Hypergeometric Functions and Carlitz Differential Equations over Function Fields
Anatoly N. Kochubei Institute of Mathematics, National Academy of Sciences of Ukraine, Tereshchenkivska 3, Kiev, 01601 Ukraine E-mail: kochubei@i.com.ua
Abstract
The paper is a survey of recent results in analysis of additive functions over function fields motivated by applications to various classes of special functions including Thakur's hypergeometric function. We consider basic notions and results of calculus, analytic theory of differential equations with Carlitz derivatives (including a counterpart of regular singularity), umbral calculus, holonomic modules over the Weyl-Carlitz ring. 2000 Mathematics Subject Classification: Primary 11S80, 12H25, 33E50. Secondary 05A40, 11G09, 16S32, 32C38.
1 INTRODUCTION
Let
be the field of formal Laurent series
with coefficients
from the Galois field
,
if
,
,
, where
is a prime number. It is well known that any non-discrete locally compact field of characteristic
is isomorphic to such
. The absolute value on
is given by
,
. The ring of integers
is compact in the topology corresponding to the metric
. The absolute value
can be extended in a unique way onto the completion
of an algebraic closure of
.
Analysis over
and
initiated by Carlitz [3] and developed subsequently by Wagner, Goss, Thakur, the author, and others (see the bibliography in [12, 37] ) is very different from the classical calculus. The new features begin with an appropriate version of the factorial invented by Carlitz – since the usual factorial
, seen as an element of
, vanishes for
, Carlitz introduced the new one as
|
(1.1)
|
An important feature is the availability of many non-trivial
-linear functions, that is such functions
defined on
-subspaces
that
for any
,
. Such are, for example, polynomials and power series of the form
, in particular, the Carlitz exponential
|
(1.2)
|
and its composition inverse, the Carlitz logarithm
|
(1.3)
|
where
(
),
. The notion of the Carlitz exponential obtained a wide generalization in the theory of Drinfeld modules (see [12, 37] ).
Among other special classes of
-linear functions there are various polynomial systems (see below), an analog of the Bessel functions [4, 34] , and Thakur's hypergeometric function [35, 36, 37] .
The latter is defined as follows.
For
,
, denote
|
(1.4)
|
Then, for
, such that the series below makes sense, we set
|
(1.5)
|
Thakur [35, 36, 37] has carried out a thorough investigation of the functions (1.5) and obtained analogs of many properties known for the classical situation. In particular, he found an analog of the hypergeometric differential equation. Its main ingredients are the difference operator
introduced by Carlitz [3] , the nonlinear (
-linear) operator
, and the
-linear Frobenius operator
. For example, the function
is a solution of the equation
|
(1.6)
|
Here we touch only a part of Thakur's results (he considered also hypergeometric functions corresponding to other places of
, a version of (1.5) with parameters from
and its extensions etc).
The Carlitz exponential
satisfies a much simpler equation of the same kind:
so that the operator
may be seen as an analog of the derivative. The operator
is an analog of the multiplication by
, so that
is the counterpart of
.
The same operators appear in the positive characteristic analogs of the canonical commutation relations of quantum mechanics [15, 16] . In the analog of the Schrödinger representation we consider, on the Banach space
of continuous
-linear functions on
, with values from
(with the supremum norm), the “creation and annihilation operators”
(
is the identity operator). Then
|
(1.8)
|
the operator
possesses the orthonormal (in the non-Archimedean sense [32] ) eigenbasis
,
|
(1.9)
|
and
act upon the basis as follows:
|
(1.10)
|
Here
is the sequence of normalized Carlitz polynomials
|
(1.11)
|
which forms an orthonormal basis in
. The spectrum of the “number operator”
is the set of elements
, so that even this notation (proposed by Carlitz in 1935) becomes parallel to the usual quantum mechanical situation.
An analog of the Bargmann-Fock representation is obtained if we consider the operators of almost the same form,
but on the Banach space
of power series
with
,
as
. These new operators satisfy the same relations (1.8)-(1.10), but this time instead of the Carlitz polynomials
we get the eigenfunctions
.
