Holonomic Modules in Positive Characteristic
Anatoly N. Kochubei
*
*
Partially supported by CRDF under Grant UM1-2567-OD-03 Institute of Mathematics, National Academy of Sciences of Ukraine, Tereshchenkivska 3, Kiev, 01601 Ukraine E-mail: kochubei@i.com.ua
Abstract
We study modules over the Weyl-Carlitz ring, a counterpart of the Weyl algebra in analysis over local fields of positive characteristic. It is shown that some basic objects of function field arithmetic, like the Carlitz module, Thakur's hypergeometric polynomials, and analogs of binomial coefficients arising in the function field version of umbral calculus, generate holonomic modules. Key words:
-linear function; holonomic module; holonomic function; Carlitz derivative
1 INTRODUCTION
The theory of holonomic modules over the Weyl algebra and more general algebras of differential or
-difference operators is becoming increasingly important, both as a crucial part of the general theory of D-modules and in view of various applications (see, for example, [1, 6, 10, 22] ). Well-known pathological properties of differential operators over fields of positive characteristic make the available, for this case, analogs of the theory of D-modules much more complicated [3, 19] . More importantly, the resulting structures are not connected with the existing analysis in positive characteristic based on a completely different algebraic foundation.
Any non-discrete locally compact field of a positive characteristic
is isomorphic to the field
of formal Laurent series with coefficients from the Galois field
,
,
. The field
is endowed with a non-Archimedean absolute value as follows. If
,
then
. This valuation can be extended onto the field
, the completion of an algebraic closure of
.
Analysis over
and
, which was initiated in the great paper by Carlitz [5] and developed subsequently by Wagner, Goss, Thakur, the author, and many others (see the bibliography in [9, 25] ) is very different from the classical calculus. An important feature is the availability of many non-trivial additive (actually,
-linear) polynomials and power series of the form
.
Taking into account the fact that the usual factorial
, seen as an element of
, vanishes for
, Carlitz introduced the new factorial
|
(1)
|
the
-linear logarithm and exponential (which obtained a wide generalization later, in the theory of Drinfeld modules), as well as an important polynomial system, the Carlitz polynomials. Subsequently many other
-linear special functions, such as Thakur's hypergeometric function [23, 24, 25] and further special polynomial systems, were introduced and investigated. The difference operator
|
(2)
|
introduced in [5] became the main ingredient of the
-linear calculus and analytic theory of differential equations over
developed in [13, 14, 15] . The role of a derivative is played by the
-linear operator
(the Carlitz derivative). The latter appears also in the
-linear umbral calculus [16] where an important role belongs to the following new analog of binomial coefficients
|
(3)
|
The meaning of a polynomial coefficient in a differential equation of the above type is not a usual multiplication by a polynomial, but the action of a polynomial in the Frobenius operator
,
.
The above developments show that in the positive characteristic case a natural counterpart of the Weyl algebra is, for the case of a single variable, the ring
generated by
, and scalars from
, with the relations [12]
|
(4)
|
Some algebraic properties of
were studied in [14] – it is left and right Noetherian, with no zero divisors, its center coincides with
.
The aim of this paper is to initiate the dimension theory for modules over
and more general “several variable” rings. The definition of the latter is not straightforward. If, for example, we consider the natural action of the Carlitz derivatives
and
on an
-linear monomial
, we notice immediately that
is not a polynomial, nor even a holomorphic function in
, if
(since the action of
is not linear and involves taking the
-th root). Moreover, it follows from the relation
and the last commutation relation in (4) that
and
do not commute even on monomials
with
.
A reasonable generalization is inspired by Zeilberger's idea (see [6] ) to study holonomic properties of sequences of functions making a transform with respect to the discrete variables, which reduces the continuous-discrete case to the purely continuous one (simultaneously in all the variables). In our situation, if
is a sequence of
-linear polynomials with
, we set
and
is well-defined. In the variable
, we consider not
but the linear operator
. The latter does not commute with
either, but satisfies the commutation relations
so that the resulting ring
resembles a universal enveloping algebra of a solvable Lie algebra.
Similarly we define
for
.
Introducing in
an analog of the Bernstein filtration and considering filtered modules over
, we find that basic principles of the theory of algebraic D-modules [7] carry over to this case without serious complications. However, the nonlinearity of
and
brings new phenomena. In particular, already the ring
possesses non-trivial finite-dimensional representations. Therefore an analog of the Bernstein inequality does not hold here without some additional assumptions.
