118
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Kummer–Witt–Jackson algebras

Received 05 Oct 2023, Accepted 29 Apr 2024, Published online: 06 Jun 2024

Abstract

This paper is concerned with the construction of a small, but non-trivial, example of a polynomial identity algebra, which we call the Jackson algebra, that will be used in sequels to this paper to study non-commutative arithmetic geometry. In this paper this algebra is studied from a ring-theoretic and geometric viewpoint. Among other things it turns out that this algebra is a “non-commutative family” of central simple algebras and thus parameterizes Brauer classes over extensions of the base.

2020 Mathematics Subject Classification:

1 Introduction

The geometric study of polynomial identity (PI) algebras, and in particular maximal orders, has seen a growing interest at least since the early 1990’s, in particular by the school following Michael Artin. For a recent example see [Citation4]. Most of this study is concerned with algebras over algebraic surfaces (the case of curves is rather well-understood) and there are beautiful results already, but a complete classification, in particular over non-algebraically closed fields (which is my main interest), seems to be out of reach at the moment.

However, the algebras that will appear in this paper are algebras over higher-dimensional schemes, and the main example, denoted Jx, will turn out to live over a three-fold, say X for now. This three-fold actually seems to be rational in many cases. On the other hand, the restriction of Jx to divisors on X will give a particular example of the algebras studied in the literature (see the already mentioned [Citation4] and the references therein), at least when extending to the projective closure of the center Spec(Z(Jx)). In fact, X=Spec(Z(Jx)).

My impetus for studying this algebra is actually two-fold (no pun intended). The first is that this algebra appeared as a q-deformation of the Lie algebra sl2a long time ago in a paper I wrote with S.D. Silvestrov [Citation16]. In that paper was left a few open questions regarding some ring-theoretic and homological properties concerning this algebra which we weren’t able to solve at that moment. A few years later I was able to do this, but I never really wrote it up in any readable manner and it was shelved. Then a few years ago I became interested in applying non-commutative deformation theory to arithmetic geometry and I needed a simple, but non-trivial, algebra to use as testing ground. I remembered the algebra in quarantine. Whence the second reason for doing a somewhat detailed ring-theoretic analysis of this algebra.

The construction of this algebra Jx will proceed in several steps. First, following the origins of the q-deformed sl2, we construct a family of hom-Lie algebras (see Section 2) and we also introduce a new class of such, infinitesimal hom-Lie algebras, that will play a specific rôle in a later paper concerning torsion points on elliptic curves. Then, in Section 3, we construct “enveloping algebras” of the hom-Lie algebras and restrict to a specific type of algebras which we call Kummer–Witt hom-Lie algebras and their enveloping algebras, Kummer–Witt algebras. We do all this globally over a general scheme, which is strictly not necessary for the rest.

Primarily in order to simplify notation, we therefore restrict to the affine case and study in Section 4.2, the ring-theoretic and homological properties of these Kummer–Witt algebras. In particular, the algebra Jx will make its entrance as a canonical subalgebra of such a Kummer–Witt algebra (see Section 4.4). The section begins with a reminder on the infinitesimal structure of polynomial identity algebras and the definition of “Auslander-regularity”.

In Section 4.5 we will use a (unpublished) method of A. Bell and S. P. Smith [Citation22] to compute the center and prove, among other things, that after a base extension to the algebraic closure of the base field, the center has rational singularities (see Theorem 4.17).

Section 5 begins by introducing a simplified version of what is to be viewed as a non-commutative scheme for us. A detailed and general definition can be found in [Citation15]. The crucial property of a non-commutative space is that it has a more intricate infinitesimal structure captured by the existence of non-trivial Ext1-groups between different points. Elements in these groups are to be interpreted as “tangents” between points. Also, in this section is a definition of what is to be meant by an “L-rational point” on a non-commutative scheme. Next we define a family of quadratic divisors on the non-commutative space, XJx, associated Jx.

After that, in Section 5.2, comes the first arithmetic discussion of Jx. Namely, we prove that fibers in XJx over an open subscheme S of the central subscheme X=Spec(Z(Jx)) parameterizes classes in Brauer groups by these fibers containing symbol algebras. We also prove that this fibration over X includes quantum Weyl algebras over the complement of S in X.

Finally, in Section 6 we look at rational points on XJx. First we find all one-dimensional points (the “commutative points”) and compute the tangent structure of these points, i.e., compute the Ext1-groups. After that we look at higher-dimensional points and use a (slightly modified) construction of D. Jordan [Citation13] to construct families of rational points that, over the algebraic closure, contains all rational points up to isomorphism. There are two types of higher-dimensional points: torsion points and torsion-free points. We prove that there are only a discrete set of torsion points, but a continuous family of torsion-free points. We conclude by giving a somewhat detailed example.

Notation

We will adhere to the following notation throughout.

  • All rings are unital.

  • For a general algebra (not necessarily commutative) Mod(A) denotes the category of left A-modules.

  • The notation Max(A) denotes the set of maximal ideals, while Specm(A) denotes the maximal spectrum of A (if A is commutative).

  • The notation Mod(A), denotes the set (groupoid) of isoclasses of all A-modules. All modules are left modules unless otherwise explicitly specified. The class of modules with annihilator ideal being prime is denoted ModΔ(A).

  • Z(A) denotes the center of A.

  • For pa prime in A, k(p) denotes the residue class field of p.

  • Abelian sheaves are denoted with scripted letters.

  • All schemes and algebras are noetherian. Schemes are also assumed to be separated. Many results surely hold without this assumption but it is cumbersome to keep track of this hypothesis in any given situation so we make the blanket assumption of separatedness throughout for simplicity.

2 Algebras of twisted derivations

Let X/S be an S-scheme and let A be a coherent sheaf of commutative OX-algebras. Assume that σ is an algebra endomorphism on A. Then a global (σ-)twisted S-derivation on A is an operator ΔU(ab)=ΔU(a)b+σ(a)ΔU(b),a,bA(U),with UX open,

Example 2.1.

The canonical example of a σ-derivation is a map AA on the form Δσ(U):=aU(idσ):A(U)A(U),aUA(U),with UX open.

In fact, for many algebras these types of maps are the only σ-derivations available.

Below UX will always denote an open subset of the scheme X. We define the A-module AnnA(Δ) as AnnA(Δ)(U):={aA(U)|aΔ(b)=0,for all bA(U)}.

Assume that ΔU°σ=qU·σ°ΔU,qUA(U),σ(Ann(Δ))Ann(Δ),and form the left A-module A·Δ by (A·Δ)(U):=A(U)·ΔU.

On A·Δ we introduce the product ​​​·,·​​​ by ​​​a·ΔU,b·ΔU​​​ U:=σ(a)·ΔU(b·ΔU)σ(b)·ΔU(a·ΔU),a,bA(U).

We now have the following theorem.

Theorem 2.1.

The above product is OS-linear and satisfies

  1. ​​​a·ΔU,b·ΔU​​​ U=(σ(a)ΔU(b)σ(b)ΔU(a))·ΔU;

  2. ​​​a·ΔU,a·ΔU​​​ U=0;

  3. a,b,c(​​​σ(a)·ΔU,​​​b·ΔU,c·ΔU​​​ U​​​ U+qU·​​​a·ΔU,​​​b·ΔU,c·ΔU​​​ U​​​ U)=0,

where a,b,cA(U).

The proof of this global version is simply a standard descent argument using the affine version as given in [Citation10].

The following definition was introduced in [Citation14] as a generalization of a hom-Lie algebra. For our purposes the definition as given below is certainly overkill but we introduce it in its full generality nonetheless. Example 2.2 gives the main example relevant for us.

Let G denote a finite group scheme acting on X over S, and let A be an OX[G]-sheaf of OX-algebras. This means that A is an OX-algebra together with a G-action, compatible with the G-action on X in the sense that σ(xa)=σ(x)σ(a), xOX, aA.

Let AG denote the skew-group algebra of G over A. Recall that this is the free algebra A{g1,g2,,gn}, giG, with product defined by the rule (agi)·(bgj)=agi(b)gigj.

This defines an associative algebra structure.

Definition 2.1.

Given the above data, a (G-)equivariant hom-Lie algebra on X over A is an AG-module M together with, for each open UX, an OX-bilinear product ​​​·,·​​​ U on M(U) such that

(hL1.) ​​​a,a​​​ U=0, for all aM(U);

(hL2.) for all σG and for each σ a qσA(U), the identity a,b,c{​​​σ(a),​​​b,c​​​ U​​​ U+qσ·​​​a,​​​b,c​​​ U​​​ U}=0,holds.

A morphism of equivariant hom-Lie algebras (M,G) and (M,G) is a pair (f,ψ) of a morphism of OX-modules f:MM and ψ:GG such that f°σ=ψ(σ)°f, and f(U)(​​​a,b​​​ M;U)=​​​f(U)(a),f(U)(b)​​​ M;U.

Hence, an equivariant hom-Lie algebra is a family of (possibly isomorphic) products parameterized by G.

Definition 2.2.

A product ​​​·,·​​​ σ in the equivariant structure, for fixed σG, is a hom-Lie algebra on M.

Example 2.2.

The main example for us in this paper is: the A-module A·Δσ, σG defines an equivariant hom-Lie algebra by Theorem 2.1. Each σ gives a hom-Lie algebra structure on A·Δσ.

Some reasons why σ-derivations are important (and actually prevalent in abundance) in arithmetic and geometry, can be found in [Citation14] and the references therein.

2.1 Infinitesimal hom-Lie algebras

Let X! be an infinitesimal thickening of X. This means that X is defined as a closed subscheme of X! by a nilpotent sheaf of ideals I. The order of the thickening is defined as the least integer n such that In=0. By construction X and X! have the same underlying topological space. Notice that OX! is an OX-algebra.

Let σ be an OS-linear automorphism of X!. This induces, by definition, an OS-linear automorphism σX on X, i.e., σX=σ|X. We will for simplicity assume that σ is a lift of the identity on X. In other words, σ|X=idX.

Put, for all open UX, Δσ(U):=aU(idσ):OX!(U)OX!(U),aUOX(U).

Then (OX!·Δσ,​​​,​​​) is called an infinitesimal hom-Lie algebra on X.

The canonical example is the following. To simplify the discussion, we restrict to affine schemes. Everything globalizes without problem. So, let X=Spec(R), with RCom(k). Put t¯:={t1,t2,,td}. Then X!:=Spec(R[t¯]/(t¯)n) is an n-th order infinitesimal thickening of X=Spec(R)=Spec((R[t¯]/(t¯)n)/(t¯)).

We will look at the particular case of d = 1. Put R!:=R[t]/(t)n. Then e¯i:=tiΔσ, 0in1, is a basis for R!·Δσ as an R!-module. We will consider the automorphism σ(t)=qt, with qk×, q1, and aR. Notice that σR=id.

Let Δσ:=a(1q)1(idσ). Then a simple induction argument gives that Δσ(ti)=a[i]qti, where we have put [i]q:=1qn1q=1+q+q2++qi1. Notice that [0]q=0 and [1]q=1. Using Theorem 2.1 (i) a small computation gives that ​​​e¯i,e¯j​​​=a(qi[j]qqj[i]q)e¯i+j.

Observe that when i+jn, then ​​​e¯i,e¯j​​​=0 since in that case ti+j=0. In addition, ​​​e¯0,e¯i​​​=a[i]qe¯i, for all i.

Example 2.3.

When n = 2, we get the solvable R-Lie algebra ​​​e¯0,e¯1​​​=ae¯1.

Example 2.4.

When n = 3, we also get a solvable R-Lie algebra: ​​​e¯0,e¯1​​​=ae¯1,​​​e¯0,e¯2​​​=a[2]qe¯2,​​​e¯1,e¯2​​​=0.