The above results motivated the author to begin to develop analysis and theory of differential equations for
-linear functions over
and
, that is for the case which can be seen as a concentrated expression of features specific for the analysis in positive characteristic. This paper is a brief survey of some achievements in this direction. In particular, we consider the counterparts of the basic notions of calculus, analytic theory of differential equations (in the regular case and the case of regular singularity), their applications to some special functions, like the power function, logarithm and polylogarithms, Thakur's hypergeometric function etc.
An umbral calculus and a theory of holonomic modules are initiated for this case. Like in the classical situation (see [5] ), it is shown that some basic objects of the function field arithmetic generate holonomic modules.
2 Calculus
2.1. Higher Carlitz operators
are introduced recursively,
|
(2.1)
|
For
, the formula (2.1) coincides with the definition of
, if we set
.
The first application of these operators is the reconstruction formula [16] for the coefficients
of a power series
. Note that the classical formula does not make sense here because it contains the expression
where both the numerator and denominator vanish.
Theorem 2.1.
If
, then
For a continuous non-holomorphic
-linear function
the behaviour of the functions
near the origin measures the smoothness of
. We say that
if
can be extended to a continuous function on
. This includes the case (
) of differentiable functions.
The next theorem proved in [16] gives a characterization of the above smoothness in terms of coefficients of the Fourier-Carlitz expansion. It includes, as a particular case (
), the characterization of differentiable
-linear functions obtained by Wagner [41] .
Theorem 2.2.
A function
belongs to
if and only if
In this case
For a generalization to some classes of not necessarily
-linear functions see [43] .
Similarly [16] , a function
is analytic (that is,
) on the ball
if and only if
, as
. A more refined result by Yang [42] , useful in many applications, which makes it possible to find an exact domain of analyticity, is as follows (again we consider only
-linear functions while in [42] a more general class is studied).
Theorem 2.3 (Yang).
A function
is locally analytic if and only if
|
(2.2)
|
and if (2.2) holds, then
is analytic on any ball of the radius
,
2.2. Viewing
as a kind of a derivative, it is natural to introduce an antiderivative
setting
where
is a solution of the equation
, with the normalization
.
It is easy to find
explicitly if
is given by its Fourier-Carlitz expansion (see [16] ).
Next, we introduce a Volkenborn-type integral of a function
(see [32] for a similar integration theory over
) setting
The integral is a
-linear continuous functional on
,
possessing the following “invariance” property (related, in contrast to the case of
, to the multiplicative structure):
Calculating the integrals of some important functions we obtain new relations between them.
In addition to the Carlitz exponential
and the Carlitz polynomials
(see (1.2) and (1.11)), we mention the Carlitz module function
|
(2.3)
|
Note that if
, then only the terms with
are different from zero in (2.3).
We have
For the proofs see [16] .
3 Differential Equations for
-Linear Functions
3.1. Let us consider function field analogs of linear differential equations with holomorphic or polynomial coefficients. Note that in our situation the meaning of a polynomial coefficient is not a usual multiplication by a polynomial, but the action of a polynomial in the operator
.
We begin with the regular case and consider an equation (actually, a system)
|
(3.1)
|
where for each
,
,
|
(3.2)
|
are
matrices with elements from
,
, and it is assumed that the series (3.2) have positive radii of convergence. The action of the operator
upon a vector or a matrix is defined component-wise, so that
for
.
We seek a
-linear solution of (3.1) on some neighbourhood of the origin, of the form
|
(3.3)
|
where
is a given element, so that the “initial” condition for our situation is
|
(3.4)
|
The next theorem, proved in [17] , is the function field analog of the Cauchy theorem from the classical analytic theory of differential equations.
Theorem 3.1.
For any
the equation (3.1) has a unique local solution of the form (3.3), which satisfies (3.4), with the series having a positive radius of convergence.
Thus, regular equations with Carlitz derivatives behave more or less as their classical counterparts. The situation is different for singular equations. Let us consider scalar equations of arbitrary order
|
(3.5)
|
where
,
are power series having (as well as the one for
) positive radii of convergence.
