Euler factors of equivariant –functions
of Drinfeld modules and beyond
Abstract.
In [2], the first author and his collaborators proved an equivariant Tamagawa number formula for the special value at of a Goss–type –function, equivariant with respect to a Galois group , and associated to a Drinfeld module defined on and over a finite, integral extension of . The formula in question was proved provided that the values at of the Euler factors of the equivariant –function in question satisfy certain identities involving Fitting ideals of certain –cohomologically trivial, finite –modules associated to the Drinfeld module. In [2], we prove these identities in the particular case of the Carlitz module. In this paper, we develop general techniques and prove the identities in question for arbitrary Drinfeld modules. Further, we indicate how these techniques can be extended to the more general case of higher dimensional abelian –modules, which is relevant in the context of the proof of the equivariant Tamagawa number formula for abelian –modules given by N. Green and the first author in [4]. This paper is based on a lecture given by the first author at ICMAT Madrid in May 2023 and builds upon results obtained by the second author in his PhD thesis [12].
Key words and phrases:
Drinfeld modules, –motives, local shtukas, étale cohomology, crystalline cohomology, equivariant motivic –functions, equivariant Tamagawa number formula, Euler factors2010 Mathematics Subject Classification:
11G09, 11M38, 11F801. Introduction
Let be a tower of finite field extensions, where , is a power of a prime , and is Galois, of abelian Galois group . Further, let and denote the integral closure of in and , respectively, and consider a Drinfeld module defined on and with coefficients in .
In [2], the authors associated to the set of data a –equivariant Goss–type –function
where is a certain Goss space of mixed characteristic containing naturally a copy of and is the completion of the algebraic closure of with respect to the unique extension of the valuation of of uniformizer (See the Introduction in [2] for the precise definitions.)
As a natural –equivariant generalization of the Goss –function associated to and studied by Taelman in [13], the –function is defined as an infinite product of Euler factors
where runs over all the primes in the maximal spectrum of which are tamely ramified in and of good reduction for . For each such prime , is a polynomial with coefficients in , very closely related to the well understood characteristic polynomial
of the action of a Frobenius morphism associated to on the –adic Tate module associated to the rank Drinfeld module at any prime , viewed as a free, rank module over the completion of at . In fact, one has equalities (see [2], Introduction):
where the infinite product converges in , is a choice of Frobenius for in , and is the idempotent in associated to the trivial character of the inertia group of in .
The main result of [2] is the proof of a Tamagawa number formula of the type
where is a certain Euler product completion of at primes which are either wildly ramified in or of bad reduction for , constructed out of some additional arithmetic data (called a taming module) for . The numerator and denominator of the right side in the above formula are both volumes (with values in and defined precisely in [2]) of certain compact –modules of Arakelov type and .
The above formula, proved in [2] for Drinfeld modules and extended in [4] for higher dimensional abelian –modules, generalizes to the –equivariant setting Taelman’s celebrated class–number formula [13] for the value of the Goss zeta–function associated to . The interested reader should also consult [1] for a slightly different proof in the abelian –module case, using deformation theory, which only gives the desired result under certain restrictive conditions.
However, the proof of the above formula in [2] and [4] (and the same applies to the proofs given in [13] and [1]) hinges upon an essential equality at the level of Euler factors, namely
for all good and tame primes as above, where the numerator and denominator of the right side are certain special generators of the Fitting ideals of the finite –modules of finite projective dimension and , respectively. In the Appendix of [2], we prove the above equality only in the simplest case where is the Carlitz module, and Taelman does the same in [13], in the case where is trivial.
In this paper, we develop techniques which allow us to prove the above equality for arbitrary Drinfeld modules (see Theorem 2.3 and its proofs in the unramified case, given in §5 and tamely ramified case, given in §6.) In §8, we also indicate how our techniques can be easily extended to prove the above formula for higher dimensional abelian –modules, satisfying a certain purity condition. The case of arbitrary abelian –modules will be treated in a separate, upcoming paper.
2. The statement of the problem
Let be a prime number and let be a power of . In what follows, denotes the rational function field in one variable over For any commutative -algebra we denote by the -power Frobenius endomorphism of . We denote by the twisted polynomial ring in , with the property that
Let be a finite, separable extension of and let be a finite abelian extension of with Galois group We also assume that the field of constants in is i.e.