One would be tempted to conjecture that these are Lie algebras for all n. However, this is not true as the case n = 4 shows.

Example 2.5.

So when n = 4 we get ​​​e¯0,e¯i​​​=a[i]qe¯i,​​​e¯1,e¯2​​​=aqe¯2,​​​e¯1,e¯2​​​=0,​​​e¯1,e¯2​​​=0.

This is not a Lie algebra since, for instance, 0,1,2​​​e¯0,​​​e¯1,e¯2​​​​​​=a2q(q1)e¯3,which is not zero unless a = 0 (trivial) or q = 1.

That the case n = 2 gives a Lie algebra is quite natural, but it seems that this should be the case also for n = 3, is more of a coincidence.

3 Enveloping algebras

We will now use Theorem 2.1 to construct an “enveloping” algebra. By this we mean an associative algebra E constructed on the given non-associative structure.

This algebra E is constructed as follows. Let A be a finitely generated (commutative) OX-algebra, σAutOX(A) and let ΔD​rσ(A), such that over UX, ΔU:=αU·(idσ),αUA(U).

Then Theorem 2.1 endows A·Δ with a non-associative algebra structure. It is clear that, over U, the elements e¯k¯:=y1k1y2k2ynkn·ΔU,k¯Z0n,form a basis over U for A·Δ as an A-module, where y1,y2,,yn are generating sections of A over U. Then we have the relations (e¯k¯)σ·e¯l¯(e¯l¯)σ·e¯k¯=​​​e¯k¯,e¯l¯​​​,

so we can form E(A·Δ)(U):=OX(U){e¯k¯|k¯Z0n}((e¯k¯)σ·e¯l¯(e¯l¯)σ·e¯k¯​​​e¯k¯,e¯l¯​​​).

Obviously this is in general a very complicated algebra because it is infinitely presented, exactly as the universal enveloping algebra of the classical Witt–Lie algebra for instance. Things simplify considerably if A is finite as an OX-module.

So assume that A is locally free of (constant) finite rank as OX-algebra, given over U by A(U)=OX(U)y1OX(U)y2OX(U)yn.

Put e¯i:=yi·ΔUand eij:=(e¯i)σ·e¯j(e¯j)σ·e¯i​​​e¯i,e¯j​​​.

Then (3.1) E(A·Δ)(U)=OX(U){e¯1,e¯2,,e¯n}/(eij).(3.1)

3.1 Enveloping algebras of infinitesimal hom-Lie algebras

We continue with the instance d = 1 as to not get too bogged down in awkward notation. That is, we consider the thickening R!=R[t]/(t)n. This means that A from the previous section, corresponds to R!.

Here e¯iσ=(tiΔσ)σ=σ(ti)Δσ=qitiΔσ, with Δσ=idσ. From this we see (e¯i)σ·e¯j(e¯j)σ·e¯i=σ(ti)Δσ(tjΔσ)σ(tj)Δσ(tiΔσ)=qitiΔσ(tjΔσ)qjtjΔσ(tiΔσ)=qie¯ie¯jqje¯je¯i,whence eij=qie¯ie¯jqje¯je¯ia(qi[j]qqj[i]q)e¯i+j.

Therefore, E(R!·Δσ)=R{e¯0,e¯1,,e¯n1}(e¯ie¯jqjie¯je¯ia([j]qqji[i]q)e¯i+j),where we have divided by q i for esthetic reasons.

We continue the examples from Section 2.1.

Example 3.1.

When n = 2 we get E(R!·Δσ)=R{e¯0,e¯1}(e¯0e¯1qe¯1e¯0ae¯1).

By changing basis e¯0e¯0+a(1q)1, we see that this algebra is in fact isomorphic to the famous quantum plane QR,q2=R[x,y]/(xyqyx). Observe that the algebra ​​​e¯0,e¯1​​​=ae¯1 is a Lie algebra but E(R!·Δσ) is not the universal enveloping algebra for this Lie algebra. In other words, first order thickenings (or deformations) give quantum planes!

Example 3.2.

In the case n = 3 we compute the relations e¯0e¯1qe¯1e¯0=ae¯1,e¯0e¯2q2e¯2e¯0=a[2]qe¯2,e¯1e¯2qe¯2e¯1=0,and so E(R!·Δσ)=R{e¯0,e¯1,e¯2}(e¯0e¯1qe¯1e¯0ae¯1e¯0e¯2q2e¯2e¯0a[2]qe¯2e¯1e¯2qe¯2e¯1).

Using the same change of basis as in the previous example, we get the isomorphic algebra (3.2) E(R!·Δσ)QR,q3:=R{e¯0,e¯1,e¯2}(e¯0e¯1qe¯1e¯0,e¯0e¯2q2e¯2e¯0,e¯1e¯2qe¯2e¯1).(3.2)

The ring QR,q3 is a quantum affine three-space. This means that QR,q3 comes associated with a solvable Lie algebra.

We will be able to say more concerning these algebras later as they have nice ring-theoretic properties (see Section 4.3).

We leave the case n = 4 for the reader.

3.2 Kummer–Witt hom-Lie algebras

We keep the notation from above and further denote the algebra structure on A over U by yiyj=k=0naijkyk,aijkOX(U),where the yi are the algebra generators of A over U. Let σ be the OX-linear algebra morphism on A defined by σ(yi)=qiyi, qiOX(U) and let Δ be the σ-derivation from the previous section. Put e¯i:=yi·Δ. Then, from Theorem 2.1, the pair (3.3) WAσ:=(A·Δ,​​​,​​​),​​​e¯i,e¯j​​​=ak=0n(qiqj)aijke¯k(3.3) defines a hom-Lie algebra structure on A·Δ. We call WAσ the Witt hom-Lie algebra over X attached to A and σ.

Remark 3.1.

The construction just given is obviously not dependent on the particular choice σ(yi)=qiyi. Any other automorphism can be used. However, the result will, of course, be more complicated and harder to write out.

From now on we assume that the n-th roots of unity are included in OX. Fix a primitive such root ζ=ζn and consider the case when A is a uniform cyclic extension of OX. In other words, we have an invertible L and a section t = tU over each U, such that A(U)=OX(U)[t]/(tnxU),xUOX(U).

Then, with yi=ti, and σ(t)=ζrt, we see that σ(yi)=ζriyi, and the product on WAσ becomes (3.4) ​​​e¯i,e¯j​​​=ζri(1ζr(ji))xU e¯{i+j​​​  mod n},ij.(3.4) where () means that () is included when i+jn. We call the resulting hom-Lie algebra the Kummer–Witt hom-Lie algebra of level r and denote it WA1/n(r). The hom-Lie algebras WA1/n(r1) and WA1/n(r2), r1r2, are in general non-isomorphic. The algebra WA1/n(r) is called the r-th twist of WA1/n:=WA1/n(1). Clearly WA1/n(0) is the abelian hom-Lie algebra.

Observe that WA1/n(μn):={WA1/n(r)|0rn1} is the equivariant hom-Lie structure associated with A and G=μn, the group (scheme) of n-th roots of unity. The structure WA1/n(μn) is a μn-torsor in a natural way. In addition, note that we are not making any assumptions on n being invertible on the base.

The above gives immediately that E(WA1/n), the Kummer–Witt algebra, is given by the relations (3.5) e¯ie¯jζr(ji)e¯je¯i(1ζr(ji))xU e¯{i+j​​​  mod n},(3.5) for ij.

Remark 3.2.

The reason we refer to (3.3) as a Witt-hom-Lie algebra is the similarity in form and construction between this algebra and the Witt-Lie algebra. In fact, the infinitesimal hom-Lie algebras are completely analogous to the classical Witt-Lie algebra in characteristic p, and the algebra (3.3) is a finite-rank analogue of the infinite-dimensional Witt-Lie algebra (sometimes called the “centreless Virasoro algebra”) appearing in conformal field theory, for instance.

Remark 3.3.

The same game can clearly be played with Artin–Schreier extensions. We invite the reader to write out the corresponding relations for him/herself.

4 Non-commutative rings from cyclic covers

4.1 Polynomial identity algebras

Let A be a polynomial identity (PI) algebra. Recall that there are two disjoint subsets of Specm(Z(A)), the Azumya locus, azu(A), and the ramification locus, ram(A), that describe the behavior of A as a module over Z(A). A maximal ideal m is in azu(A) if and only if A/m=AZ(A)k(m) is a central simple algebra over k(m). Then ram(A):=Specm(Z(A))azu(A). It is known that ram(A) is the support of a Cartier divisor in Spec(Z(A)) (see e.g., [Citation12, III.2.5]).

Let M:=A/M and N:=A/N, be two simple A-modules, with M,NMax(A), such that m:=MZ(A)=NZ(A). We now have the following theorem.

Theorem 4.1

(Müller’s theorem). Let A be an affine PI-algebra over a field K. Then ExtA1(M,N)MZ(A)=NZ(A).

Proof.

This is a reformulation of Müller’s theorem as stated in [Citation2, Theorem III.9.2] using [Citation2, Lemma I.16.2]. □

In other words, the “tangent spaces” ExtA1(M,N) are non-zero precisely over the ramification locus, ram(A).

We will primarily be interested in PI-algebras that are furthermore finite as modules over their center (or central subalgebra).

4.2 Some ring-theoretical properties

The following definition is one generalization of regularity to non-commutative rings. It is not necessary to understand the definition beyond knowing that this is a regularity property suitable for non-commutative algebraic geometry.

Definition 4.1.

  1. Let R be a ring and M an R-module. Then the grade of M is defined as j(M):=min{i|ExtRi(M,R)0}

or j(M)= if no such i exists.
  1. R is Auslander–Gorenstein if for every left and right Noetherian R-module M and for all i0 and all R-submodules NExtRi(M,R), we have j(N)i.

  2. R is Auslander-regular if it is Auslander–Gorenstein and has finite global dimension.

  3. Let R be a K-algebra, for K a field. Then R is Cohen–Macaulay (CM) if j(M)+GKdim(M)<,

for every R-module M. Here GKdim denotes Gelfand–Kirillov dimension with respect to K.

Note that when R is commutative we get the ordinary (Serre) regularity as defined in commutative algebra.

For this section we can work slightly more generally and assume that B is an admissible k-algebra. Recall that an admissible (commutative) ring (or scheme) is a ring which is of finite type over a field or excellent Dedekind domain. We assume in addition that B is a regular domain. Essentially everything can be made global if one is careful, but for simplicity we only consider the situation over a fixed affine patch. The crucial difficulty arises when considering viewing Ore extensions in a global setting.

We put W:=B{e¯0,e¯1,,e¯n1}(e¯ie¯jqi1qje¯je¯ik=0n1(1qi1qj)aijke¯k)and give W the standard ascending filtration by degree with Fil0:=B and the generators {e¯i} in degree one. It is important to notice that, if we globalize W, we find that (3.5) is a special case.

Recall that an Ore extension (or skew-polynomial ring) of a ring A is a twisted polynomial ring A[x;σ,δ], where σ is a ring morphism on A and δ a σ-derivation on A, twisted in the sense that xa=σ(a)x+δ(a).

Consider now the iterated Ore extension (4.1) R:=B[y0][y1;ϕ1][yn1;ϕn1],withϕi(yj)=qijyj,0j<i,(4.1) where we have put qij:=qi1qj and where qiB× for all 0in1. Notice that R is in fact gr(W) with W given the standard filtration and yi=gr(e¯i).

Proposition 4.2.