We investigate formal solutions of (3.5), of the form
|
(3.6)
|
One can apply an operator series
(even without assuming its convergence) to a formal series (3.6), setting
and
where the factor
is omitted for
. These formal manipulations are based on the identity
Using also the relation
now we can give a meaning to the notion of a formal solution of the equation (3.5).
Theorem 3.2.
Let
be a formal solution (3.6) of the equation (3.5), where the series for
,
, and
, have positive radii of convergence. Then the series (3.6) has a positive radius of convergence.
This result (proved in [17] ) is in a strong contrast to the classical theory. Note that in the
-adic case a similar phenomenon takes place for equations satisfying certain strong conditions upon zeros of indicial polynomials [1, 6, 25, 33] . In our case such a behavior is proved for any equation, which resembles the (much simpler) case [25] of differential equations over a field of characteristics zero, whose residue field also has characteristic zero.
3.2. The equations (3.1) and (3.5) behave like linear equations, though they are actually only
-linear. Theorem 3.1 can be extended [20] to the case of strongly nonlinear equations (containing self-compositions
).
On the other hand, it is natural to consider some equations of this kind in wider classes of
-linear functions resembling meromorphic functions of a complex variables. The set
of locally convergent
-linear holomorphic functions forms a non-commutative ring with respect to the composition operation (the pointwise multiplication violates the
-linearity). The non-commutativity of
makes the algebraic structures related to Carlitz differential equations much more complicated compared to their classical counterparts. So far their understanding is only at its initial stage. It is known, however, that
can be imbedded into a skew field of
-linear “meromorphic” series containing terms like
(see [20] ). A deep investigation of bi-infinite series of this kind convergent on the whole of
has been carried out by Poonen [24] .
A specific class of equations with solutions meromorphic in the above sense is the class of scalar Riccati-type equations
|
(3.7)
|
where
,
(note that the right-hand side of (3.7) does not contain the linear term). The following theorem is proved in [20] .
Theorem 3.3.
If
,
,
for all
, then the equation (3.7) possesses solutions of the form
where the series converges on the open unit disk
.
4 Regular Singularity
4.1. In analysis over
, a typical class of systems with regular singularity at the origin
over
consists of systems of the form
|
(4.1)
|
where
are constant matrices, and the series converges on a neighbourhood of the origin.
Such a system possesses a fundamental matrix solution of the form
where
is holomorphic on a neighbourhood of zero,
is a constant matrix,
is defined by the obvious power series. Under some additional assumptions regarding the eigenvalues of the matrix
, one can take
. For similar results over
see [9] .
In order to investigate such a class of equations in the framework of
-linear analysis over
, one has to go beyond the class of locally analytic functions. Instead of power series expansions we can use the expansions in Carlitz polynomials on the compact ring
. The property of local analyticity, if it takes place, can be recovered with the use of Theorem 2.3. Note that our approach would fail if we consider equations over
instead of
(our solutions may take their values from
, but they are defined over subsets of
). In this sense our techniques are different from the ones developed for both the characteristic zero cases.
We begin with the simplest model scalar equation
|
(4.2)
|
whose solution may be seen as a function field counterpart of the power function
.
We look for a continuous
-linear solution
of the equation (4.2), with the “initial condition”
, in the form
|
(4.3)
|
where
.
It is easy to see that the equation (4.2) has no continuous solutions if
. If
, then the solution
is unique, continuous on
, and the coefficients from (4.3) have the form
The function
is analytic on
if and only if
for some
; in this case
. If
for any integer
, then
is locally analytic on
if and only if
, and in that case
for
. The relation
holds for all
,
, and for all
. For the proofs see [18] .
Similarly, if in (4.1)
is is a
matrix with elements from
, and we look for a matrix-valued solution of (4.1), then such a solution is given by the series (4.3) with the matrix coefficients
(
is the unit matrix), if
.
4.2. The analog, for our situation, of the system (4.1) is the system
|
(4.4)
|
where
is a matrix-valued analytic function, so that
. We assume that
,
, for all
,
. Denote by
a solution of the equation
.
Let
be the eigenvalues of the matrix
.
Theorem 4.1.