Let us denote by Note that if denotes an arbitrary normalized valuation on and denotes the normalized valuation of uniformizer , then
Let and denote the integral closures of in and , respectively. In what follows, we abuse notation and use the same letter for normalized valuations and the associated maximal ideals of elements of strictly positive valuation.
Next, we consider a Drinfeld module of rank defined on with values in More precisely, is given by an -algebra morphism
where , for all and This gives rise to a functor
In other words, for any -module we denote by the module whose underlying –module is and the -action is given by
Let and let and denote the completions of and with respect to the valuation . For all , we denote by the -module of -torsion points of i.e.
The -adic Tate module of is defined as
Since is a PID, we also have
where the transition maps in the projective limit are given by multiplication with a generator of , while
Recall that and are free modules of rank over and , respectively, and are endowed with obvious -linear, continuous -actions, where
Let MSpec(), such that Fix a choice of decomposition group and a Frobenius morphism Then, it is known (see [2] and the references therein) that if has good reduction at (i.e. ), the -representation is unramified at and the polynomial
is independent of and actually lies in Above, denotes the identity matrix.
Definition 2.1.
Let be an -module which is free of rank as an -module. Then it is known (see [2] Proposition A.4.1) that the Fitting ideal is principal and has a unique –monic generator of degree We denote this generator by i.e.
The following is Proposition A5.1. from the Appendix in [2]:
Proposition 2.2.
Assume that is tamely ramified in and let be any Drinfeld module as above. Let denote the prime in sitting below and let Then the following hold:
-
(1)
The -modules and are free of rank and therefore and are monic polynomials of –degree .
-
(2)
We have an equality
where denotes the unique monic generator of and .
Let denote the inertia and decomposition groups of in , respectively. Let be the image of via the Galois restriction map . Our main goal in this paper is the proof of the following.
Theorem 2.3.
Assume that is tamely ramified in and that has good reduction at . Then, we have an equality in
where is the idempotent of the trivial character of in
A proof of the above statement in the case where is the Carlitz module , defined by , was given in the Appendix of [2]. Below, we develop techniques which settle the above theorem for a general Drinfeld module .
Proposition 1.2(2) gives us a good understanding of . Therefore a major portion of our work is directed towards understanding the relation between and
3. The reduction of modulo
In this section, we fix a prime such that has good reduction at We are not assuming that is necessarily tamely ramified in . Let us denote by the prime in that lies below After reduction of (i.e. reduction of the coefficients of modulo ), we obtain the rank Drinfeld module defined over , given by the -algebra morphism
where with being the obvious map. The Drinfeld module has rank , characteristic , and height . By definition, is the unique integer which, for any generator of , satisfies the equality
where , , and . (See [3], Section 4.5 for the existence of .)
Recall that, by the notation introduced above, we have a field isomorphism Consequently, the Frobenius morphism is an endomorphism of and it acts naturally and linearly on all the Tate modules associated to .
Next, we fix , and consider the characteristic polynomial of the action of the -power Frobenius morphism, viewed as an endomorphism of , on the free –module of rank :
Then, is independent of and lies in . (See §4.12 in [3].) By Theorem 4.12.15 in [3] and the discussion preceding that, we have the following.
Proposition 3.1.
Any root of satisfies the following properties:
-
(1)
for all finite places of except for exactly one place above
-
(2)
There is only one place of lying above
-
(3)
where denotes the unique extension to of the normalized absolute value of corresponding to
-
(4)
divides
Further, one can consider the characteristic polynomial of (viewed as endomorphism of ) acting on the free –module of rank (see [3], Section 4.5 for the calculation of the rank)
This is a monic polynomial in of degree In §? below we will prove the following.
Proposition 3.2.
The polynomial divides the polynomial in .
Proof.
See §7 below for the proof of a stronger statement. ∎
Let and be the completions at of and , respectively. Our choice of decomposition group corresponds to choosing an embedding at the level of separable closures of and , such that Galois restriction induces a group isomorphism . Since has good reduction at and the Galois representations are unramified at , it is not difficult to see that we have
where is the integral closure of in the maximal unramified extension of in . Moreover, the reduction map induces isomorphisms of –modules
(1) |
where
The group isomorphism above sends (the image of our choice of Frobenius in ) to . Consequently, we have an equality of characteristic polynomials in :
(2) |
Consequently, Proposition 3.1 gives us information on the roots of the characteristic polynomial . The following corollary regarding the coefficients of will be particularly useful in what follows.