The case we study in a greater detail is that of holonomic submodules of the
-module of
-linear functions
, polynomial in
and holomorphic near the origin in
.
Following [6] we call a function
holonomic if such is the module
. We prove general conditions for a function
to be holonomic and verify them for basic objects of this branch of analysis – the Carlitz polynomials, Thakur's hypergeometric polynomials, and the
-binomial coefficients (3), making the above transition (*) from discrete variables to continuous ones.
Considering the
-binomial coefficients we use this occasion to prove also the fact that they belong to the ring of integers not only for the field
, but for any place of the global function field
. Together with the results of [16] , this property supports the case for considering the expressions (3) as “proper” analogs of the classical binomial coefficients. For other analogs of the latter see [25] .
2 The Weyl-Carlitz Ring
2.1. Denote by
the set of all germs of functions of the form
|
(5)
|
where
are such that all the series are convergent on some neighbourhoods of the origin. We do not exclude the case
where
will mean the set of all
-linear power series
convergent on a neighbourhood of the origin.
will denote the set of all polynomials from
, that is the series (5) in which only a finite number of coefficients is different from zero.
The ring
is generated by the operators
on
defined in the Introduction, and the operators of multiplication by scalars from
. To simplify the notation, we will write
instead of
and identify a scalar
with the operator of multiplication by
. The operators
are
-linear, so that
|
(6)
|
while the operators
satisfy the commutation relations (4). In the action of each operator
(acting in a single variable), other variables are treated as scalars. The operator
acts simultaneously on all the variables and coefficients, so that
It follows from (2) that
|
(7)
|
the second equality can be included in the first one, if we set
. Similarly
|
(8)
|
Since
for any
, the action of operators from
does not spoil convergence of the series (5).
The identity
, together with (7) and (8), implies the commutation relations
|
(9)
|
verified by applying both sides of each equality to an arbitrary monomial.
Using the commutation relations (4), (6), and (9), we can write any element
as a finite sum
|
(10)
|
Proposition 1.
The representation (10) of an element
is unique.
Proof. Suppose that
|
(11)
|
Applying the left-hand side of (11) to the function
with
we find that
whence
for each
. Writing this in the form
|
(12)
|
where
and taking into account that (12) holds for arbitrary
, that is for an infinite set of values of
, we find that
. Repeating this reasoning we get the equality
for all
.
Suppose that
for
and arbitrary
. Then we apply the left-hand side of (11) to the function
and proceed as before coming to the equality
for all
.
It is easy to prove by induction with respect to
(using the commutation relations (9) and the result from [14] regarding the case
) that
has no zero-divisors.
2.2. Let us introduce a filtration in
denoting by
,
, the
-vector space of operators (10) with
where the maximum is taken over all the terms contained in the representation (10). It is clear that
is a filtered ring (for the definitions see [20] ). Setting
,
,
, we introduce the associated graded ring
It is generated by scalars
and the images
of the elements
respectively, which satisfy, by virtue of (4), (6), and (9), the relations
It is clear that
is a (left and right) almost normalizing extension of the field
(see Chapter 1, §6 in [20] ), so that the rings
and
are left and right Noetherian. Let us compute the dimension of the
-vector space
. Note that
so that
coincides with the dimension of the appropriate space appearing in the natural filtration in
.
Lemma 1.
For any
Proof. The number
coincides with the number of non-negative integral solutions
of the inequality
, so that
where
is the number of different representations of
as sums of
non-negative integers.
It is known (Proposition 6.1 in [17] ) that
. Then (see Sect. 1.3 from [21] )
as desired.
3 Filtered Modules
3.1. Let
be a left module over the Weyl-Carlitz ring
. Suppose we have a filtration
of
, that is
|
(13)
|
and
for any
. We assume that each
is a finite-dimensional vector space over
. Below we write
and
if
and
.
In a standard way [7] we define the graded module
over
, associated with the filtration (13). As usual, the filtration (13) is called good, if
is finitely generated.
Main properties of filtered modules over the Weyl algebra (see [2, 7] ) carry over to our situation without any substantial changes, both in their formulations and proofs. In fact, the only technical difference is that the operators
and
are semilinear, not linear. However, as it is explained in Appendix I to Chapter 2 of [4] , basic notions of linear algebra remain valid for semilinear mappings – a semilinear mapping of a vector space into itself can be interpreted as a linear mapping between two different vector spaces, and, for instance, dimensions of the kernel and cokernel are not changed in this interpretation. Note that everywhere in this paper we consider vector spaces over the algebraically closed field
, on which
induces an automorphism. Below, as before,
means the dimension over
.