If
|
(4.5)
|
then the system (4.4) has a matrix solution
where the series for
has a positive radius of convergence.
The paper [18] contains, apart from the proof of Theorem 4.1, a discussion of some situations (the Euler type equations) where its conditions are violated, as well as of the meaning of the conditions (4.5). Here we only mention that in the scalar case
the condition (4.5) is equivalent to the assumption
, so that it excludes the case where solutions of the equation
has pathological properties.
4.3. For the above equation, continuous solutions were found as Fourier-Carlitz expansions
|
(4.6)
|
and we had to impose certain conditions upon coefficients of the equation, in order to guarantee the uniform convergence of the series on
(which is equivalent to the fact that
).
However formally we could write the series (4.6) for the solutions without those conditions.
Thus, it is natural to ask whether the corresponding series (4.6) converge at some points
.
Note that (4.6) always makes sense for
(for each such
only a finite number of terms is different from zero). The question is whether the series converges on a wider set; if the answer is negative, such a formal solution is called strongly singular.
The available results regarding strong singularity of solutions of some equations are based on the following general fact [18] .
Theorem 4.2.
If
for all
(where
is some natural number), then the function (4.6) is strongly singular.
It follows from Theorem 4.2 that non-trivial formal solutions of the equation (4.2) with
are strongly singular. A more complicated example is provided by the equation
|
(4.7)
|
for Thakur's hypergeometric function
.
A holomorphic solution of (4.7) is given by an appropriate specialization of (1.5). Classically (over
), there exists the second solution with a logarithmic singularity. Here the situation is different. Looking for a solution of the form (4.6) we obtain a recursive relation
|
(4.8)
|
Taking arbitrary initial coefficients
we obtain a solution
defined on
. On each step we have to solve the equation
If
and
, then in the equation for
we have
.
It can be shown [18] that the equation (4.9) has a unique solution
, for which
, and
other solutions
,
. It is natural to call a solution generic if, starting from a certain step of finding the coefficients
, we always take the most frequent option corresponding to a solution of (4.9) with
. Now Theorem 4.2 implies the following fact.
Theorem 4.3.
A generic solution of the equation (4.7) is strongly singular.
Of course, in some special cases the recursion (4.8) can lead to more regular solutions, in particular, to the holomorphic solutions found by Thakur.
5 Polylogarithms and a Zeta Function
5.1. The Carlitz differential equations can be used for defining new special functions with interesting properties. Some examples are given in this section.
An analog of the function
is defined via the equation
|
(5.1)
|
a counterpart of the classical equation
. The next results are taken from [21] where the equation (5.1) is considered for an arbitrary finite place of
.
Let
be a
-linear holomorphic solution of (5.1) with the zero initial condition (in the sense of (3.4)). Then it is easy to show that
|
(5.2)
|
and the series in (5.2) converges for
.
Note that
is different from the well-known Carlitz logarithm
(see (1.3)). Analogies motivating the introduction of special functions are not so unambiguous, and, for instance, from the composition ring viewpoint,
is an analog of
, though in other respects it is a valuable analog of the logarithm. By the way, another possible analog of the logarithm is a continuous function
,
, satisfying the equation
(an analog of
) and the condition
. In fact,
, the first hyperdifferential operator; see [14] .
Now we consider continuous non-holomorphic extensions of
.
Theorem 5.1.
The equation (5.1) has exactly
continuous solutions on
coinciding with (5.2) as
. These solutions have the expansions in the Carlitz polynomials
where
is an arbitrary solution of the equation
, higher coefficients are found from the relation
and the coefficient
is determined by the relation
Below we denote by
an arbitrary fixed “branch” of extensions of (5.2).
The polylogarithms
are defined recursively by the equations
|
(5.3)
|
which agree with the classical ones
. Analytic
-linear solutions of (5.3), such that
as
, are found easily by induction:
|
(5.4)
|
Theorem 5.2.
For each
, there exists a unique continuous
-linear solution of the equation (5.3) coinciding for
with the polylogarithm (5.4). The solution is given by the Carlitz expansion
with
5.2. Now that the above polylogarithms have been extended onto the disk
, we can interpret their values at
as “special values” of a kind of a zeta function. In order to define the latter, we introduce the operator
,
, a function field analog of the Hadamard fractional derivative
from real analysis (see [31] ).