Corollary 3.3.
Let , with . Then, we have
-
(1)
and , for all .
-
(2)
is a polynomial of degree in with the same leading coefficient as .
-
(3)
, for some , where is the unique monic generator of
Above, denotes the degree in of a polynomial in
Proof.
Let denote the roots of in the integral closure of . Then
Let denote an extension to of the normalized absolute value of corresponding to (also denoted by below.) By Proposition 1.4, we have , for all . Therefore, we have
Furthermore, since is non-archimedean, we have
This concludes the proof of part (1).
Part (2) is a direct consequence of part (1). Further, since , part (3) is a direct consequence of Proposition 3.1(1)–(3). ∎
4. Fitting ideals of Tate modules and consequences
We begin by stating a general commutative algebra result regarding Fitting ideals of modules over certain rings of equivariant Iwasawa algebra type. For a proof of this result, see Proposition 4.1 in [5].
Proposition 4.1 (Greither–Popescu).
Let be a semi-local, compact topological ring, and let be a pro-cyclic group, topologically generated by Suppose that is an –module which is free of rank as an –module. Let denote the matrix of the action of on some -basis of Then, we have an equality of –ideals
Proof.
See the proof of Proposition 4.1 in [5]. ∎
An immediate consequence of the above proposition is the following.
Corollary 4.2.
For all , we have the following equalities of –ideals.
-
(1)
If , then
-
(2)
If , then
Proof.
Apply the proposition above to , , and the module
which is –free of rank , if and, respectively, the module
which is –free of rank , if ∎
Next, we fix a prime lying above . We let , and , where . Then, is a choice of Frobenius for and its image corresponds via the group isomorphism to , viewed as an endomorphism of .
Lemma 4.3.
Let . Then, we have the following canonical isomorphisms of –modules:
-
(1)
-
(2)
, assuming that .
Above, we let for any –module .
Proof.
By the definition of , we have
Consequently, part (2) follows from part (1) via the second isomorphism in (1) above.
In order to prove part (1), apply the functor to the exact sequence of –modules
Since the –module is divisible and therefore injective (as is a PID), the above functor is exact. Therefore, we obtain the following exact sequence of –modules:
(3) |
Now, it is easy to see that one has an isomorphism of –modules
for all and all , where is the class of in .
Now, Proposition 3.1(3) and Proposition 3.2 show that the eigenvalues of acting on the –vector space are all different from . Consequently, is an automorphism of this –vector space. Consequently, when one takes the –invariants and coinvariants in the exact sequence (3) above, one obtains an isomorphism of –modules
which concludes the proof of the Lemma.
∎
Corollary 4.4.
For all , the following equalities of –ideals hold:
Further, if , then we have an equality of –ideals
Here, , is the image of in , and is the localization of at .
Proof.
First, note that the isomorphism of –modules in Lemma 4.3(1) can be rewritten as an isomorphism of –modules
where the ring morphism is the –linear map given by Galois restriction, which maps The isomorphism above permits us to apply the well known base–change property of Fitting ideals which, combined with Corollary 4.2 above, leads to the equalities of –ideals
Next, assume that and observe that since is finite (and therefore –torsion), we have isomorphisms
Consequently, base–change for Fitting ideals applied to the ring extension and the last equality of ideals displayed above gives
However, since the ring extension is faithfully flat (because is), we have
Above, we used the fact that if is a faithfully flat extension of commutative rings and is an ideal in , then . (See [11], Chapter 2, Section 4, 4.C(ii).) ∎
In the next two sections, we provide a proof of Theorem 2.3. For technical reasons which will become apparent shortly, we treat first the unramified case.
5. The unramified case
We keep the notations and assumptions of the previous section. In addition, we assume that the prime is unramified in Consequently, we have and throughout.
Lemma 5.1.
Under the current assumptions, we have equalities of ideals.
-
(1)
for all , with .
-
(2)
Proof.
We are ready to prove the following refinement of Theorem 2.3 in the unramified case.
Proposition 5.2.
Proof.
According to Corollary 3.3 (see (1) and (3) in loc.cit.), and , viewed as polynomials in , have degrees equal to and leading coefficients equal to . Therefore, and are indeed monic polynomials of common –degree . Further, Corollary 3.3(3) shows that which, if combined to Proposition 2.2(2), proves part (1) of the statement above.