Assume that the qi’s are primitive mi-th roots of unity. Then,

  1. gr(W) is a noetherian, Auslander-regular domain;

  2. Kdim(gr(W))=Kdim(B)+n;

  3. gl.dim(gr(W))=gl.dim(B)+n;

  4. we have the central subalgebra

    B[y0N,y1N,,yn1N]Z(R)=Z(gr(W)),

    where N is the least common multiple of the mi;

  5. R=gr(W) is finite as a module over its center and hence a polynomial identity (PI) algebra, and

  6. R is a maximal order in its quotient ring of fractions, which is a division algebra.

Proof.

By [Citation6] an iterated Ore extension of a noetherian regular domain is a noetherian Auslander-regular domain, proving (i). The next two statements follow from [Citation18, Theorem 7.5.3] and [Citation18, Proposition 6.5.4], respectively. Clearly, B[y0N,y1N,,yn1N]Z(R). The whole ring R is finite as a module over B[y0N,y1N,,yn1N], since any monomial y0l0y1l1yn1ln1 can be written as y0l0y1l1yn1ln1=(y0l0Ns0y1l1Ns1yn1ln1Nsn1)·y0Ns0y1Ns1yn1Nsn1,with s0,s1,sn10 and each liNsi<N. Therefore, R is finite over Z(R). From this follows that R is PI by [Citation18, Corollary 13.1.13(iii)]. The last claim, that R is a maximal order in its quotient ring of fractions, follows from [Citation17, Proposition V.2.3] since R is prime (every Ore extension over a domain is a domain, hence prime). This is a division ring since R is an Ore domain by [Citation18, 2.1.15] together with (i) above, and the claim then follows by [Citation18, 2.1.14]. □

Proposition 4.3.

The B-algebra W is an Auslander-regular, noetherian PI-domain. Consequently the ring of fractions is a division algebra.

Proof.

Notice that the monomials in e¯0,e¯1,,e¯n1 form a basis for W as a B-module and the relations between the e¯i’s are on the form e¯ie¯jqije¯je¯i=k=0n1aijke¯k.

Furthermore, for fB, we have that e¯if=fe¯i, for all i. Then [Citation9, Theorem 1 and Corollary 2] implies that W is Auslander-regular (at this point we could also have used that gr(W) is Auslander-regular). By assumption B is a noetherian domain. From this, and the fact that the standard filtration is separated, follows that gr(W) is a noetherian domain if and only if W is a noetherian domain. Since gr(W) is an iterated Ore extension of a noetherian domain it is itself a noetherian domain. To prove that W is a PI-algebra, we note that W is finite as a module over the commutative subalgebra B[e¯1N,e¯2N,,e¯n1N]W. Hence we can conclude by [Citation18, Corollary 13.1.3(iii)]. That the ring of fractions is a division algebra follows as in the proof of Proposition 4.2. □

Remark 4.1.

The B-algebras W are not Ore extensions in general.

We will not use the the following two results but I include them for interest’s sake. Let T be a commutative ring, A an T-algebra and M a finitely generated A-module. Then M is called generically free if there is a non zero-divisor sA such M[s1]:=MAA[s1] is free.

Proposition 4.4.

Let B be an admissible domain and endow W with the standard filtration (with generators in degree one). Then every finite W-module is generically free.

Proof.

We have seen that gr(W) is an Ore extension and by [Citation1, Proposition 4.4], gr(W) is strongly noetherian. We can now apply [Citation1, Theorem 0.3] to conclude. □

For a ring B, K0(B) is the Grothendieck group of projective B-modules, and Pic(B) is the Picard group, i.e., the group of locally free B-modules of rank one.

Proposition 4.5.

We have K0(W)K0(B),

and if B=oK, the ring of integers in a number field K, we have K0(W)K0(oK)Pic(oK)Z.

Proof.

We filter W with the standard filtration with Fil0=B and generators in degree one. The associated graded gr(W) is flat over B since it is an (iterated) Ore extension of B. Since W is Auslander-regular the global dimension is finite. This implies that every cyclic W-module has finite projective dimension (e.g., [Citation18, 7.1.8]) and so is right regular. Therefore, the first isomorphism now follows from Quillen’s theorem [18, Theorem 12.6.13].

The isomorphism K0(oK)=Pic(oK)Z, comes from the Chern character (see [Citation19, III.6], for instance): ch:K0(oK)Pic(oK)Z,Edet(E)rk(E).

Therefore, K0(W)Pic(oK)Z. □

Notice that since oK is the ring of integers in K, Pic(oK), is nothing but the class group of K.

Corollary 4.6.

The morphism K0(Z(W))K0(W) induces a group morphism K0(Z(W))Pic(oK)Z,via K0(oK).

Remark 4.2.

In order to explicitly transfer projectives between W and B, it would be interesting to know an explicit isomorphism between K0(W) and K0(B).

We say that a B-algebra T is fiber-wise Cohen–Macaulay if TBk(p) is Cohen–Macaulay for all pSpec(B).

Put W/p:=WBk(p), pSpec(B). Assume that q¯i0, i.e., that qip. Reducing the ring R from (4.1) modulo p gives R/p:=k(p)[y¯0][y¯1;ϕ¯1][y¯n1;ϕ¯n1]with ϕ¯i(y¯j)=q¯ijy¯j, q¯ij:=q¯i1q¯j. Giving W/p the standard filtration with Fil0(W/p)=k(p) and all generators in degree one, we find R/p=gr(W/p).

Proposition 4.7.

The following holds:

  1. GKdim(gr(W/p))=Kdim(gr(W/p))=gl.dim(gr(W/p))=n;

  2. GKdim(W/p)=GKdim(gr(W/p));

  3. tr.deg(Z(gr(W/p)))=n;

  4. W is fiber-wise Cohen–Macaulay;

  5. gr(W/p) is a maximal order in its division ring of fractions, and

  6. Z(gr(W/p)) is an integrally closed domain.

Proof.

The first three statements follow from [Citation18, 13.10.6] and [Citation18, Proposition 8.1.14]. Taking Λ=k(p)[y¯0] in [Citation9, Theorem 3] shows (iv) as the associated graded gr(Λ)=gr(k(p)[y¯0])=k(p)[y¯0],with k(p)[y¯0] given the standard filtration, is Cohen–Macaulay; (v) follows from [Citation17, Proposition V.2.3] again (the division ring claim follows as in Proposition 4.2). Finally, (vi) follows from [Citation18, Proposition 5.1.10 b(i)]. □

4.3 Kummer–Witt algebras

Recall the assumption that ζB, where ζ is a primitive n-th root of unity. We will consider a special case of the construction in Section 3.2 from which all else that follow will be built.

As in Section 3.2 let A be the cyclic ring extension (4.2) A:=B[t]/(tnx)=i=0n1Bei,(4.2) with ei:=ti and xB. Clearly, eiej=xe{i+j​​​  mod n}.

This means that A is a Kummer extension of B. The element xB is the A-ramification divisor. This element determines a canonical subscheme in the ramification locus of a non-commutative space attached to A. Observe that x is a ramification invariant in two senses: (1) as the divisor in B over which A is ramified (i.e., p|xp ramified); and (2) as an element giving a subscheme of the ramification locus in a certain non-commutative space.

Put Wx1/n(r):=E(WA1/n(r)). Explicitly this means that (4.3) Wx1/n(r)=B{e¯0,e¯1,,e¯n1}(e¯ie¯jζr(ji)e¯je¯i(1ζr(ji))x e¯{i+j​​​  mod n}).(4.3)

If x = 0, the algebra W01/n(r) is the enveloping algebra of an infinitesimal hom-Lie algebra over B of order n and so XW01/n(r) can be viewed as a non-commutative fat point of order n.

Observe that Wx1/n(r) is a special case of the algebra W in the previous section.

Remark 4.3.

The above construction globalizes immediately. Let x:=(Ui,xi) be a Cartier divisor on a scheme X over a base including a ζn. Then we define an OX-algebra Wx1n(r) by giving it locally by (4.3), with x replaced by xi. In order to keep it simple, we have opted to only write out the affine case. Every algebra, when given by generators and relations, in this paper can be globalized. Be aware, however, that the isomorphism in Proposition 4.10 cannot be sheafified.

4.3.1 Fibers of Wx1/n(r)

We begin by observing that the center commutes with taking fibers Z(Wx1/n(r)Bk(p))=Z(Wx1/n(r))Bk(p).

This follows since, if a1Z(Wx1/n(r)Bk(p)) then, for x1Wx1/n(r)Bk(p), (a1)(x1)=ax1and(x1)(a1)=xa1so that ax = xa. The other inclusion follows since any element in Z(Wx1/n(r))Bk(p) commutes with any element in Wx1/nk(p).

Let pSpec(B) be a prime. Observe that the reduction of ζ modulo p (i.e., the image of ζ in k(p)), ζ¯, is non-zero, since ζB×. Then (4.4) Wx1/n(r)/p=k(p){e¯0,e¯1,,e¯n1}(e¯ie¯jζ¯r(ji)e¯je¯i(1ζ¯r(ji))x e¯{i+j​​​  mod n}).(4.4)

We record the following for easy reference.

Proposition 4.8.

We have the following three possibilities when reducing modulo a prime p:

  1. ζ¯=1, in which case we get

    Wx1/n(r)/p=k(p){e¯0,e¯1,,e¯n1}(e¯ie¯je¯je¯i)=k(p)[e¯0,e¯1,,e¯n1],

    the commutative polynomial algebra;

  2. x¯=0, in which case we get

    Wx1/n(r)/p=k(p){e¯0,e¯1,,e¯n1}(e¯ie¯jζ¯r(ji)e¯je¯i),

    a quantum affine space;

  3. and the generic case (4.4) with relations unchanged.

    It is important to note that in all three cases, the reduced algebra is a domain (see Proposition 4.2).

Proof.

Obvious. □

The ring-theoretic properties of Wx1/n(r) are summarized in the following theorem.

Theorem 4.9.

The algebra Wx1/n(r) satisfies the following:

  1. it is an Auslander-regular, noetherian PI-domain, finite over its center;

  2. Kdim(Wx1/n(r))=gl.dim(Wx1/n(r))=n+Kdim(B);

  3. it is fiber-wise Cohen–Macaulay with

    GKdim(Wx1/n(r)/p)=tr.deg(Wx1/n(r)/p)=n;

  4. it is fiber-wise a maximal order in its fiber-wise division rings of fractions;

  5. every finitely generated Wx1/n(r)-module is generically free, and

  6. K0(Wx1/n(r))K0(B).

Proof.

The point (i) follows from Proposition 4.3. By definition, tr.deg of a PI-algebra S over a field is tr.deg(Z(S)). From [Citation18, Proposition 13.10.6] we have Kdim(Wx1/n(r)/p)=tr.deg(Wx1/n(r)/p)=GKdim(Wx1/n(r)/p).

Since k(p)[e¯0n,e¯2n,,e¯n1n]Z(Wx1/n(r)/p), we have that tr.deg(Z(Wx1/n(r)/p))Z(k(p)[e¯0n,e¯2n,,e¯n1n])=n,implying that Kdim(Wx1/n(r)/p)n. For a filtered ring S we have by [Citation18, Lemma 6.5.6] that Kdim(S)Kdim(gr(S)), and by Proposition 4.7 we have that Kdim(gr(Wx1/n(r)/p))=n, so Kdim(Wx1/n(r)/p)=n. The same applies to GKdim(Wx1/n(r)/p)=n. Lifting to B we get Kdim(Wx1/n(r)/p)=n+Kdim(B). By [Citation18, Corollary 7.6.18], gl.dim(Wx1/n(r)/p)gl.dim(gr(Wx1/n(r)/p)). For noetherian prime PI-rings S we have that Kdim(S)gl.dim(S) by [Citation20, Theorem 1.7(i)], so n=Kdim(Wx1/n(r)/p)gl.dim(Wx1/n(r)/p)gl.dim(gr(Wx1/n(r)/p))=n, proving (ii) and (iii). The rest follows directly from Proposition 4.7 as Wx1/n(r) is a special case of W from that section. □

Remark 4.4.