Denote by
,
,
, the sequence of hyperdifferentiations defined initially on monomials by the relations
,
,
,
where it is assumed that
for
.
is extended onto
by
-linearity, and then onto
by continuity [40] . The sequence
is an orthonormal basis of the space of continuous
-linear functions on
[14, 7] .
Let
,
,
. Denote
. For an arbitrary continuous
-linear function
on
we define its “fractional derivative”
at a point
by the formula
The function
is continuous and
-linear.
Our understanding of
as a kind of a fractional derivative is justified by the following properties:
for any
.
5.3. We define
,
, setting
,
and
if
,
.
is a continuous
-linear function on
with values in
.
In particular, we have
The above definition is of course inspired by the classical polylogarithm relation
In contrast to Goss's zeta function defined on natural numbers and interpolated onto
(see [12, 37] ), the above
is purely an object of the characteristic
arithmetic.
Let us write some relations for special values of our
; for the details see [21] .
As we saw,
Next, let us consider the double sequence
,
,
These elements appear as the coefficients of the expansion [40] of a hyperdifferentiation
in the normalized Carlitz polynomials, as well as in the expression [13] of the operators
from (2.1) via the iterations
. Here we have the identity
which may be seen as a distant relative of Riemann's functional equation for the classical zeta.
Finally, consider the coefficients
of the Carlitz expansion of
(see Theorem 5.1). They are expressed via zeta values:
By Theorem 5.1, for
we have
|
(5.5)
|
The series in (5.5) may be seen as an analog of
. This analogy becomes clearer if, for a fixed
,
, we consider the set
of all convergent power series
corresponding to sequences
. Let us introduce the multiplication
in
setting
and extending the operation distributively (for a similar construction in the framework of
-analysis in characteristic 0 see [23] ). Denoting by
the product in S of elements indexed by prime numbers we obtain in a standard way the identity
(the infinite product is understood as a limit of the partial products in the topology of
), an analog of the Euler product formula.
6 Umbral Calculus
6.1. Classical umbral calculus [29, 28] is a set of algebraic tools for obtaining, in a unified way, a rich variety of results regarding structure and properties of various polynomial sequences. There exists a lot of generalizations extending umbral methods to other classes of functions. However there is a restriction common to the whole literature on umbral calculus – the underlying field must be of zero characteristic. An attempt to mimic the characteristic zero procedures in the positive characteristic case [10] revealed a number of pathological properties of the resulting structures. More importantly, these structures were not connected with the existing analysis in positive characteristic based on a completely different algebraic foundation.
A version of umbral calculus inmplementing such a connection was developed by the author [19] , and we summarize it in this section. Its basic notion is motivated by the following identity for the non-normalized Carlitz polynomials
:
|
(6.1)
|
where the “
-binomial coefficients”
are defined as
Computing the absolute values of the Carlitz factorials directly from their definition (1.1), it is easy to show that
In fact,
, and we can consider also other places of
, that is other non-equivalent absolute values. It can be proved [22] that
belongs to the ring of integers for any finite place of
.
We see the relation (6.1) as a function field counterpart of the classical binomial identity [29, 28] satisfied by many classical polynomials. Now, considering a sequence
of
-linear polynomials with coefficients from
, we call it a sequence of
-binomial type if
and for all
|
(6.2)
|
As in the conventional umbral calculus, the dual notion is that of a delta operator. However, in contrast to the classical situation, here the delta operators are only
-linear, not linear.
Denote by
the operator of multiplicative shift,
. We call a linear operator
, on the
-vector space
of all
-linear polynomials, invariant if it commutes with
for each
.
A
-linear operator
, where
is a linear invariant operator on
, is called a delta operator if
and
for
. A sequence
of
-linear polynomials is called a basic sequence corresponding to a delta operator
, if
,
,
for
,
|
(6.3)
|
or, equivalently,
|
(6.4)
|
It is clear that
is a delta operator. It follows from well-known identities for the Carlitz polynomials
[11] (see also (1.10)) that the sequence
is basic with respect to the operator
.