In particular, for a good filtration there exist a polynomial
and a number
, such that
The number
, called the (Gelfand-Kirillov) dimension of
, and the leading coefficient of
multiplied by
, called the multiplicity
of
, do not depend on the choice of a good filtration on
. A filtration
is good if and only if there exists such
that
If
and
are a submodule and the corresponding quotient module, with the induced filtrations, then
, and if
, then
. For a direct sum
we have
.
In particular, if we consider
as a feft module over itself, then by Lemma 1
|
(14)
|
It follows from (14) and the above general facts that for any finitely generated left
-module
By (14), the bound in (15) in general cannot be improved. However, if
is a non-zero left ideal in
, then
|
(16)
|
The proof of (16) is identical to the proof of Corollary 9.3.5 from [7] .
3.2. Let us consider the set
of polynomials (5) as a
-module. A filtration
can be introduced by setting
to be the collection of all the polynomials (5), in which the maximal indices
corresponding to non-zero coefficients
do not exceed
. This filtration is obviously good.
Proposition 2.
For the module
,
|
(17)
|
Proof. Let us compute
. For a fixed
, the quantity of
-tuples
of non-negative integers, for which
, is added up from those
-tuples where
numbers are equal to
while
numbers are strictly larger and can take
values.
Therefore the above quantity equals
. Next,
possible values of
in (5) correspond to each
-tuple. Thus,
Denote
,
. Performing the Abel transformation we get
where
.
It is known ([11] , Chapter 15) that
where
are the Bernoulli numbers. Therefore we find that
where
is a polynomial of the degree
. This implies (17).
It is natural to call an
-module
holonomic if
. Thus,
is an example of a holonomic module.
3.3. Let us look at possible values of
for
-modules. The next result demonstrates a sharp difference from the case of modules over the Weyl algebras.
Theorem 1.
-
(i) For any
, there exists such a nontrivial
-module
that
(
means the dimension over
), that is
.
-
(ii) Let
be a finitely generated
-module with a good filtration. Suppose that there exists a “vacuum vector”
, such that
and
for all
. Then
.
Proof. (i) Let
. Denote by
the standard basis in
, that is
, with 1 at the
-th place. Let
be a
matrix over
, such that
if
, while the diagonal elements satisfy the equation
. We define the action of
and
on
as follows:
with subsequent additive continuation onto
.
If
,
, then we have
so that
and we have indeed an
-module.
(ii) It follows from the relation
(see [14] ) that
that is
is an eigenvector of a linear operator
on
(considered as a
-vector space) corresponding to the eigenvalue
. Therefore the vectors
are linearly independent.
It follows from the existence of the Hilbert polynomial
implementing the dimension
that
.
4 Holonomic Functions
4.1. Let
,
is a left ideal in
. The left
-module
is isomorphic to the submodule
– an element
corresponds to the class of
in
. A natural good filtration in
is induced from that in
– the subspace
is generated by elements
with
.
As we know (see (16)), if
, then
. We call a function
holonomic if the module
is holonomic, that is
. The condition
means that
is a solution of a “differential equation”
,
. For
, we have the following easy result.
Theorem 2.
If a non-zero function
satisfies an equation
,
, then
is holonomic.
Proof. It is sufficient to show that
. In fact, the sequence
is linearly independent because otherwise we would have such a finite collection of elements
, some of which are different from zero, that
|
(18)
|
for all
from a neighbourhood of the origin in
. It follows from (18) that
takes only a finite number of values. By the uniqueness theorem for non-Archimedean holomorphic functions,
on some neighbourhood of the origin. Due to the
-linearity,
, and we have come to a contradiction.
In particular, any
-linear polynomial of
is holonomic, since it is annihilated by
, with a sufficiently large
.
4.2. If
, the situation is more complicated. We call the module
(and the corresponding function
) degenerate if
(by the Bernstein inequality, there is no degeneracy phenomena for modules over the complex Weyl algebra). We give an example of degeneracy for the case
.
Let
where the function
belongs to
and satisfies an equation
,
. Then
is degenerate.
Indeed, by the general rule,
is spanned by elements
with
. In the present situation,
so that an element
is a linear combination of elements
with
,
. Therefore
is contained in the linear hull of elements
,
. By Theorem 2, the
-dimension of the latter does not exceed a linear function of
, so that
. On the other hand, since, as in the proof of Theorem 2, the system of functions
is linearly independent, we find that
.