A natural question is to what extent the above theorem, and the results that will follow, can be extended to other ring extensions, besides Kummer extensions.

4.3.2 Transfer of structure

It is obviously interesting to transfer structures such as modules or subspaces of A/B to corresponding structures over Wx1/n(r). The following easy observation allows us to do just that. Write Wx1/n(r) as B{e¯0,e¯1,,e¯n1}/I, with I the two-sided ideal of relations in B{e¯0,e¯1,,e¯n1} from (4.3). We can construct a B-module morphism ξ:AWx1/n(r),ei=tie¯i,0in1.

Then, if SA is a subset, we can transfer S to Wx1/n(r) via the map ξ, to get the subset ξ(S)Wx1/n(r). In particular, if S generates an ideal in A, ξ(S) generates a two-sided ideal in Wx1/n(r), and we can consider the quotient Wx1/n(r)/ξ(S).

Similarly, if M is an A-module over B, we have an action of ei on M via some structure morphism ρ:AEndB(M). Via the association ξ we can transfer the action of ei on M to e¯i to get a morphism B{e¯0,e¯1,,e¯n1}ρEndB(M),ρ(e¯i)=ρ(ξ1(ei)):=ρ(ei).

Taking the invariants of M under ρ(I) we get a Wx1/n(r)-module via the induced structure morphism χ:Wx1/n(r)EndB(Mρ(I)).

4.4 The algebra Jackson algebra Jx(r)

When n > 2 there is a canonical subalgebra of Wx1/n(r) that we will now study in some detail. Let me remark that at some points one needs to be a bit careful when the characteristic is two.

First, notice that since ζn=1, we have that ζ(n1)=ζ and ζ(n2)=ζ2. We put Jx(r):=B{e¯0,e¯1,e¯n1}(e¯0e¯1ζre¯1e¯0(1ζr)e¯1e¯n1e¯0ζre¯0e¯n1(1ζr)e¯n1e¯n1e¯1ζ2re¯1e¯n1x(1ζ2r)e¯0).

This is clearly a subalgebra of Wx1/n(r). However, it is more beneficial to work with an isomorphic algebra:

Proposition 4.10.

Let ζ2r1. Then the algebra Jx(r) is isomorphic over B[(1ζ2r)1], in particular fibre-wise, to the algebra (4.5) Jx(r):=B{e¯0,e¯1,e¯2}(e¯0e¯1ζre¯1e¯0,e¯2e¯0ζre¯0e¯2,e¯2e¯1ζ2re¯1e¯2xe¯0x(1ζ2r))(4.5)

and Jx(r) is an iterated Ore extension.

Proof.

By changing basis e¯0(1ζ2r)1e¯0+1 we can transform the relations for Jx(r) to the ones in (4.5). Construct the iterated Ore extension B[e¯0][e¯1;τ][e¯2;τ1,δ], with τ(e¯0)=ζre¯0,τ1(e¯1)=ζ2re¯1,δ(e¯1)=xe¯0+x(1ζ2r),δ(e¯0)=0.

It is easy to see that this Ore extension is isomorphic to Jx(r). Notice that we have implicitly extended τ to e¯1 in the proof. □

The algebra Jx(r) is isomorphic to the (enveloping algebra of the) “Jackson-sl2”, which is a q-deformation of the Lie algebra sl2, from [Citation16]. Therefore the following definition is natural.

Definition 4.2.

We call the algebra Jx(r) the Jackson algebra (of level r) associated with the cover Spec(A)Spec(B). If r is irrelevant for the discussion, we will often drop it from the notation.

Remark 4.5.

We make the following (long) series of remarks.

  1. If n = 2 the algebra Wx1/n(r) only have two generators, so Jx(r) cannot be a subalgebra in this case. Still, abstractly, it is well-defined as given by generators and relations.

  2. When ζ2r=1, the algebra Jx(r) is either isomorphic to the commutative polynomial algebra B[t1,t2,t3] (when ζ = 1) or to the B-algebra on generators e¯0, e¯1, e¯2 and with relations

    e¯0e¯1+e¯1e¯0=0,e¯2e¯0+e¯0e¯2=0,e¯2e¯1e¯1e¯2=xe¯0.

    However, in this case Jx(r) and Jx(r) are not isomorphic.

  3. The algebra defined by (4.5) is isomorphic to the down-up algebra Dζr over B (see [Citation3] for the definition) defined by the relations

    d2u=ζr(1+ζr)dudζ3rud2+a(1ζ2r)(1ζr)d,du2=ζr(1+ζr)uduζ3ru2d+a(1ζ2r)(1ζr)u.

    To see this, solve for ae¯0 in (4.5) and insert in the other two relations and simplify.

  4. If x = 0, we get, after a renaming of generators, the same relations as in (3.2) and hence a quantum A3.

  5. The algebras Wx1/n(r) are not Ore extensions in general.

  6. Notice that Jx(r) includes two copies of the quantum plane QB(r):=B{t1,t2}/(t1t2ζrt2t1). Hence Jx(r) is constructed by “glueing” the quantum planes via the third relation in (4.5).

  7. The algebra Jx/(e¯0) is the first quantum Weyl algebra.

The following deserves its own remark:

Remark 4.6.

Since A/B is a Kummer extension and the ramification properties of such extensions are intimately related to the divisor x, it is natural to assume that the ramification is related to the algebra Jx. This is indeed the case, and one of the reasons I started this project. It will be a fundamental part of a sequel to the present paper to study this connection in more depth.

We give the relations in the cases n = 3 and n = 4.

Example 4.1.

When n = 3 we get r=0:{e¯0e¯1e¯1e¯0=0e¯2e¯0e¯0e¯2=0e¯2e¯1e¯1e¯2=xe¯0}r=1:{e¯0e¯1ζ3e¯1e¯0=0e¯2e¯0ζ3e¯0e¯2=0e¯2e¯1ζ32e¯1e¯2=xe¯0+x(1ζ32)}r=2:{e¯0e¯1ζ32e¯1e¯0=0e¯2e¯0ζ32e¯0e¯2=0e¯2e¯1ζ3e¯1e¯2=xe¯0+x(1ζ3)}

The different cases are non-isomorphic. Observe that the case r = 0 is in fact the universal enveloping algebra of a solvable 3-dimensional Lie algebra.

Example 4.2.

In a sense the case n = 4 is more interesting. Recall that ζ4 is chosen to be primitive. This means in particular that ζ42 is not primitive, hence the case r = 2 is rather special. r=1:{e¯0e¯1ζ4e¯1e¯0=0,e¯2e¯0ζ4e¯0e¯2=0,e¯2e¯1+e¯1e¯2=xe¯0+2x}r=2:{e¯0e¯1+e¯1e¯0=0,e¯2e¯0+e¯0e¯2=0,e¯2e¯1e¯1e¯2=xe¯0}r=3:{e¯0e¯1+ζ4e¯1e¯0=0,e¯2e¯0+ζ4e¯0e¯2=0,e¯2e¯1+e¯1e¯2=xe¯0+2x}.

Obviously, the case r = 0 is the same for all n.

4.5 Ring-theoretic and geometric properties of Jx(r)

4.5.1 The center

We will use a result of A. D. Bell and S.P. Smith from [Citation22]. Unfortunately, as far as I’m aware, this result is not publicly available so, for completeness, I include their proof. Therefore, except for Corollary 4.12, there is nothing original (apart for minor modifications) in the following section. Any mistakes are certainly my own.

To be consistent with the notation in Bell and Smith’s work we rearrange the last relation in Jx(r): (4.6) Jx(r)=k(p){e¯0,e¯1,e¯2}(e¯0e¯1ζre¯1e¯0,e¯2e¯0ζre¯0e¯2,e¯1e¯2ζ2re¯2e¯1ae¯0b),(4.6) where we have put a:=xζ2r and b:=xζ2r(1ζ2r)=x(1ζ2r), for simplicity.

Put w¯:=e¯1e¯2 and W:=k(p)[e¯0,w¯]. Then W is a commutative subalgebra of Jx(r) and Jx(r) is free as a module over W.

Define σ(e¯0)=ζre¯0and σ(w¯)=ζ2r(w¯ae¯0b)=e¯2e¯1.

Then σAut(W) (this is part of [Citation22, Lemma 3.2.2]). One sees immediately that (4.7) we¯1=e¯1σ(w)ande¯2w=σ(w)e¯2for all wW.(4.7)

Put Jx(r)n:=We¯2n=e¯2nW and Jx(r)n:=We¯1n=e¯1nW. Then Jx(r)=nZJx(r)n=Jx(r)Jx(r)+,

with Jx(r):=n<0Jx(r)n,Jx(r)+:=n0Jx(r)n.

One can easily show that Jx(r)+=W[e¯2;σ] and Jx(r)=W[e¯1;σ1] are both Ore extensions. The following proposition is part of proposition 3.2.4 in Bell and Smith [Citation22].

Proposition 4.11.

Put s:=ord(σ). Observe that s must be such that sr is a multiple of n by the definition of σ. Then Z(Jx(r))=Wσ[e¯1s,e¯2s],where Wσ is the invariant subring under σ.

Proof.

An element in Jx(r) is central if and only if every term is. Let wW and consider we¯2i. Then for wW we have, by (4.7), we¯2iw=σ(w)iwe¯2iand, since Jx is a domain, we¯2i commutes with all wW if and only if σi=id. We similarly see that [e¯2,we¯2i]σ(w)=w.

Suppose now that σi=id and that σ(w)=w. Then, since Jx is a domain, we¯2i commutes with e¯1 if and only if (4.8) e¯2e¯1we¯2i=e¯2we¯2ie¯1.(4.8)

Now, e¯2e¯1we¯2i=σ(w¯)we¯2i=(w¯)e¯2iand e¯2we¯2ie¯1=we¯2i+1e¯1=we¯2ie¯2e¯1=we¯2iσ(w¯)=wσi+1(w¯)e¯2i=(w¯)e¯2i.

Hence, (4.8) is proven and so we¯2iZ(Jx)σi=id and σ(w)=w.

The exact same reasoning applies to we¯1i, thereby completing the proof. □

From this follows:

Corollary 4.12.

We have, still with w¯=e¯1e¯2, Wσ={k(p)[e¯0l],if a,b0k(p)[e¯0l,w¯l],if a=b=0and so Z(Jx(r))={k(p)[e¯0l,e¯1l,e¯2l],if a,b0k(p)[e¯0l,e¯1l,e¯2l,w¯t],if a=b=0, where l is the least integer such that lr0 (mod n) and t minimal with the property that 2tr0 (mod n).

Observe that the congruences imply that 2trlr (mod n), i.e., n|(2tl)r.

Proof.

We begin by determining Wσ. First, that e¯0mWσ means that σ(e¯0m)=ζmre¯0m, so mr0 (mod n). We can thus assume that m = l, the least such integer such that lr0 (mod n). An induction argument shows that σk(w¯)=ζ2krw¯aζ(k+1)r[k]ζre¯0ζ2r[k]ζ2rb,

and from this, together with σk(e¯0)=ζkre¯0, follows that the order of σ must be l. Let e¯0lw¯tW. Then, σ(e¯0lw¯t)=σ(e¯0)lσ(w¯)t=ζlre¯0l(ζ2r(w¯ae¯0b))t=ζ(2t+l)re¯0l(w¯ae¯0b)t.