Theorem 6.1.
For any delta operator
, there exists a unique basic sequence
, which is a sequence of
-binomial type. Conversely, given a sequence
of
-binomial type, define the action of
on
by the relations (6.4), extend it onto
by linearity and set
. Then
is a delta operator, and
is the corresponding basic sequence.
The analogs of the higher Carlitz difference operators (2.1) in the present general context are the operators
. The identity
|
(6.5)
|
holds for any
. If
is a
-linear polynomial,
, then a generalized Taylor formula
|
(6.6)
|
holds for any
. For the Carlitz polynomials
, the formulas (6.5) and (6.6) are well known [11] . It is important that, in contrast to the classical umbral calculus, the linear operators involved in (6.6) are not powers of a single linear operator.
Any linear invariant operator
on
admits a representation
|
(6.7)
|
The infinite series in (6.7) becomes actually a finite sum if both sides of (6.7) are applied to any
-linear polynomial. Conversely, any such series defines a linear invariant operator on
.
Let us consider the case where
, so that
. The next result leads to new delta operators and basic sequences.
Theorem 6.2.
The operator
, where
is a delta operator if and only if
|
(6.8)
|
Example 1. Let
for all
, that is
|
(6.9)
|
Estimates of
which follow directly from (1.1) show that
, so that (6.8) is satisfied. Comparing (6.9) with a classical formula from [29] we may see the polynomials
for this case as analogs of the Laguerre polynomials.
Example 2. Let
. For this case it can be shown [19] that
,
;
for all
(of course,
), and
,
for
.
6.2. As in the
-adic case [38, 39, 27] , the umbral calculus can be used for constructing new orthonormal bases in
.
Let
be the basic sequence corresponding to a delta operator
,
|
(6.10)
|
The sequence
,
, called the normalized basic sequence, satisfies the identity
another form of the
-binomial property. Though it resembles its classical counterpart, the presence of the Frobenius powers is a feature specific for the case of a positive characteristic.
Theorem 6.3.
If
,
for
, then the sequence
is an orthonormal basis of the space
– for any
there is a uniformly convergent expansion
where
,
as
,
By Theorem 6.3, the Laguerre-type polynomial sequence from Example 1 is an orthonormal basis of
. The sequence from Example 2 does not satisfy the conditions of Theorem 6.3.
Note that the conditions of Theorem 6.3 imply that
for all
, so that the series (6.10) considered in Theorem 6.3 always correspond to delta operators.
In [19] recursive formulas and generating functions for normalized basic sequences are also given.
7 The Weyl-Carlitz Ring and Holonomic Modules
7.1. 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, [2, 5, 30] ).
It is clear from the above results 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
|
(7.1)
|
The ring consists of finite sums
|
(7.2)
|
and the representation of an element in the form (7.2) is unique.
The next result proved in [17] shows that basic algebraic properties of
are similar to those of the Weyl algebra in characteristic 0 and quite different from the case of the algebra of usual differential operators over a field of positive characteristic [26] .
Theorem 7.1.
The ring
is left and right Noetherian, without zero divisors; its centre coincides with
.
possesses no non-trivial two-sided ideals stable with respect to the mapping
A generalization of
to the case of several variables is not straightforward because the Carlitz derivatives
and
do not commute on a monomial
, if
.
Moreover, if
, then
is not a polynomial, nor even a holomorphic function in
(since the action of
is not linear and involves taking the
-th root).
A reasonable generalization is inspired by Zeilberger's idea (see [5] ) 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
|
(7.3)
|
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.
More generally, denote by
the set of all germs of functions of the form
|
(7.4)
|
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 (7.4) in which only a finite number of coefficients is different from zero.
The ring
is generated by the operators
on
, and the operators of multiplication by scalars from
. To simplify the notation, we write
instead of
and identify a scalar
with the operator of multiplication by
. The operators
are
-linear, so that
|
(7.5)
|
while the operators
satisfy the commutation relations (7.1). 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. We have the relations involving
:
|
(7.6)
|
Using the commutation relations (7.1), (7.5), and (7.6), we can write any
, in a unique way, as a finite sum
|
(7.7)
|
Let us introduce a filtration in
(an analog of the Bernstein filtration) denoting by
,
, the
-vector space of operators (7.7) with
where the maximum is taken over all the terms of (7.7). Then
is a left and right Noetherian filtered ring.