Next, we focus on the proof of equality (2) in the Proposition above. In order to simplify the notation, let
Then, and are both monic polynomials in , of degrees equal to . (See Proposition 2.2(2) and Corollary 3.3.) Further, Lemma 5.1(1) and the definition of imply that they satisfy the following equalities
Now, it is easy to check that the total ring of fractions of is . Since and are monic, they are not zero–divisors in and . Therefore, the equalities above imply that
This implies that there exists and , such that
where is the unique monic generator of the maximal ideal .
We claim that . In order to prove that, let us note that Lemma 5.1 implies that
On the other hand, Proposition 3.2 implies that
Consequently, since , we have
Therefore, , which implies that
However, since , this shows that the divisibility above happens in . Therefore, we have , as claimed and, consequently
However, since and are monic polynomials in of common degree , we must have
which concludes the proof of part (2) of the Proposition. Part (3) is a consequence of parts (1) and (2).
∎
Remark 5.3.
Note that since Corollary 3.3 is valid in general, regardless of the ramification status of in , the polynomials , , and are monic of degree in in general, even in the tamely ramified case. Also, the proof given to part (1) of the Proposition above is valid in general, so even in the tamely ramified case.
6. The tamely ramified case
Now, suppose that is tamely ramified in , i.e. . As above, we denote by
the idempotent in associated to the trivial character of the inertia group of in . The goal of this section is to prove the following analogue of Proposition 5.2 in this case.
Proposition 6.1.
Proof.
Observe that part (3) is a consequence of parts (1) and (2). Since Remark 5.3 settles everything but part (2) in the more general, tamely ramified case, we will focus below on proving part (2).
Note that, if and is a fixed prime in above , as before we have isomorphisms of –modules
Consequently, base–change for Fitting ideals and the isomorphism show that it suffices for us to work locally and prove that we have an equality of –ideals
(4) |
where is the ring of integers in the –adic completion of and is now viewed as , where , etc. have the obvious meanings. Also, let us note that the faithful flatness of the ring extension combined with part (1) of the Proposition give us
Let denote the maximal sub-extension of , which is unramified at Let denote the prime in lying below . Since is unramified at , Proposition 5.2 leads to an equality of ideals
where denotes the Frobenius element of in . Since (as is totally ramified at ), we also have an equality of ideals
(5) |
For simplicity, let Then, Proposition A.5.1. in [2] gives isomorphisms of –modules
Since , the isomorphisms above lead to an equality of –ideals
(6) |
Remark 6.2.
The definition of Fitting ideals implies that if is a finitely generated –module, then
We will use equality (6) to prove (4). The obvious ring isomorphism which sends and equality (5) give
(7) |
Next, we will calculate For that, we need a definition and a couple of lemmas.
Definition 6.3.
Let be a commutative ring and a finitely presented –module. is called quadratically presented if there exists an and an exact sequence of –modules
is called locally quadratically presented if is a quadratically presented –module for all .
The following computationally useful result is due to Johnston and Nickel. (See [8]).
Lemma 6.4 (Johnston–Nickel).
If is a commutative ring, is a locally quadratically presented and
is an exact sequence of finitely presented –modules, then
A supply of locally quadratically presented –modules in the context at hand is given by the following Lemma, whose proof uses concepts of group cohomology discussed in detail in the Appendix of [2].
Lemma 6.5.
Under the above assumptions, the following hold.
-
(1)
The –submodule of is –cohomologically trivial (–c.t.), for all .
-
(2)
For all , the –modules and are locally quadratically presented.
Proof.
(1) Fix . Since the –module is annihilated by and , we have
Consequently, the inflation–restriction sequences in group–cohomology give us exact sequences of –modules
for all . However, , for some . Since is unramified, we have isomorphisms of –modules
where the second isomorphism is multiplication with , where is a generator of (as an ideal of ). This shows that the –module is –induced and therefore –c.t. Now, the exact sequence above implies that
which shows that, indeed, is –c.t., for all .
(2) Part (1) implies that the modules are –c.t. (since and are. Therefore, since is a PID, they have projective dimension at most over . However, is finite and since is infinite, must have projective dimension exactly over . Therefore, there are , with and exact sequences of –modules
for all . Again, by the finiteness of , one concludes that , which gives the desired local quadratic presentations, concluding the proof. ∎
Corollary 6.6.