In order to exclude the degenerate case, we introduce the notion of a non-sparse function.
A function
of the form (5) is called non-sparse if there exists such a sequence
that, for any
, there exist sequences
(depending on
), such that
as
(
), and
.
Lemma 2.
If a function
is non-sparse, then the system of functions
(
) is linearly independent over
.
Proof. Suppose that
|
(19)
|
for some
,
. Substituting (5) into (19) and collecting coefficients of the power series we find that
|
(20)
|
for all
.
We see from (20) that the polynomial
has an infinite sequence of different roots, so that
for all
, and for each
. Repeating this reasoning we find that all the coefficients
are equal to zero.
Now the above arguments regarding
yield the following result.
Theorem 3.
If a function
is non-sparse, then
. If, in addition,
satisfies an equation
,
, then
is holonomic.
As in the classical situation, one can construct holonomic functions by addition.
Proposition 3.
If the functions
are holonomic, and
is non-sparse, then
is holonomic.
Proof. Consider the
-module
. Since
and
are both holonomic, we have
. Next, let
be a submodule of
consisting of such pairs
that
. Then
, so that
.
On the other hand, we have an injective mapping
, which maps
to the image of
in
. Therefore
.
It remains to use Theorem 3.
4.3. We use Theorem 3 to prove that the functions (5) obtained via the sequence-to-function transform (
) or its multi-index generalizations, from some well-known sequences of polynomials over
are holonomic.
a) The Carlitz polynomials. The sequence
of normalized Carlitz polynomials forms an orthonormal basis of the space of all
-linear continuous functions on the ring of integers of the field
. Its transform (
), the function
|
(21)
|
called the Carlitz module, is one of the main objects of the function field arithmetic [9, 25] .
It is known [5, 8] that
where
(
),
. By (1), we have
|
(22)
|
so that
For large values of
, an elementary investigation of the function
,
, shows that
so that
for all
with
. Therefore the series (21) converges for small
, so that the Carlitz module function belongs to
.
Since
for
, and
[8] , we see that
. Clearly, the function
is non-sparse. Therefore the Carlitz module function is holonomic, jointly in both its variables.
b) Thakur's hypergeometric polynomials. We consider the polynomial case of Thakur's hypergeometric function [23, 24, 25] :
|
(23)
|
where
,
|
(24)
|
It is seen from (24) that the terms in (23), which make sense and do not vanish, are those with
. Let
|
(25)
|
We prove as above that all the series in (25) converge near the origin. Thus,
.
It is known ([25] , Sect. 6.5) that
|
(26)
|
if all the parameters
are different from zero. If at least one of them is equal to zero, then the left-hand side of (26) equals zero. This property implies the identity
, the same as that for the Carlitz module function. Since
is non-sparse, it is holonomic.
In the next section we will see that the
-binomial coefficients (3) correspond to a holonomic function satisfying a more complicated equation containing also the operator
.
5
-Binomial Coefficients
5.1. Let us consider the
-binomial coefficients (3). It follows from (22) that
Since
, it is natural to consider also other places of
, that is other non-equivalent absolute values on
. It is well known ([26] , Sect. 3.1) that they are parametrized by monic irreducible polynomials
. The absolute value
,
, is defined as follows. We write
where
,
, and
does not divide
. Then
,
where
; as usual,
. The absolute value
used elsewhere in this paper corresponds to
.
Proposition 4.
For any monic irreducible polynomial
, the
-binomial coefficients (3) satisfy the inequality
Proof. First we compute
. It follows from Lemma 2.13 of [18] that
Writing
, with
,
, we find that
Similarly we can write
, with
,
, and get that
If
, then we obtain a similar representation for
simply by adding those for
and
, so that
If
, then
where
. In this case
since
.
Below we will use only the valuation with
, that is, as above, consider the field
.
5.2. Let us derive, for the
-binomial coefficients (3), analogs of the classical Pascal and Vandermonde identities.
Proposition 5.
The identity
|
(27)
|
holds, if
and it is assumed that
.
Proof. Let
be the “non-normalized” Carlitz polynomials. They satisfy the main
-binomial identity [5, 16]
|
(28)
|
which holds, for example, for any
.