In a standard way (see [8] ) we define filtered left modules over
. All the basic notions regarding a filtered module
(like those of the graded module
, dimension
, multiplicity
, good filtration etc) are introduced just as their counterparts in the theory of modules over the Weyl algebra.
If we consider
as a left module over itself, then
|
(7.8)
|
For any finitely generated left
-module
, we have
. By (7.8), this bound cannot be improved in general. However, if
is a non-zero left ideal in
, then
|
(7.9)
|
For the module
of
-linear polynomials (7.4), we have
The proofs of all these results, as well as the ones given in this section below, can be found in [22] .
It is natural to call an
-module
holonomic if
. Thus,
is an example of a holonomic module.
The next theorem demonstrates, already for the case of
-modules, a sharp difference from the case of modules over the Weyl algebras. In particular, we see that an analog of the Bernstein inequality (see [8] ) does not hold here without some additional assumptions.
Theorem 7.2.
-
(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
.
7.3. Let us consider the case of holonomic submodules of the
-module
, consisting of
-linear functions (7.4) polynomial in
and holomorphic near the origin in
.
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
.
As we know (see (7.9)), if
, then
. We call a function
holonomic if the module
is holonomic, that is
. The condition
means that
is a solution of a non-trivial “differential equation”
,
. The case
is quite simple.
Theorem 7.3.
If a non-zero function
satisfies an equation
,
, then
is holonomic.
In particular, any
-linear polynomial of
is holonomic, since it is annihilated by
, with a sufficiently large
.
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). The simplest example of a degenerate function (for
) is
where the function
belongs to
and satisfies an equation
,
. It can be shown that
.
In order to exclude the degenerate case, we introduce the notion of a non-sparse function.
A function
of the form (7.4) is called non-sparse if there exists such a sequence
that, for any
, there exist sequences
(depending on
), such that
as
(
), and
.
Theorem 7.4.
If a function
is non-sparse, then
. If, in addition,
satisfies an equation
,
, then
is holonomic.
7.4 We use Theorem 7.4 to prove that the functions (7.4) obtained via the sequence-to-function transform (7.3) or its multi-index generalizations, from some well-known sequences of polynomials over
are holonomic. In all the cases below the non-sparseness is evident, and we have only to prove that the corresponding function satisfies a non-trivial Carlitz differential equation.
a) The Carlitz polynomials. The transform (7.3) of the sequence
is the Carlitz module function
; see (2.3). It is easy to check that
. 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 (1.5), that is
|
(7.10)
|
where
. It is seen from (1.4) that the terms in (7.10), which make sense and do not vanish, are those with
. Let the function
be given by
It is known ([37] , Sect. 6.5) that
|
(7.11)
|
if all the parameters
are different from zero. If at least one of them is equal to zero, then the left-hand side of (7.11) equals zero. This property implies the identity
, the same as that for the Carlitz module function. Thus,
is holonomic.
c).
-binomial coefficients. It can be shown [22] that the
-binomial coefficients
(see Sect. 6) satisfy the Pascal-type identity
|
(7.12)
|
where
and it is assumed that
.
Consider a function
associated with the
-binomial coefficients, that is
|
(7.13)
|
The identity (7.12) implies the equation
for the function (7.13). Therefore
is holonomic.
Acknowledgements This work was supported in part by CRDF (Grant UM1-2567-OD-03), DFG (Grant 436 UKR 113/72), and the Ukrainian Foundation for Fundamental Research (Grant 01.07/027).
References
-
F. Baldassarri, Differential modules and singular points of
-adic differential equations, Adv. Math. 44 (1982), 155–179.
-
Yu. Berest and A. Kasman,
-modules and Darboux transformations, Lett. Math. Phys. 43 (1998), 279–294.