Under the above assumptions, the following hold.
-
(1)
For all , the identity map induces an isomorphism of –modules
-
(2)
We have equalities of –ideals
Proof.
(1) The identity map induces an isomorphism of –modules between the two modules in question. However, since , acts trivially on . Therefore, we have
which shows that the identity is also an isomorphism of –modules.
(2) Consider the short exact sequences of –modules
for all Apply Lemma 6.5(2) and Lemma 6.4 to conclude that
for all . Consequently, we obtain the following equalities of –ideals.
Now, we combine the above equalities with the –module isomorphisms in part (1) to obtain
Now, since , we have
which concludes the proof of the Corollary. ∎
7. The –motive and local –shtukas associated to
This section is dedicated to the proof of a more precise version of Proposition 3.2, which will be stated below. The arguments below ensued from a lengthy and enlightening e-mail exchange [6] between the first author and Urs Hartl, to whom we would like to express our gratitude.
As before, is the reduction modulo of the original Drinfeld module defined on and over . So, is a Drinfeld module of rank , defined on , over the finite field , and of characteristic . In §3, we denoted by the height of , picked a prime , and considered
where is the –power Frobenius endomorphism of , acting naturally and linearly on its various Tate modules. In the same section, we stated the well known result that is independent on and in fact (which will also be proved very briefly below) and claimed without a proof that in . The precise goal of this section is to prove the following.
Proposition 7.1.
Let be the absolute integral closure of . Then the following hold.
-
(1)
The roots of in are exactly those roots of which satisfy
where is the unique extension to of the normalized absolute value on .
-
(2)
If is such a root, then its multiplicities in and are the same.
-
(3)
In particular, in
Obviously, part (3) of the above Proposition is a direct consequence of parts (1) and (2). In order to prove parts (1) and (2), we need to introduce first the –motive associated to . We follow most of the notations and conventions in [3], Chapter 5. Let and let be a fixed algebraic closure. In what follows, denotes the additive affine line, viewed as a scheme over . We think of as a functor from the category of –algebras to the category of –modules
where is endowed with a natural –module structure via the –algebra (injective) morphism
Definition 7.2.
As is a perfect field containing (the field of definition of ), we follow loc.cit. and define the –motive over associated to as the left –module
endowed with the left –module structure given by
Remark 7.3.
It is important to note that the –module has some distinctive properties (see loc. cit. for proofs): First, it is obvious that is a free –module of rank (which is the dimension of the –motive ) and (less obvious) that it is a free –module of rank (which is the rank of the –motive .) Second, it is important to note that since is perfect, is an –submodule of and, as a consequence of the definition of , we have
where is the obvious –algebra map of kernel .
It is not difficult to check that the evaluation (perfect) pairing
gives rise to an isomorphism of –modules
where acts on by raising the coefficients of the Laurent series in question to the –th power and acts via multiplication. For every , this leads to a natural isomorphism of –modules
after identifying via the isomorphism .
Now, we fix an arbitrary and let denote the monic generator of . We let
Note that if , then we have natural isomorphisms of topological rings
Further, note that since is a free –module of rank (see the Remark above), then is a free –module of rank and, consequently, a free –module of rank . In addition, if we view as the canonical topological generator of , then the free –module is endowed with a –semilinear endomorphism, abusively denoted , and given by
Definition 7.4.
The data consisting of the free –module of rank together with its –semilinear endomorphism defined above is called the local –shtuka over associated to at .
The link between the local shtuka and the Tate module is obtained by taking the projective limit as of the isomorphisms defined above, to get an isomorphism of –modules
The following useful result is due to Hartl–Singh [7], building upon earlier work of Laumon [10].
Proposition 7.5.
For all , the local –shtuka over splits canonically as a direct sum of local –shtukas over
where is the maximal –submodule of on which the restriction of is bijective and is the maximal –submodule of on which the restriction of is topologically nilpotent (i.e. there exists an such that .)
Proof.
