Constructing reductions for creative telescoping
The general differentially finite case

Joris van der Hoeven

Laboratoire d'informatique, UMR 7161 CNRS

Campus de l'École polytechnique

1, rue Honoré d'Estienne d'Orves

Bâtiment Alan Turing, CS35003

91120 Palaiseau

Email: vdhoeven@lix.polytechnique.fr

Revised version, November 4, 2023

The class of reduction-based algorithms was introduced recently as a new approach towards creative telescoping. Starting with Hermite reduction of rational functions, various reductions have been introduced for increasingly large classes of holonomic functions. In this paper we show how to construct reductions for general holonomic functions, in the purely differential setting.

Keywords: creative telescoping, holonomic function, Hermite reduction, residues

A.C.M. subject classification:

I.1.2 Algebraic algorithms

A.M.S. subject classification: 33F10, 68W30

1.Introduction

Let be an effective field of characteristic zero and let be a non-zero polynomial. Consider the system of linear differential equations

(1.1)

where is an matrix with entries in and is a column vector of unknown functions. Notice that any system of linear differential equations with can be rewritten in this form by taking to be a multiple of all denominators.

Let be a formal solution of (1.1) and consider the -module of linear combinations where is a row vector. Then has the natural structure of a D-module for the derivation defined by

A -linear mapping is said to be a reduction for (1.1) if for all . Such a reduction is said to be confined if its image is a finite dimensional subspace of over and normal if for all . In this paper, we propose a solution to the following problem:

Problem 1.1. Design an algorithm that takes the equation (1.1) as input and that returns a (possibly normal) confined reduction for (1.1), in the form of an algorithm for the evaluation of .

Confined reductions are interesting for their application to creative telescoping. After its introduction by Zeilberger in [23], the theory of creative telescoping has known a rapid development. For a brief history of the topic and further references, we point the reader to [12]. In section 2 we recall how confined reductions can be used for the computation of so-called telescopers; see also [7, 14]. It is worth to notice that Problem 1.1 concerns univariate differential equations, whereas creative telescoping is a priori a multivariate problem.

Reduction-based algorithms have appeared recently as a particularly promising approach in order to make creative telescoping more efficient and to understand its complexity. The simplest kind of reduction is Hermite reduction [22, 16, 2, 7], in which case and . More precisely [14, Proposition 21], given with , and writing for the square-free part of , there exist unique with such that

Then the Hermite reduction of is defined by . Confined reductions have been constructed in increasingly general cases: hyperexponential functions [3] (see also [15]), hypergeometric terms [9, 20], mixed hypergeometric-hyperexponential terms [5], algebraic functions [11], multivariate rational functions [6], and Fuchsian differential equations [8].

The existence of a reduction-based algorithm for general differential equations was raised as open Problem 2.2 in [10]. Problem 1.1 is essentially a more precise form of this problem, by specifying the space on which the reduction acts. From a cohomological point of view, reductions can be regarded as an explicit way to exhibit elements in cokernels. An abstract proof of the fact that the cokernel of the derivation on is finite-dimensional was given in [21]. Our solution to Problem 1.1 in particular yields a new constructive proof of this fact.

After section 2, we will leave applications to the theory of creative telescoping aside and focus on the construction of confined reductions. This construction proceeds in two stages. In section 3, we first consider the -submodule of of linear combinations with . We will construct a -linear head reduction such that and is bounded from above for all . Here we understand that for all . The head reduction procedure relies on the computation of a head chopper using an algorithm that will be detailed in section 5. We also need a variant of row echelon forms that will be described in section 4.

The head reduction may also be regarded as a way to reduce the valuation of in , at the point at infinity. In section 6 we turn to tail reductions, with the aim to reduce the valuation of at all other points in and its algebraic closure . This is essentially similar to head reduction via a change of variables, while allowing ourselves to work in algebraic extensions of . In the last section 7, we show how to combine the head reduction and the tail reductions at each of the roots of into a global confined reduction on . Using straightforward linear algebra and suitable valuation bounds, one can further turn this reduction into a normal one, as will be shown in section 7.3.

Our solution to Problem 1.1 is made precise in Theorems 5.9, 3.3, and 7.1. As far as we aware of, these results are new, and provide a positive answer to [10, Problem 2.2]. The application to creative telescoping is well known; see for instance [14, section 1.2.1]. Some of the techniques that we use are similar to existing ones. First of all, the construction of head choppers bears some similarities with Abramov's EG-eliminations [1]. Our procedure for head reduction is reminiscent of Euclidean division and classical algorithms for computing formal power series solutions to differential equations: first find the leading term and then continue with the remaining terms. In [11, section 5], a similar “polynomial reduction” procedure has been described in the particular case when . Finally, the idea to glue “local reductions” together into a global one is also common in this area [3, 5, 8].

Subsequently to the publication of a preprint version of this paper [18], the results have been further sharpened and generalized. In [4], an analogue algorithm was proposed for the case of higher order linear differential equations instead of first order matrix equations. This paper is mostly based on similar techniques, but also introduced a new tool: the Lagrange identity. In the terminology of the present paper, this makes it possible to avoid introducing the formal parameter , after which the operator from section 5 simply becomes multiplication with . Such simplifications make it easier to extend the theory beyond the setting of differential equations (1.1): see [19] for generalizations to difference equations. The original preprint version of this paper [18] also contained degree and valuation bounds for head and tail choppers; one of our motivations was to use these to derive polynomial complexity bounds for creative telescoping. Using the Lagrange identity technique from [4], it is possible to prove even sharper bounds. We refer to the follow-up paper [19] for more information on degree and valuation bounds and how to exploit them for proving polynomial complexity bounds.