If this element shall be invariant under σ we must thus have ζ(2t+l)re¯0l(w¯ae¯0b)t=e¯0lw¯t,which, since W is a domain, is equivalent to (4.9) ζ(2t+l)r(w¯ae¯0b)t=w¯t.(4.9)

The trinomial identity allows us to expand the parentheses in the left-hand-side to obtain the condition ζ(2t+l)ri+j+k=t(1)j+kt!i!j!k!ajbke¯0jw¯i=w¯t.

From this follows that if a,b0, this can never occur unless t = 0.

If a=b=0, then (4.9) implies that ζ(2t+l)rw¯t=w¯t. Since lr0 (mod n) we see that we must have 2tr0 (mod n). Take, t minimal (possibly zero) with this property. Then Proposition 4.11 shows the claim concerning the center, thereby completing the proof. □

4.5.2 Algebraic geometry of Jx(r)

Theorem 4.13.

The center of Jx(r) is given fibre-wise as follows.

  1. For x = 0:

    Z(J0(r))/p=k(p)[e¯0l,e¯1l,e¯2l,w¯t],

    where l is the least integer such that lr0 (mod n) and t minimal with the property that 2tr0 (mod n), i.e., t is minimal such that 2tr is a multiple of n.

  2. For x0:

    Z(Jx(r))/p=k(p)[e¯0l,e¯1l,e¯2l]

    with lN as in (i).

    In both cases we have

  3. Jx(r) is an Auslander-regular, noetherian, fibre-wise Cohen–Macaulay domain, finite as a module over its center and hence a polynomial identity ring (PI) of pideg(Jx(r))=n;

  4. Jx(r) is a maximal order in its division ring of fractions;

  5. K0(Jx(r))K0(B);

  6. Spec(Z(Jx(r))) is a normal, irreducible scheme of dimension three, for all x.

  7. (vii) Spec(Z(Jx(r))) is in addition fibre-wise Cohen–Macaulay in the commutative sense, i.e.,

    Spec(Z(Jx(r)/p))=Spec(Z(Jx(r))Bk(p))

    is a Cohen–Macaulay scheme for all pSpec(B) and all x.

Proof.

Properties (i) and (ii) is included in corollary (4.12).

Continuing with (iii), the same argument as in the proof of Proposition 4.2 shows that Jx(r) is finite over its center, and from this follows that Jx(r) is PI by [Citation18, Corollary 13.1.13(iii)] and [Citation17, Proposition V.2.3] once again shows that it is a maximal order in its division ring of fractions (as in Proposition 4.2 again). Since Jx(r) is an iterated Ore extension over a noetherian domain, it is Auslander-regular and a noetherian domain itself.

As for the pi-degree, it looks at a first glance like the rank of Jx(r) over its center is n3, which it obviously cannot be (it must be a square). On the other hand, on closer inspection of the relations involved, we see that e¯0 can be eliminated from Jx(r) over Z(Jx(r)). Hence we effectively only have two generators, and so pideg(Jx(r))=n also in this case.

Part (v) follows by the same argument as in Proposition 4.5 and part (vi) follows from [Citation18, 5.1.10 b(i)] and (iv). That the dimension is three is a consequence of [Citation18, Proposition 13.10.6]. Finally, when x = 0 the claim concerning the Cohen–Macaulayness in (vii) follows from [Citation5, Lemma 2.2] and the case x0 is obvious by (ii). □

Corollary 4.14.

With the notation as above:

  1. The algebra Jx(r) is finite as a module over k(p)[e¯0l,e¯1l,e¯2l].

  2. Hence, the ring extension k(p)[e¯0l,e¯1l,e¯2l]Z(Jx(r)) is finite, and consequently the morphism

    ψ:Spec(Z(Jx(r)))A(l)3:=Spec(k(p)[e¯0l,e¯1l,e¯2l])

    is finite as a morphism of schemes. We put l in the notation to indicate that we have a weighted version of the affine three-space.

  3. Any maximal ideal M in Jx(r) intersects the center uniquely at a maximal ideal m.

  4. In the other direction, any maximal m in k(p)[e¯0l,e¯1l,e¯2l] splits into i maximal ideals in Jx(r), where 1im and where m is the rank of Jx(r) as a module over k(p)[e¯0l,e¯1l,e¯2l].

Proof.

The first claim is clear from the theorem and the rest then follows from (i). □

It is in fact quite easy to find an explicit presentation of the scheme in Theorem 4.13 (i).

Theorem 4.15.

Put ui:=e¯il for 0i2, and u3:=w¯t. Then Z(J0(r))/p=k(p)[u0,u1,u2,u3]/(u3ru1au2a),where a is the minimal integer such that tr = an. From this also follows that t = l and that r and a cannot both include a factor 2. The integer a is uniquely determined by n and r.

Proof.

A simple induction argument shows that w¯k=ζk(k1)re¯1ke¯2k, for all kN. Recall that t is the minimal integer such that 2tr0 (mod n)2tr=sn, for some s. We must have that s=2a for some a, since otherwise u1 and u2 are not defined (we lose commutativity). Therefore tr = an and so, w¯tr=ζtr(tr1)re¯1tre¯2tr(w¯t)r=ζan(an1)e¯1ane¯2anu3r=u1au2a.

This also proves that t = l by minimality of t and that r and a cannot both include a factor 2, since otherwise the center would not be a domain. □

It is now easy to convince oneself of the validity of the following corollary:

Corollary 4.16.

Let Xr be the family of affine surfaces Xr:=Spec(k(p)[u1,u2,u3](u3ru1au2a)).

Then Spec(Z(J0(r))/p)=A1×Xr.

The surface Xr furthermore satisfies:

  1. X1 is a regular and rational.

  2. Xr, for r > 1, are ramified r-covers of A2, with branch locus the coordinate axes u1=u2=0, and singular at the origin.

Observe that Spec(Z(J0(r))/p) is a trivial A1-fibration for all r and that a is uniquely determined by r and n.

We can also prove that the singularities in the above corollary are rational:

Theorem 4.17.

Let K be a field of characteristic zero and let Jx(r)/Kal be the base change Jx(r)/KKKal of Jx(r)/K to the algebraic closure Kal. Then Spec(Z(Jx(r)/Kal)) has rational singularities for all x. Hence Jx(r) has rational singularities on the generic fibre over Spec(B) (recall that B is a domain).

Proof.

When x0 the center is A3 (with some weight) by Theorem 4.13 (ii) so we can assume that x = 0.

We know that Jx(r)/Kal is Auslander-regular and Cohen–Macaulay by Theorem 4.13(iii). This implies that Jx(r)/Kal is homologically homogeneous (which we won’t define here) by [Citation24, Corollary 3.8(ii) and Note 3.5(i)]. Note that for affine pi-algebras, being Macaulay (as is discussed in this reference) is equivalent to being Cohen–Macaulay, and being Macaulay implies being locally Macaulay.

Now the main theorem of [Citation23] implies that Z(Jx(r)/Kal) has rational singularities. □

There’s probably an easier way to show rationality, simply by looking at the coordinate ring. Also, it could be that the singularities are rational even before going to the algebraic closure.

5 Geometry of XJxA(l)3

Even though we haven’t formally defined what should be meant by a “non-commutative scheme”, we will use the language of schemes in what follows. The following is a “soft” definition of a non-commutative scheme. For the rigorous definition see [Citation15].

Let A be a B-algebra, where B is a commutative ring. Then we define the non-commutative scheme or non-commutative space of A to be XA:=(Mod(A),OA),OA:=A.

We often identify M with its annihilator.

Let M:={M1,M2,,Mr} be a family of A-modules. Then the tangent space of XA at M is the collection of Ext1-groups TM:={ExtA1(Mi,Mj)|1i,jr}.

The tangent space TM controls the simultaneous non-commutative deformations of the modules, as a family.

The ring object OA, which we simply have put equal to A here, is actually built from TM via matric Massey products, by taking the projective limits over all families of A-modules. See [Citation7] or [Citation15]. For simplicity we view OA as a global object rather than the as a local object which it actually is. Hence the assignment OA:=A.

Definition 5.1.

Suppose A is a K-algebra with K a field, L/K a field extension and M an L-vector space. Then an L-rational point on XA is a K-linear algebra morphism ρ:AEndL(M) such that kerρ is a maximal ideal. The set of all L-rational points are denoted XA(L).

5.1 A family of divisors of degree 2

Let B be a commutative Dedekind domain with ζ:=ζnB.

Proposition 5.1.

Put a:=(a0,a1,a2)A/B3.

  1. The elements

    Ωa:=a0e¯2e¯1a1e¯1e¯2+a2e¯02xa0a1ζ11ζre¯0x(a0a1),

    defines a family of normal elements in Jx(r), parameterized by A3.

  2. In fact, Ωa defines an automorphism γ of Jx(r) as Ωa·t=γ(t)·Ωa with

    γ(e¯0)=e¯0,γ(e¯1)=ζ2re¯1,γ(e¯2)=ζ2re¯2.

  3. Hence, we can view Ωa as defining a flat family of non-commutative quadric surfaces

    Ya:=XJx(r)/Ωa

    embedded in XJx(r) and parameterized by aA3.

  4. The spaces Ya (i.e., the algebras Jx(r)/Ωa) are Auslander-regular for all aA3.

Proof.

Straightforward (but long) computations show (a) and (b) and then point (c) follows directly from definition. It is well-known that a quotient of an Auslander-regular algebra by a normal element is Auslander-regular, showing (d). □

The intersection with the center, i.e., the image of ϕ, is given by the (commutative) quadric family Z(Ωa)=(a0a1)u1u2+a2u02xa0a1ζ11ζru0x(a0a1),

with u0:=e¯0l, u1:=e¯1l and u2:=e¯2l, l being the least integer such that lr = n. This means that we have a fibration ϕa:=ϕ|Ya:YaSpec(Z(Jx(r))/Z(Ωa))Spec(B).

Notice that Pa:=Spec(Z(Jx(r))/Z(Ωa)) is an affine quadric, and Ya is an non-commutative space over this surface, defined by the order Jx(r)/Ωa. The quadric Pa is the “commutative shadow” of Ya.

Remark 5.1.

If Ωa and Ωa are two normal elements in the above family, then ΩaΩa is also a normal element. Therefore, it is possible to construct more complicated subspaces (of higher degree) from elements in the family Ωa.

5.2 Fibers over A(n)3

5.2.1 Geometric and arithmetic fibers

There is an important distinction to make concerning the notion of “fibre” in the non-commutative context. Let A be an R-algebra with RZ(A). Contraction of prime ideals defines a morphism α:XASpec(R).

Let pSpec(R). We can then talk of the “fibre of α over p” as α1(p):=XARk(p).

In the context of PI-algebras there are two types of fibers depending on whether pazu(A) or pram(A). If pazu(A), then A/p:=ARk(p) is a central simple algebra over k(p) and so XA/p is one simple module as, both A/p-module and A-module.

On the other hand, if pram(A), the “fibre algebra” A/p is not central simple. However, it is certainly artinian so A/p/rad(A/p), where rad is the Jacobson radical, is semi-simple and A/p and A/p/rad(A/p) have the same simple modules. The point here is that, even if pram(A), it can happen that A/p/rad(A/p), and hence also A/p, only has one simple module. This apparent contradiction is resolved by being careful what algebra we mean: there is only one simple module of A/p as A/p-module, but as A-module there are more.

Therefore, when we speak of “fibre” when need to be careful what we mean: do we mean the points of α1(p) as A-modules or as A/p-modules. We say that α1(p)=XA/p is the geometric fibre at p, i.e., we view α1(p) as a set of A/p-modules; the set Φ(p):={PXA|PR=p}is the arithmetic fibre. By the fibre of A at p, we mean the algebra A/p=ARk(p).