See the proofs of Propositions 2.7 and 2.9 in [7] and keep in mind that the field is perfect in our context. Also, note that in loc.cit. the authors work with the –module structures (local) rather than the —module structures (semi-local), but the transition local/semi–local is seamless. ∎
In light of the above result, since obviously acts as an isomorphism of (i.e. is an étale local –shtuka over ), the isomorphism of –modules defined above can be rewritten as
The next crucial step is provided by the following extension to the case of of Lang’s well known theorem on (see [9]).
Lemma 7.6 (Hartl, [6]).
Under the above assumptions, the following hold.
-
(1)
The map taking is surjective.
-
(2)
Any free –module of finite rank , endowed with a bijective, –semilinear endomorphism satisfies the property that the standard map
is an isomorphism of –modules.
Proof.
Since we have a ring isomorphism , it suffices to prove part (1) for . So, given a matrix , i.e.
we need to find a matrix given by
such that the matrices satisfy the relations
Lang’s theorem (see loc.cit.) implies that part (1) is true for , so we can find a matrix satisfying the –th relation above. After multiplying the –th relation above to the right by we obtain the equivalent relation
which consists of one Artin–Schreier equation for each entry of . Since is algebraically closed, these equations have solutions. Therefore, inductively, one can find matrices , for all , as desired.
Part (2) follows immediately from part (1) in a standard way: take a basis of over and let be the matrix of in that basis. Let such that . Then is an -basis of which is contained in . This concludes the proof. ∎
By applying the Lemma above to and , we conclude that we have the following natural isomorphisms of –modules
The above isomorphism leads to a further isomorphism of –modules
which, if composed with the map gives an isomorphism of –modules
This prompts the following.
Definition 7.7.
The first –adic étale cohomolgy group of is defined by
Note that the maps lead to the following –module isomorphisms.
Further, the first –adic crystalline cohomology group of is defined by
Note that for all we have isomorphisms and inclusions of –modules
The following holds at primes different from the characteristic of . (See [3], Chapter 5 as well.)
Lemma 7.8.
If is an element in different from the characteristic of , then
In other words, is bijective on and we have canonical isomorphisms of –modules
Proof.
(sketch) It is easy to show that since in , is injective and therefore bijective on the finite dimensional –vector spaces , for all . The bijection of on is obtained now by taking the projective limit as . ∎
Now, we have all the necessary ingredients to prove parts (1) and (2) of Proposition 7.1.
Proof.
First, take and observe that, based on the previous Lemma and Definition, we have canonical isomorphisms of –modules
As a consequence, from the definition of , we have
The last equality proves that is independent of and that it has coefficients in . Further, if one applies the analogues of Lemmas 7.6 and 7.8 to the finite, étale –shtukas over (see [7]), one concludes that has coefficients in . Since (intersection viewed inside ), has coefficients in , as stated before.
Now, from the definitions, we also have similar natural isomorphisms of –modules
Therefore, when combining these with the second note in Definition 7.7, we obtain equalities
Firstly, this shows that divides in . However, since are both in the polynomial ring , this divisibility holds in . Secondly, note that
Since (which is a power of ) acts bijectively on the étale piece and topologically nilpotently on the nilpotent piece of , the roots of are all –adic units and the roots of are in the maximal ideal of an absolute integral closure of . This concludes the proof. ∎
8. Final thoughts: the more general case of pure, abelian -modules
Most of the techniques developed in this paper can be extended to the more general case of pure, abelian –modules, as indicated briefly below. The very general case of abelian –modules, whose equivariant –functions are the main object of study in [4], will be treated in upcoming work.
With notations as in §2, let be the ring of matrices with entries in , for a fixed .
Definition 8.1.
A –module of dimension , defined over is an -algebra morphism
where , , and .
For any –module , one denotes by
the additive group , endowed with the –module structure given by and (evaluation of at , for all ), respectively.
Note that Drinfeld modules defined on and over are the –dimensional –modules defined in §2 above. However, unlike the –dimensional case, where the rank of the Drinfeld module is simply equal to , in the case of higher dimensional –modules, introducing the notion of rank is much more subtle. We indicate briefly how this is done below. This will lead to the more restrictive notion of an abelian –module.
Let be the separable closure of . As in §7, we let (where is viewed as a group scheme over ), endowed with the –module structure induced by the structural morphism
As in §7, we define the –motive over associated to by
endowed with the –module structure given by
As an –module, is clearly free of rank (which is called the dimension of the motive). However, as an –module this is not finitely generated, in general, which brings us to the following.
Definition 8.2.