-
L. Carlitz, On certain functions connected with polynomials in a Galois field, Duke Math. J. 1 (1935), 137–168.
-
L. Carlitz, Some special functions over
, Duke Math. J. 27 (1960), 139–158.
-
P. Cartier, Démonstration “automatique” d'identités et fonctions hypergéometriques (d'après D. Zeilberger), Astérisque 206 (1992), 41–91.
-
D. N. Clark, A note on the
-adic convergence of solutions of linear differential equations, Proc. Amer. Math. Soc. 17 (1966), 262–269.
-
K. Conrad, The digit principle, J. Number Theory 84 (2000), 230–237.
-
S. C. Coutinho, A Primer of Algebraic D-modules, Cambridge University Press, 1995.
-
B. Dwork, G. Gerotto, and F. J. Sullivan, An Introduction to
-Functions, Princeton University Press, 1994.
-
L. Ferrari, An umbral calculus over infinite coefficient fields of positive characteristic, Comp. Math. Appl. 41 (2001), 1099–1108.
-
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.
-
S. Jeong, Continuous linear endomorphisms and difference equations over the completions of
, J. Number Theory 84 (2000), 276–291.
-
S. Jeong, Hyperdifferential operators and continuous functions on function fields, J. Number Theory 89 (2001), 165–178.
-
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.
-
A. N. Kochubei, Strongly nonlinear differential equations with Carlitz derivatives over a function field, Ukrainian Math. J. (to appear), math.NT/ 0405542.
-
A. N. Kochubei, Polylogarithms and a zeta function for finite places of a function field, Contemporary Math. (to appear); math.NT/0405544.
-
A. N. Kochubei, Holonomic modules in positive characteristic, math.RA/0503398.
-
M. B. Nathanson, Additive number theory and the ring of quantum integers, math.NT/0204006.
-
B. Poonen, Fractional power series and pairings on Drinfeld modules, J. Amer. Math. Soc. 9 (1996), 783–812.
-
M. van der Put, Meromorphic differential equations over valued fields, Indag. Math. 42 (1980), 327–332.
-
M. van der Put, Differential equations in characteristic
, Compositio Math. 97 (1995), 227–251.
-
A. M. Robert, A Course in
-Adic Analysis, Springer, New York, 2000.
-
S. Roman, The Umbral Calculus, Academic Press, London, 1984.
-
G.-C. Rota, D. Kahaner and A. Odlyzko, On the foundations of combinatorial theory. VIII. Finite operator calculus, J. Math. Anal. Appl. 42 (1973), 684–760.
-
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.
-
S. G. Samko, A. A. Kilbas, and O. I. Marichev, Fractional Integrals and Derivatives: Theory and Applications, Gordon and Breach, New York, 1993.
-
W. Schikhof, Ultrametric Calculus, Cambridge University Press, 1984.
-
M. Setoyanagi, Note on Clark's theorem for
-adic convergence, Proc. Amer. Math. Soc. 125 (1997), 717–721.
-
D. Sinnou and D. Laurent, Indépendence algebrique sur les
-modules, Compositio Math. 122 (2000), 1–22.
-
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.
-
L. Van Hamme, Continuous operators, which commute with translations, on the space of continuous functions on
. In:
-Adic Functional Analysis (J. M. Bayod et al., eds.), Lect. Notes Pure Appl. Math. 137, Marcel Dekker, New York, 1992, pp. 75–88.
-
A. Verdoodt, Umbral calculus in non-Archimedean analysis. In:
-Adic Functional Analysis (A. K. Katsaras et al., eds.), Lect. Notes Pure Appl. Math. 222, Marcel Dekker, New York, 2001, pp. 309–322.
-
J. F. Voloch, Differential operators and interpolation series in power series fields, J. Number Theory 71 (1998), 106–108.
-
C. G. Wagner, Linear operators in local fields of prime characteristic, J. Reine Angew. Math. 251 (1971), 153–160.
-
Z. Yang, Locally analytic functions over completions of
, J. Number Theory 73 (1998), 451–458.
-
Z. Yang,
-functions over completions of
at finite places of
, J. Number Theory 108 (2004), 346–374.