Providence, RI, 2016. MR 3495343
- [Ple74] P. A. B. Pleasants, The number of generators of the integers of a number field, Mathematika 21 (1974), 160–167. MR 382237
- [Poo06] Bjorn Poonen, The moduli space of commutative algebras of finite rank, Journal of the European Mathematical Society 10 (2006).
- [Poo17] Bjorn Poonen, Rational points on varieties, American Mathematical Society, 2017.
- [Ser79] Jean-Pierre Serre, Local fields, Graduate Texts in Mathematics, vol. 67, Springer-Verlag, New York-Berlin, 1979, Translated from the French by Marvin Jay Greenberg. MR 554237
- [Ser06] E. Sernesi, Deformations of algebraic schemes, Grundlehren der mathematischen Wissenschaften, Springer Berlin Heidelburg, 2006.
- [Sta20] The Stacks Project Authors, Stacks Project, http://stacks.math.columbia.edu, 2020.
- [The19] The Sage Developers, Sagemath, the Sage Mathematics Software System (Version 8.7), 2019, https://www.sagemath.org.
- [Voj04] Paul Vojta, Jets via Hasse-Schmidt Derivations, arXiv Mathematics e-prints (2004), math/0407113.
- [Wes11] Craig Westerland, Configuration spaces in topology and geometry, Austral. Math. Soc. Gaz. 38 (2011), no. 5, 279–283. MR 2896208