The –module and the –motive are called abelian if is a finitely generated and (automotically, see [3] Chapter 5) free –module. In that case, the rank over of is called the rank of and also called the rank of .
It is not hard to see that if is a Drinfeld module of classical rank , then is –free of rank . So, Drinfeld modules are –dimensional abelian –modules and the two notions of rank coincide.
Assuming now that is an abelian –module as above, of dimension and rank , a prime is called a prime of good reduction for if the –module of dimension , defined over by
(composing with the usual mod reduction map) is abelian of the same rank as .
Now, assume that is an abelian –module as above, of dimension and rank , and let and , with and such that has good reduction at . One can show without difficulty (see [3] Chapter 5) that one has isomorphisms of –modules, for all
Also, the –representation is unramified at all with and the polynomials
are independent of and have coefficients in . (See [4] and the references therein.) It is also true (see [4]) that if is tamely ramified in , then the finite –modules and are free over and therefore the monic polynomials
are well defined in . The abelian –module analogue of Theorem 2.3 is the equality
The above equality was used in [4] to prove an equivariant Tamagawa number formula for the value at of the –equivariant –function associated to the data as above.
In order to prove the above displayed equality in the general case of abelian –modules , one can try to extend to the higher dimensional context the techniques developed in the previous sections of this paper. And, indeed, it is not very difficult to see that the techniques developed in §4, §6 and §7 extend in full to the higher dimensional context. In particular, Propositions 3.2 and 7.1 hold true in general. However, a higher dimensional analogue of Proposition 3.1 does not hold true, in general. A slightly weaker version of Proposition 3.1 (see Theorem 5.6.10 in [3]), which is strong enough for our purposes, holds true for a smaller class of abelian –modules and primes of good reduction , namely those for which (the reduction of mod ) is, in addition, a pure –module. The notion of purity is technical and will not be described here. Instead, we refer the interested reader to §5.5 in [3]. To summarize, if we assume purity of , then the last displayed equality above can be proved with the techniques developed in this paper. If, for example, is a tensor product of Drinfeld modules (in the category of –modules), then purity is achieved. The more general case of arbitrary abelian –motives will be treated in an upcoming paper.
References
- [1] T. Beaumont, On equivariant class number formulas for –modules, https://arxiv.org/abs/2110.15696, 36 pages.
- [2] J. Ferrara, N. Green, Z. Higgins, and C.D. Popescu, An equivariant Tamagawa number formula for Drinfeld modules and applications, J. Algebra and Number Theory 16 (2022), no. 9, 2215–2264.
- [3] D. Goss, Basic structures of function field arithmetic, Springer Berlin, Heidelberg, 1998.
- [4] N. Green and C.D. Popescu, An equivariant Tamagawa number formula for –modules and applications, https://arxiv.org/abs/2206.03541 (June 2022), 27 pages.
- [5] C. Greither and C.D. Popescu, The Galois module structure of -adic realizations of Picard 1-motives and applications, Int. Math. Res. Not. 2012 (2012), no. 5, 986–1036.
- [6] U. Hartl, Private email communication with C.D. Popescu, (2024).
- [7] U. Hartl and R.K. Singh, Local shtukas and divisible local Anderson modules, Canad. J. Math. 71 (2019), no. 5, 1163–1207. MR 4010425
- [8] H. Johnston and A. Nickel, Noncommutative Fitting invariants and improved annihilation results, J. Lond. Math. Soc. (2) 88 (2013), no. 1, 137–160. MR 3092262
- [9] S. Lang, Algebraic groups over finite fields, Amer. J. Math. 78 (1956), 555–563. MR 86367
- [10] G. Laumon, Cohomology of Drinfeld modular varieties. Part I, Cambridge Studies in Advanced Mathematics, vol. 41, Cambridge University Press, Cambridge, 1996, Geometry, counting of points and local harmonic analysis. MR 1381898
- [11] H. Matsumura, Commutative algebra, The Benjamin/Cummings Publishing Co. Inc., Second Edition, 1980.
- [12] N. Ramachandran, Some Fitting ideal computations in Iwasawa theory over and in the theory of Drinfeld modules, Ph.D. thesis, University of California San Diego (2024).
- [13] L. Taelman, Special -values of Drinfeld modules, Ann. of Math. (2) 175 (2012), no. 1, 369–391. MR 2874646