It is known [5, 8] that
|
(29)
|
Let us rewrite the left-hand side of (28) in accordance with (29), and apply to each term the identity (28) with
substituted for
. We have
By (29),
,
, whence
Note that
|
(30)
|
Indeed, the left-hand side of (30) equals
and coincides with the right-hand side. Therefore the last sum in the expression for
equals
Using (29) again we find that
and the comparison with (28) yields
for any
.
Since the Carlitz polynomials are linearly independent, we obtain that
for any
, and it remains to note that
if
,
, by the definition of the Carlitz polynomials.
More generally, we have the following Vandermonde-type identity. Let
be integers,
.
Proposition 6.
Define
by the recurrent relation
|
(31)
|
and the initial conditions
for
and
,
. Then, for any
,
|
(32)
|
Proof. The identity (32) is trivial for
. Suppose it has been proved for some
. Let us transform the right-hand side of (32) using the identity (27). Then we have
Since we assume that
, the summation in both the above sums can be performed from 0 to
. Using (31) we obtain the required identity (32) with
substituted for
.
5.3. Now we consider a function
associated with the
-binomial coefficients, that is
|
(33)
|
Obviously,
is non-sparse.
Proposition 7.
The function (33) satisfies the equation
|
(34)
|
so that
is holonomic.
Proof. Let us compute
. We have
Using Proposition 5 we find that
where
Note that
so that
|
(35)
|
Next, we have
and also
whence
Therefore
As above,
, so that
Together with (35), this implies (34).
References
-
Yu. Berest and A. Kasman,
-modules and Darboux transformations, Lett. Math. Phys. 43 (1998), 279–294.
-
J.-E. Björk, Rings of Differential Operators, North-Holland, Amsterdam, 1979.
-
R. Bögvad, An analogue of holonomic D-modules on smooth varieties in positive characteristics, Homology Homotopy Appl. 4, No. 2 (2002), 83–116.
-
N. Bourbaki, Algebra I, Springer, Berlin, 1989.
-
L. Carlitz, On certain functions connected with polynomials in a Galois field, Duke Math. J. 1 (1935), 137–168.
-
P. Cartier, Démonstration “automatique” d'identités et fonctions hypergéometriques (d'après D. Zeilberger), Astérisque 206 (1992), 41–91.
-
S. C. Coutinho, A Primer of Algebraic D-modules, Cambridge University Press, 1995.
-
D. Goss, Fourier series, measures, and divided power series in the theory of function fields, K-Theory 1 (1989), 533–555.
-
D. Goss, Basic Structures of Function Field Arithmetic, Springer, Berlin, 1996.
-
M. A. Guest, Quantum cohomology via D-modules, math.DG/0206212.
-
K. Ireland and M. Rosen, A Classical Introduction to Modern Number Theory, Springer, New York, 1982.
-
A. N. Kochubei, Harmonic oscillator in characteristic
, Lett. Math. Phys. 45 (1998), 11–20.
-
A. N. Kochubei,
-linear calculus over function fields, J. Number Theory 76 (1999), 281–300.
-
A. N. Kochubei, Differential equations for
-linear functions, J. Number Theory 83 (2000), 137–154.
-
A. N. Kochubei, Differential equations for
-linear functions II: Regular singularity, Finite Fields Appl. 9 (2003), 250–266.
-
A. N. Kochubei, Umbral calculus in positive characteristic, Adv. Appl. Math. 34 (2005), 175–191.
-
S. K. Lando, Lectures on Generating Functions, AMS, Providence, 2003.
-
R. Lidl and H. Niederreiter, Finite Fields, Addison-Wesley, Reading, 1983.
-
G. Lyubeznik, F-modules: applications to local cohomology and D-modules in characteristic
, J. Reine Angew. Math. 491 (1997), 65–130.
-
J. C. McConnell and J. C. Robson, Noncommutative Noetherian Rings, AMS, Providence, 2001.
-
J. Riordan, Combinatorial Identities, Wiley, New York, 1968.
-
C. Sabbah, Systèmes holonomes d'équations aux q-differences. In: “D-modules and Microlocal Geometry” (M. Kashiwara et al., eds.), Walter de Gruyter, Berlin, 1993, pp. 125–147.
-
D. S. Thakur, Hypergeometric functions for function fields, Finite Fields and Their Appl. 1 (1995), 219–231.
-
D. S. Thakur, Hypergeometric functions for function fields II, J. Ramanujan Math. Soc. 15 (2000), 43–52.
-
D. S. Thakur, Function Field Arithmetic, World Scientific, Singapore, 2004.
-
A. Weil, Basic Number Theory, Springer, Berlin, 1967.