5.2.2 Ramification over a central subalgebra

Suppose now that A is a PI-algebra, finite over a central subring R and consider the inclusions RfZ(A)gA.

We say that A is ramified over R if either f is a ramified map (i.e., the extension RZ(A) is ramified) or ram(A) (or both). The ramification locus of g°f is ramR(A):=ram(g°f):=branch(f)(ram(A)R).

The complement of ram(g°f) is the azumaya locus, azuR(A):=azu(g°f), of g°f. Here we see that slightly unfortunate clash between the algebraic-geometric and the PI-theoretic notions of ramification.

Notice that, even if pazuR(A) there can be more than one prime of A over p since the p might split in Z(A).

5.2.3 The fibers

For the rest of this section we work over a field B = K. Recall the assumption that ζK is primitive. It is important to note that we, to simplify the discussion, now only consider the case r = 1. Hence l = n and we write Jx:=Jx(1)=K{e¯0,e¯1,e¯2}(e¯0e¯1ζe¯1e¯0,e¯2e¯0ζe¯0e¯2,e¯2e¯1ζ2e¯1e¯2xe¯0x(1ζ2)).

The characteristic p of K is arbitrary.

We begin by recalling that a symbol algebra or cyclic algebra (see e.g., [Citation8, Corollary 2.5.5]) over a field K (or a commutative ring) is an algebra (a,b)ξ with generators x and y such that xyξyx=0andxn=a,yn=b where a,bK×,ξn=1,with n invertible in K (in particular np). Symbol algebras are central simple algebras.

The actual construction goes as follows. Let L be the cyclic extension np. Since L is cyclic over K, there is σGal(L/K) such that σ(t)=ξt. Then, for cK, (a,c)ξ is the L-vector space (5.1) (a,c)σ=(a,c)ξ:=i=0n1Lei,with en=c and et=σ(t)e=ξte.(5.1)

Clearly this is the quotient of the Ore extension (K[t]/(tna))[e;σ] by the central element enc.

From [Citation8, Exercise 4.10] we have that (a,b)ξ(a,c)ξb1cNm(L),where Nm is the ordinary norm function Nm:LK. It also turns out that L is a splitting field for (a,b)ξ, i.e., (a,b)ξKLMr(L) for some r1.

Put Λ:=K[u0,u1,u2]=K[e¯0n,e¯1n,e¯2n],Z:=Spec(Z(Jx)),A(n)3:=Spec(Λ)

and denote by ψ the scheme morphism ψ:ZA(n)3induced by restriction of primes. Remember that ψ might actually be the identity in some cases, depending on Z.

Pick a maximal ideal m=m(a,b,c)=(u0a,u1b,u2c),a,b,ck(m),in Λ. The fibre (as algebra) of Jx over m is Jm:=Jx/m=JxΛk(m)=Jx(e¯0na,e¯1nb,e¯2nc).

Recall from Remark 4.5(6) that Jx includes two copies of the quantum plane QK=QK(1). Recall also that QK is independent on x. It is well-known that the center of QK is K[xn,yn] (if we use x and y as generators for QK) and that ram(QK) is the union of the coordinate axes. Hence reduction modulo a maximal ideal mMax(Z(QK)) gives an algebra of the type (5.2) (a,b)ζ:=QKΛk(m)=k(m){x,y}(xna,ynb,xyζxy).(5.2)

In fact we can consider this from another angle. Indeed, we can view (a,b)ζ as the fibers over azu(QK) of the inclusion Z(QK)QK. Over ram(QK), the fibers (a,b)ζ become non-commutative fat points (and certainly not central simple).

Consider the following diagram of inclusions of algebras

where Q1 and Q2 are the quantum planes Q1:=k(m){e¯0,e¯1}(e¯0e¯1ζe¯1e¯0)andQ2:=k(m){e¯0,e¯2}(e¯2e¯0ζe¯0e¯2)inside Jx.

The point is that this diagram parameterizes cyclic algebras over K and Brauer classes. Furthermore, this includes a non-commutative family of central simple algebras.

First of all, we have seen in (5.2) that the center Z(QK) of a quantum plane parameterizes cyclic algebras by its very construction. The previous two propositions give information on this parameterization by extending the centers Z(Q1) and Z(Q2) to the central algebra Λ.

Proposition 5.2.

Let m be the maximal ideal m:=m(a,b,c)azuΛ(Jx),with a,b,ck(m) such that abc0,and put L:=k(m)[t]/(tna). Then the following statements hold.

  1. There are at least two symbol algebras (a,b)ζ and (a,c)ζ in Jm. These are non-isomorphic unless c/bNm(L).

  2. In addition:

    1. if c/bNm(L), then Jm generates two classes in Br(k(m))[n];

    2. if c/bNm(L), Jm generates one class in Br(k(m))[n].

  3. Let MMax(Jx) be such that m:=MΛazuΛ(Jx) and assume that k(MZ(Jx))=k(m). Then there are two canonical central simple algebras

    (a,b)ζ,/k(m)k(m)(a,c)ζ,/k(m)(a,bc)ζ,/k(m)Br(k(m))[n](5.3)

    and

    (a,b)ζ,/k(m)k(m)(a,c)ζ,/k(m)k(m)Jx/M(a,bc)ζ,/k(m)k(m)Jx/MBr(k(m))[lcm(n,per(Jx/M))],(5.4)

    where per(A) denotes the period (i.e., the order of the class of A in Br(K)) of the K-central simple algebra A. This implies that Jx parameterizes cyclic algebras and Brauer classes in Br(k(m)).

Proof.

  1. We observe that there are two symbol algebras as subalgebras inside Jm. Namely, (a,b)ζ and (a,c)ζ1:

    (a,b)ζ=k(m){e¯0,e¯1}(e¯0na,e¯1nb,e¯0e¯1ζe¯1e¯0),(a,c)ζ1=k(m){e¯0,e¯2}(e¯0na,e¯2nc,e¯2e¯0ζe¯0e¯2).

    Since (a,c)ζ1 is the symbol algebra constructed as in (5.1), but with

    et=σ1(t)e=σn1(t)e,

    we find that (a,c)ζ1(a,c)ζ. Indeed, if gcd(n,k)=1, then (a,c)σ(a,c)σk, by a change of basis. Because gcd(n,n1)=1, the claim then follows.

    The reason for the addition of “at least” in the claim, is that, as we remarked, even if mazuΛ(Jx), there might be more than one fibre of Jx over m depending on the splitting of m inside Z(Jx).

  2. The claim follows from the above discussion of symbol algebras above, and from the equivalence

    AB[A]=[B]Br(F)and dimF(A)=dimF(B),

    for simple F-algebras A and B.

  3. Take a maximal ideal M in Jx and assume that k(MZ(Jx))=k(m), with m:=MΛ. Then Jx/M is a central simple algebra over k(m). Let k1:=k(mZ(Q1)) and k2:=k(mZ(Q2)). Then k1,k2k(m) and we can extend the fibers to k(m):

    (Q1Z(Q1)k1)k1k(m)=(a,b)ζ,/k(m)

    and similarly with Q2 to get (a,c)ζ,/k(m). Observe that the extension to k(m) might split (a,b)ζ,/k(m) or (a,c)ζ,/k(m) (or both). It is well-known that (a,b)(a,c)(a,bc) (e.g., [Citation8, Ex. 4.10(a)]) and that cyclic algebras are n-torsion in the Brauer group. Hence (5.3) follows. Since Jx/M is central simple over k(m) and has order per(Jx/M) the tensor product is lcm(n,per(Jx/M))-torsion in Br(k(m)).

The proof is complete. □

Turning now to the case when abc = 0, we have already found that the algebra Jm is not simple and so mramΛ(Jx). Therefore we see that {u0=0}{u1=0}{u2=0}ramΛ(Jx).

For instance, over the {u0=0}-part of ramΛ(Jx), we find Jm=JxΛk(m)=Jx(e¯0n,e¯1nb,e¯2nc),bc0,(e¯0)rad(Jm),since the radical is the largest nilpotent ideal and e¯0n=0. Hence, if x0, Jm/(e¯0)=k(m){e¯1,e¯2}(e¯1nb,e¯2nc,e¯2e¯1ζ2e¯1e¯2(1ζ2)x), and if x = 0 we get the cyclic algebra (b,c)ζ2.

If ζ21 and x0, we can change basis e¯1(1ζ2)xe¯1 and see that Jm/(e¯0) is isomorphic to the quotient of the first quantum Weyl algebra A1(ζ2)/k(m)=k(m){v,w}(vwζ2wv1),Z(A1(ζ2)/k(m))=k(m)[vl,wl],by the central maximal ideal (vlb,wlc), where l is the least integer such that 2l is a multiple of n. It is known (see e.g., [Citation11, Theorem 6.2]) that localizing A1(ζ2)/k(m) at a certain central element, ω, gives an Azumaya algebra. Therefore, the ramification locus in Z(A1(ζ2)/k(m)) is the zero set of this element. In fact, this ω gives the hyperbola vw=(1ζ2)l inside Z(A1(ζ2)/k(m)) and (b,c)ram(A1(ζ2)/k(m)) if and only if bc=(1ζ2)l.

Clearly, if ζ2=1, the algebra Jm/(e¯0) is a (commutative) zero-dimensional subscheme embedded into A2.

The cases b = 0 and c = 0, both of which are interchangeable, are more subtle (and to be honest, I don’t quite understand everything here myself). First recall that for any artinian algebra A, A and its semisimple quotient A/rad(A) have the same simple modules. Let m=(e¯0na,e¯1n,e¯2nc) and look at Jm=Jx/m. The radical of Jm is (e¯1) so Jm and Jm/(e¯1) have the same simple modules and we have a surjection JmJm/(e¯1)=k(m){e¯0,e¯2}(e¯0na,e¯2nc,e¯0e¯2ζe¯2e¯0,xe¯0+(1ζ2)x).

Assuming first that x0, this clearly means that Jm/(e¯1) only have the one-dimensional simple module defined by (5.5) M=k(m)·u,e¯0·u=(1ζ2)u,e¯2·u=0.(5.5)

Hence the only simple Jm-module is also M, with e¯1 acting as b = 0 on u. Observe that this is independent on a and c: regardless of the values of a and c, the action of Jm on M is given as above. This also implies that M is a simple one-dimensional Jx-module via the composition JxJmJm/(e¯1).

If x = 0, we get the symbol algebra (a,c)ζ, which is central simple.

The point of the above discussion is that even though there are several maximal ideals (simple modules) of Jx above mramR(Jx), the algebraic fibre Jm has only one simple module M, which is also then a simple module of Jx.

We summarize the discussion above in the following proposition.

Proposition 5.3.

Let abc = 0. Then the following can occur:

  1. Over the plane u0=0 the fibre Jm is a non-commutative deformation of a central quotient of the first quantum Weyl algebra A1(ζ2)/k(m),

    Jx(e¯0n,e¯1nb,e¯2nc).

    whose semisimplification is A1(ζp2)/(vpb,wpc) when x0 and (a,c)ζ when x = 0.

  2. Over the plane u1=0, Jm is a non-commutative deformation of

    1. the fat point k(m)[e¯0]/(e¯0n(1)n(1ζ2)n) if x0, ζ21;

    2. the cyclic algebra (a,c)ζ if x = 0, ζ21;

    3. the algebra

      k(m){e¯0,e¯2}(e¯02a,e¯22c,e¯0e¯2+e¯2e¯0)

      if x = 0, ζ2=1, and

    4. the algebra k(m)[e¯2]/(e¯2nc) if x0, ζ2=1.

The cases u1=0 and u2=0 are completely symmetric.

When x0, the fibre Jm has the simple module given by (5.5), which is also a simple module for Jx. In the case x = 0, the simple module is the simple module of (a,c)ζ.

We end this section with an example meant to inspire the reader to go where I dare not at this point.

Example 5.1.

The following situation should be worth pondering in some detail in view of the interest of Brauer groups and division algebras over surfaces (see for instance [Citation21]).

Let jZ:SZSpec(Z(Jx)) be a closed immersion, where SZ is a two-dimensional normal subscheme of Spec(Z(Jx)). The induced morphism over A(n)3=Spec(Λ) is denoted j:SΛA(n)3.

The sheaf Jx over Spec(Z(Jx)) and A(n)3 has pull-backs jZJx and jΛJx over SZ and SΛ, respectively. The algebras jZJx and jΛJx are maximal orders in jZJxSZK(SZ) and jΛJxSΛK(SΛ).

Now, Propositions 5.2 and 5.3 seem to give interesting information concerning the Brauer groups at the different points of the surfaces SZ and SΛ. This certainly includes the generic points. The reader is invited to examine this in more detail.

6 Rational points on XJx(r)

From now on, unless explicitly stated otherwise, we assume that B is a field K such that Kμn. We begin by looking at the “commutative points”.

6.1 The locus of one-dimensional points and its infinitesimal structure

Let KL be a field extension. A one-dimensional representation P=L·u must come from a maximal ideal M in Jx(r) and P=Jx(r)/M. Since L is a field, the ideal M must include the ideal generated by the commutators of e¯0, e¯1 and e¯2. Assume ζ2r1. Then the relations reduce to (1ζr)e¯0e¯1=0(1ζr)e¯0e¯2=0(1ζ2r)e¯1e¯2=xe¯0+x(1ζ2r).

This implies that x(e¯0+(1ζ2r))=0 so, if x0, we have that (ζ2r1,0,0) is a one-dimensional module. In addition, the conic (hyperbola) Cx given by e¯1e¯2=x in the plane e¯0=0 consists entirely of commutative points (i.e., one-dimensional modules). When x = 0, we find that the one-dimensional modules lie on the union of the coordinate axes.

Remark 6.1.

When ζ2r=1 we may assume that ζr=1. The relations e¯0e¯1+e¯1e¯0=0,e¯2e¯0+e¯0e¯2=0,e¯2e¯1e¯1e¯2=xe¯0

imply that 2e¯0e¯1=0,2e¯0e¯2=0,xe¯0=0.

Consider first when char(K)2. When x0 we get that the plane {e¯0=0} consists entirely of one-dimensional points. On the other hand, when x = 0, we get the union of the coordinate planes. Now, if char(K)=2, we get, when x0, the plane {e¯0=0} and, if x = 0, the whole A/K3.

Put u1:=e¯1n and u2:=e¯2n. The restriction, Z(Cx), of Cx to the center is then given by the equation u1u2=xn and a point M=(e¯0,e¯1b,e¯2c),b,cL,on Cx restricts to the point m=(u0,u1b,u2c),with b=bn,c=cnL,such that bc=xn.

Observe that Mij=(e¯0,e¯1ζib,e¯2ζjc),1i,jn1,also maps to m under ϕ, so Mijϕ1(m). This then means that Z(Cx) lies in the ramification locus.

Be sure to note that the divisors Ya, for all a with a0a1, intersect the {e¯0=0}-part of the ramification locus along Cx.

We will now compute the tangents between the one-dimensional modules. From now on we will for simplicity only consider the case r = 1 and leave it to the reader to insert r at the appropriate places. Therefore, put Jx:=Jx(1).

Remark 6.2.

The Ext-computations below are done in the Hochschild complex. Recall that for an K-algebra R we have the following isomorphisms ExtR1(M1,M2)HH1(R,HomK(M1,M2))DerK(R,HomK(M1,M2))/Ad,where HH1 is the Hochschild cohomology functor and Ad is the subgroup of Der(R,Hom(M1,M2)) of inner derivations.

Put P1:=Jx/M1 and P2:=Jx/M2, with M1,M2 lying over mMax(Z(Jx)), are both one-dimensional over L. It is easy to see that e¯0 must act as zero on any one-dimensional module. Therefore, we put P(b,c):=P1=L·u,withe¯1·u=bu,e¯2·u=cu,b,cL,

and similarly P(e,f):=P2. Furthermore, put δ(e¯0)u=Δ0v,δ(e¯1)u=Δ1v,δ(e¯2)u=Δ2v,Δ0,Δ1,Δ2L,where we chose v as basis for P(e,f).

Assume first that x0. The relations of Jx must map to zero in the group HomL(P(b,c),P(e,f)). This gives restrictions on the set {b,c,e,f,Δ0,Δ1,Δ2}. A small computation shows that these restrictions are:

  • Δ0=0.

  • b=ζe, and if fζc we get one free parameter; if not, we find that Δ1 and Δ2 are both free.

  • b=ζ2e: here we get that (bζe)(fζ2c)Δ1=0, so if f=ζ2c, then Δ1 and Δ2 are both free, otherwise only Δ2 is free.

  • By symmetry we can switch (b, e) and (c, f) and get the same result.

As for the inner derivations θHomL(P(b,c),P(e,f)) θ(u)=sv,sL,we get adθ(e¯1)(u)=θe¯1ue¯1θu=(be)svadθ(e¯2)(u)=θe¯2ue¯2θu=(cf)sv implying that dim(Ad)=1 unless b = e and c = f, in which case dim(Ad)=0.

The relevant computations yield the following propositions.

Proposition 6.1.

When x0 and ζ21, the Ext1-groups are ExtA1(P(b,c),P(e,f))={L,if a=ζe and f=ζcLif a=ζ2e and f=ζ2cLif a=e and c=f0,otherwise.

Similar reasoning yields the case x = 0:

Proposition 6.2.

When x = 0 and ζ21, the Ext1-groups are ExtA1(P(b,c),P(e,f))={Lif b=ζ2e,f=ζ2cL2if b=c=e=f=00otherwise.

The cases when ζ2=1 are more complicated since we then need to handle the case of characteristic two separately. Analogous computations as above gives the following propositions.

Proposition 6.3.

When x0 and ζ2=1, ExtA1(P(b,c),P(b,c))=L, and zero in all other cases, independent on characteristic.

Proposition 6.4.

When x = 0 and ζ2=1, the Ext1-groups are

  1. char(K)2: here ExtA1(P(b,c),P(±b,±c))=L and zero otherwise;

  2. char(K)=2: here ExtA1(P(b,c),P(b,c))=L3 and zero otherwise.

The above propositions determine the tangent space T{P(b,c),P(e,f)}.

6.2 Families of rational points

We will follow the construction given by D. Jordan in [Citation13] (with a few modifications so as to apply in our case) to identify some (but not necessarily all) finite-dimensional simple modules of Jx(r). Jordan’s construction gives a moduli for all simple modules over an algebraically closed field.

Put y¯:=x(1ζr)1e¯0+x and w¯:=e¯2e¯1y¯=e¯2e¯1x(1ζr)1e¯0x.

Recall from the proof of Proposition 4.10 that we had an automorphism τ on Jx, defined on e¯0 as τ(e¯0):=ζre¯0. The order of τ is lcm(r,n).

We note that e¯0w¯=w¯e¯0,e¯1w¯=ζ2rw¯e¯1,e¯2w¯=ζ2rw¯e¯2and w¯=ζ2r(e¯1e¯2τ(y¯)). We extend τ to the commutative K-algebra K[e¯0][w¯] by putting τ(w¯)=ζ2rw¯. Observe that this is not consistent with the definition of w¯ with respect to the τ as given in Proposition 4.10. For instance, τ(e¯2) is not defined. However, this is not a problem as we will see.

In addition, let m=(f)Max(K[e¯0]), for f irreducible. Recall that xK. If x = 0, clearly y¯=0.

The identity (6.1) e¯2e¯1i=(y¯ζ2riτi(y¯))e¯1i1+ζ2ire¯1ie¯2(6.1) or equivalently in our case, (6.2) e¯2e¯1i=x([i]ζre¯0+(1ζ2ri))e¯1i1+ζ2ire¯1ie¯2,(6.2) follows by a simple induction argument.

6.3 Torsion points

Suppose that dZ>0 is minimal with the property that (6.3) y¯ζ2rdτd(y¯)=x([d ]ζre¯0+(1ζ2rd))m.(6.3)

Then T(m):=Jx(r)Jx(r)(m,e¯2,e¯1d)is a simple e¯1- and e¯2-torsion module of dimension d.

Explicitly, let m=(f) be a maximal ideal in K[e¯0]. However, because of (6.3), we see that x([d ]ζre¯0+(1ζ2rd))=x[d ]ζr(e¯0+(1ζ)(1+ζrd))must be a factor in f and so m=(e¯0a), with a=(1ζ)(1+ζrd).

Put vi:=e¯1i+I, with I:=Jx(r)(m,e¯2). Then, (6.4) e¯0·vi=e¯0e¯1i+I=ζriae¯1i+I=ζriavi,ande¯1·vi=vi+1,(6.4) where the second equality follows since e¯0e¯1+I=ζre¯1e¯0+I=ζre¯1a+I. Also, (6.5) e¯2·vi=e¯2e¯1i+I=(6.2)x([i]ζre¯0+(1ζ2ri))e¯1i1+I=x(a[i]ζrζr(i1)+(1ζ2ri))vi1=ψit(a)vi1,(6.5) with ψit(a):=x(a[i]ζrζr(i1)+(1ζ2ri)).

Therefore, T(m)=i=0d1K·vi,with the actions (6.4) and (6.5). We see that when x = 0, we get e¯2·vi=0, for all i0. Observe that the torsion points are defined already over K.

Notice that ρ(e¯0)d=adI,and ρ(e¯1)d=ρ(e¯2)d=0.

When x = 0 we get ρ(e¯2)=0.

The above discussion proves the following proposition.

Proposition 6.5.

Given (d,r)Z>02, the element a is uniquely determined and so for each distinct choice (d, r) there is exactly one torsion module up to isomorphism. In other words, for each pair (d,r)Z>02 there is a unique e¯1e¯2-torsion point in @@@XJx(K) and all torsion points are K-rational.

Example 6.1.

When n=d=3, K=Q(ζ3) and x = 1, we get a family T(m)=T(a) of modules, parameterized by m=(e¯0a), aK. On matrix form T(a) is given as (with r = 1), ρa(e¯0)=(a000ζ3a000ζ32a),ρa(e¯1)=(000100010),ρa(e¯2)=(0a+ζ3+20001aζ3000)

Note that the module lies over the point (a3,0,0)Specm(Z(Jx))(K).

6.4 Torsion-free points

To determine the e¯1- and e¯2-torsion-free modules we proceed as follows.

Take m=(f)Max(K[e¯0]) such that τd(m)=m, with dZ>0 minimal with this property. Let L be the splitting field of f and aL such that f(a) = 0. Put Mb:=K[e¯0][w¯](m,w¯b)(K[e¯0]KL)[w¯]=L[e¯0][w¯],bL.

Observe that e¯0w¯=w¯e¯0 so K[e¯0][w¯] is a commutative subalgebra of Jx(r). Let f=i=0rfie¯0i be a generator for m. Then, recalling that τ(e¯0)=ζre¯0, we get τd(f)=i=0rfiτd(e¯0)i=i=0rfiζrdie¯0i,from which it follows that τd(Mb)=(i=0rfiζrdie¯0i,ζ2rdw¯b).

Now put J:=Jx(r)(Mb,e¯1sc),cL,and vi:=e¯1i+J. Since w¯=e¯2e¯1y¯, we find e¯2e¯1i=(e¯2e¯1)e¯1i1=(w¯+y¯)e¯1i1.

Furthermore, w¯e¯1+J=ζ2re¯1w¯+J=ζ2rbe¯1+J and y¯e¯1+J=(x(1ζr)1e¯0+x)e¯1+J=e¯1(x(1ζr)1ζre¯0+x)+J=(x(1ζr)1ζra+x)e¯1+J.

This implies that e¯2e¯1i+J=(ζ2r(i1)b+x(aζr(i1)(1ζr)1+1))e¯1i1+J,hence (6.6) e¯2·vi=ψitf(a,b)vi1,(6.6) where ψitf(a,b):=(ζ2r(i1)b+x(aζr(i1)(1ζr)1+1)).

Clearly, ψitf is a function ψitf:L×LL.

Modulo J we have that e¯1s=c, so e¯1sc1=1=e¯10. From this observation follows e¯2e¯10+J=e¯2c1e¯1s=c1(w¯+y¯)e¯1s1+J,so (6.7) e¯2·v0=c1ψstf(a,b)vs1.(6.7)

Clearly (6.8) e¯1·vi={vi+1,if 0i<s1cv0,if i=s1(6.8) and (6.9) e¯0·vi=ζriavi,0is.(6.9)

The case x = 0 gives (6.10) e¯2·vi={ζ2r(i1)bvi1,i0c1ζ2r(s1)bvs1,i=0.(6.10)

We have that ρ(e¯0)s=asI,ρ(e¯1)s=c,and ρ(e¯2)s=i=1sψitf(a,b).

When x = 0 we get ρ(e¯2)=(0b00000ζn2rb000000ζn2r(s2)bc1ζn2r(s1)b000)and the other two operators are the same. Observe that this implies that, when x0, the corresponding maximal ideal lies over the point (as,c,i=1sψitf(a,b))Specm(Z(Jx(r))).

There are two distinct cases to consider, giving simple e¯1-torsion-free Jx(r)-modules:

  1. b = 0. Here τd(M0)=M0 and we define

    M(m,c):=Jx(r)Jx(r)(M0,e¯1dc)=i=0d1L·vi.

  2. b0. Here Mb is periodic since ζ2r is a root of unity. The order of Mb is s=lcm(d,l), with l the least integer such that (ζ2r)l=1. Define

    M(m,b,c):=Jx(r)Jx(r)(Mb,e¯1sc)=i=0s1L·vi.

In both cases the actions are given by formulas (6.6)–(6.9).

There is an obvious e¯1-e¯2-symmetry which gives that the same construction yields the e¯2-torsion-free modules M(m,c) and M(m,b,c), as well.

When K=Kal, the above construction is a complete classification of simple finite-dimensional modules of Jx(r) up to isomorphism. When K is not algebraically closed there are finite-dimensional simple modules not isomorphic to one in the above families (in other words, a module, simple over K, might reduce over Kalg; there are “more” simple modules over Kal than over K). Studying how modules degenerate into families is actually quite subtle and should definitely be studied further.

We have already discussed the torsion-case above, so we now focus on the torsion-free situation. First,

Example 6.2.

The commutative points in Section 6.1 is gotten with a = 0 and d = 1. In other words, e¯0 is a factor in f.

Example 6.3.

We still look at n=d=3, r = 1, x = 1, m=(e¯0a), aL and K=Q(ζ3). We find ψ1tf(a,b)=b+a(1ζ3)1+1=(13ζ3+23)a+b+1,ψ2tf(a,b)=ζ32b+aζ3(1ζ3)1+1=(13ζ313)a+(ζ31)b+1,ψ3tf(a,b)=c1(ζ3b+aζ32(1ζ3)1+1)=(23ζ313)c1a+ζ3c1b+c1and matrices ρ(e¯0)=(a000ζ3a000ζ32a),ρ(e¯1)=(00c100010),ρ(e¯2)=(0ψ1tf(a,b)000ψ2tf(a,b)c1ψ3tf(a,b)00)

The module ρ lies over the point (a3,c,ψ1tf(a,b)ψ2tf(a,b)ψ3tf(a,b))Specm(Z(Jx))(L).

Explicitly, the z-coordinate is (19ζ3+19)a3c1+(ζ31)b3c1+(2ζ3+1)abc1(ζ3+1)c1.

Remark 6.3.

We make two remarks here.

  1. Fix t0,t1,t2L and mt=(u0t0,u1t1,u2t2)Max(Z(Jx)). Then the extension of mt to Jx is

    Mt=(e¯03t0,e¯13t1,e¯23t2).

    A necessary condition for the associated module to be isomorphic to one in the above family is that t03L. The same remark can be given for e¯2. But notice that e¯1 is already given on the correct form. Switching to the e¯2-torsion-free versions, we get that e¯2 is on the correct form.

    However, if t03L then ρt is not isomorphic over L to a module in the family. To get an isomorphism we need to extend to L(t03). The same applies to e¯2 and t2. Therefore, over the field K(2):=L[L1/3], every module is isomorphic to a (unique) module in the family. The same remarks apply to the general case of arbitrary n.

  2. Another thing worth remarking upon, is that in the construction above, we started from a commutative algebra and used induction to bigger rings and constructing modules along these inductions. In this process we have not used the possibility that some power of e¯0 is torsion, without e¯0 being torsion. The arguments in [Citation13] show that any module, e¯0-torsion or not, is isomorphic to one in the above families over Kal (clearly, it must be torsion with respect to some element, though). However, this might not be true over fields which are not algebraically closed.

Example 6.4.

We continue the above example and assume furthermore that x0. Let us look at the line l in Specm(Z(Jx)) parameterized by c(t):=(t,t+ζ3,t+9),tL,

and the family Mt:=JxJxmt,mt:=(e¯03t,e¯13tζ3,e¯23t9)of (left) Jx-modules.

Clearly, l intersects the ramification locus (not necessarily uniquely) in the point P:=(0,ζ3,9). Let M0 be a Jx-module lying over P. For instance, M0=M(m0,ζ3,2), with m0=(u0), is such a module since the contraction to Z(Jx) is (0,c,b3+1)=(0,ζ3,9).

However, Mi:=M(m0,ζ3,ζ3i·2), i = 1, 2, are also two modules contracting to P. Hence the arithmetic fibre over P is Φ(P)={M0,M1,M2}.

Notice that these modules are non-isomorphic over Kal since they are distinct points in the family. Therefore they are non-isomorphic also over K.

The geometric fibre is XJx/m0=Mod(Jx/m0).

(By m0 we mean the 2-sided ideal generated by m0.) The support of XJx/m0 is the point corresponding to the symbol algebra (ζ3,9)ζ3 (which is simple and so has only one simple module).

After a, not quite trivial, computation one finds ExtJx1(Mi,Mj)={L2,if i=jL,if ij,giving us the tangent structure of XJx over PSpecm(Z(Jx)).

Acknowledgments

As already indicated, this paper has traveled a long and winding road. Its primary ancestor was an attempt to answer some ring-theoretic questions raised in [Citation16] back in 2006. However, as the years went by, it evolved into something completely different. Along the way, I have benefited a lot from discussions and help from many people, but let me single out the help from Arvid Siqveland and Arnfinn Laudal as particularly important.

References

  • Artin, M., Small, L. W., Zhang, J. J. (1999). Generic flatness for strongly Noetherian algebras. J. Algebra 221(2):579–610. DOI: 10.1006/jabr.1999.7997.
  • Brown, K. A., Goodearl, K. R. (2002). Lectures on Algebraic Quantum Groups. Advanced Courses in Mathematics. CRM Barcelona. Basel: Birkhäuser Verlag.
  • Benkart, G., Roby, T. (1998). Down-up algebras. J. Algebra 209(1):305–344. DOI: 10.1006/jabr.1998.7511.
  • Chan, D., Nyman, A. (2013). Non-commutative Mori contractions and P1-bundles. Adv. Math. 245:327–381.
  • Ceken, S., Palmieri, J. H., Wang, Y.-H., Zhang, J. J. (2016). The discriminant criterion and automorphism groups of quantized algebras. Adv. Math. 286:754–801. DOI: 10.1016/j.aim.2015.09.024.
  • Ekström, E. K. (1989). The Auslander condition on graded and filtered Noetherian rings. In: Séminaire d’Algèbre Paul Dubreil et Marie-Paul Malliavin, 39ème Année (Paris, 1987/1988), volume 1404 of Lecture Notes in Mathematics. Berlin: Springer, pp. 220–245.
  • Eriksen, E., Laudal, O. A., Siqveland, A. (2017). Noncommutative Deformation Theory. Monographs and Research Notes in Mathematics. Boca Raton, FL: CRC Press.
  • Gille, P., Szamuely, T. (2006). Central Simple Algebras and Galois Cohomology, volume 101 of Cambridge Studies in Advanced Mathematics. Cambridge: Cambridge University Press.
  • Gómez-Torrecillas, J., Lobillo, F. J. (2004). Auslander-regular and Cohen-Macaulay quantum groups. Algebras Represent. Theory 7(1):35–42.
  • Hartwig, J. T., Larsson, D., Silvestrov, S. D. (2006). Deformations of Lie algebras using σ-derivations. J. Algebra 295(2):314–361. DOI: 10.1016/j.jalgebra.2005.07.036.
  • Irving, R. S. (1979). Prime ideals of Ore extensions over commutative rings. II. J. Algebra 58(2):399–423.
  • Jahnel, J. (2014). Brauer Groups, Tamagawa Measures, and Rational Points on Algebraic Varieties, volume 198 of Mathematical Surveys and Monographs. Providence, RI: American Mathematical Society.
  • Jordan, D. A. (1995). Finite-dimensional simple modules over certain iterated skew polynomial rings. J. Pure Appl. Algebra 98(1):45–55. DOI: 10.1016/0022-4049(94)00026-F.
  • Larsson, D. (2017). Equivariant hom-Lie algebras and twisted derivations on (arithmetic) schemes. J. Number Theory 176:249–278.
  • Larsson, D. (2023). Arithmetic geometry of algebras with large centres. Preprint. Available at http://www.numberlab.se.
  • Larsson, D., Silvestrov, S. D. (2007). Quasi-deformations of sl2(F) using twisted derivations. Commun. Algebra 35(12):4303–4318.
  • Maury, G., Raynaud, J. (1980). Ordres maximaux au sens de K. Asano, volume 808 of Lecture Notes in Mathematics. Berlin: Springer.
  • McConnell, J. C., Robson, J. C. (1987). Noncommutative Noetherian Rings. Pure and Applied Mathematics (New York). Chichester: Wiley. With the cooperation of L. W. Small, A Wiley-Interscience Publication.
  • Neukirch, J. (1999). Algebraic Number Theory, volume 322 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Berlin: Springer-Verlag. Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder.
  • Resco, R., Small, L. W., Stafford, J. T. (1982). Krull and global dimensions of semiprime Noetherian PI-rings. Trans. Amer. Math. Soc. 274(1):285–295. DOI: 10.1090/S0002-9947-1982-0670932-1.
  • Saltman, D. J. (2008). Division algebras over surfaces. J. Algebra 320(4):1543–1585. DOI: 10.1016/j.jalgebra.2008.02.028.
  • Smith, A. D., Paul Smith, S. (1997). Some 3-dimensional skew polynomial rings. Draft of April 1, 1997.
  • Stafford, J. T., Van den Bergh, M. (2008). Noncommutative resolutions and rational singularities. Michigan Math. J. 57:659–674. Special volume in honor of Melvin Hochster. DOI: 10.1307/mmj/1220879430.
  • Yi, Z. (1995). Injective homogeneity and the Auslander-Gorenstein property. Glasgow Math. J. 37(2):191–204. DOI: 10.1017/S0017089500031098.