2.Creative telescoping

2.1.Holonomic functions

Let be a subfield of . An analytic function on some non-empty open subset of is said to be holonomic (or D-finite) over if it satisfies a linear differential equation

(2.1)

where are rational functions and . Modulo multiplication by the common denominator, we may assume without loss of generality that are actually polynomials. Many, if not most, special functions are holonomic. Examples include , , , , Bessel functions, hypergeometric functions, polylogarithms, etc.

Instead of higher order scalar equations such as (2.1), it is also possible to consider first order linear differential systems

(2.2)

where is a non-zero polynomial and an matrix with polynomial coefficients. Given a column vector of analytic solutions to (2.2) on some non-empty open subset of , it is well-known that each component is a holonomic function. Conversely, taking and

any solution to (2.1) corresponds to a unique solution of (2.1).

The concept of holonomy extends to multivariate functions. There are again two equivalent formalizations that are respectively based on higher order scalar equations and first order systems. Let us focus on the bivariate case and let and denote the partial derivatives with respect to and . Consider a system of linear differential equations

(2.3)

where is non-zero and are such that

(2.4)

A holonomic function in two variables is defined to be a component of a solution to such a system. The compatibility relation (2.4) corresponds to the requirement , under the assumption that satisfies (2.3).

Example 2.1. The vector function

satisfies the system (2.3) for and

2.2.Creative telescoping

Assume that is an effective subfield of and let be a complex analytic solution of (2.3). By Cauchy-Kowalevski's theorem such solutions exist and can be continued analytically above . With , let be the -module generated by the entries of . Notice that is stable under both and . For any with and any non-singular contour in between two points , we may consider the integral

(2.5)

which defines a function in the single variable . It is natural to ask under which conditions is a holonomic function and how to compute a differential operator with .

The idea of creative telescoping is to compute a differential operator (called the telescoper), an element (called the certificate), and , such that

(2.6)

Integrating over , we then obtain

If the contour has the property that for all (where the equality is allowed to hold at the limit if necessary), then yields the desired annihilator with . In general, we need to multiply on the left with an annihilator of , as operators in the skew ring .

Example 2.2. With as in Example 2.1, we have . The contour that follows the real axis from to is non-singular and any function in vanishes at the limits of this contour (for fixed ). In particular, taking , the integral

is well defined for all . It can be checked that

(2.7)

whence we may take as our telescoper and as our certificate. Integrating over , it follows that

This equation admits a simple closed form solution

for some integration constant . In general, the computation of such integration constants is a difficult problem that is well beyond the scope of this paper. For our particular example, we have

whence . We could have seen this more directly by observing that the integrand is an odd function in for all . On the other hand, a similar computation for and

leads to

(2.8)

and

2.3.Reduction-based creative telescoping

We have shown how relations of the form (2.6) can be used for the computation of parametric integrals (2.5). This leaves us with the question how to find such relations. Many different approaches have been proposed for this task and we refer to [12] for a historical overview. From now on we will focus on the reduction-based approach, which is fairly recent and has shown to be particularly efficient for various subclasses of holonomic functions.

Notice that the first equation of the system (2.3) is of the form (1.1), where we recall that . Now assume that we have a computable confined reduction . Then the functions in the sequence can all be computed and they belong to a finite dimensional -vector space . Using linear algebra, this means that we can compute a relation

(2.9)

with and . Taking

we thus obtain (2.6). If the relation (2.9) has minimal order and the reduction is normal, then it can be shown [14, Proposition 16] that there exist no relations of the form (2.6) of order lower than .

One important property of reduction-based telescoping is that it allows us to compute telescopers without necessarily computing the corresponding certificates. In practice, it turns out that certificates are often much larger than telescopers; this often explains the efficiency of the reduction-based approach. Notice that the above approach can easily be adapted to compute certificates as well, when really needed: it suffices to require that the reduction procedure also produces a with . Given a relation (2.9) and with , we indeed have .

Example 2.3. Continuing Examples 2.1 and 2.2, let us show how to compute a confined reduction . Given , our algorithm to compute proceeds by induction on . If , then we take . Otherwise, we may write and . Setting

we have

whence

(2.10)

is of the form with . By the induction hypothesis, we know how to compute , so we can simply take . It is easily verified that , again by induction on , so the reduction is confined.

Applying our reduction to the functions and from Example 2.2, we find that

and

This leads to the desired relations (2.7) and (2.8).

Remark 2.4. In order to simplify the exposition, we have restricted our attention to the bivariate case. Nevertheless, the reduction-based approach extends to the case when is replaced by a finite number of coordinates and satisfies an equation with respect to each coordinate (with suitable compatibility constraints). Indeed, for each , it suffices to compute the sequence until we find a relation with . For each , this yields a non-trivial operator with .

3.Head reduction

3.1.Head choppers

Let , where and are indeterminates. We view as a parameter that takes integer values. We may regard as a Laurent polynomial with matrix coefficients :

(3.1)

If , then we denote . Setting

(3.2)

the equation (1.1) implies

(3.3)

for any constant matrix . The matrix can also be regarded as a Laurent polynomial with matrix coefficients . We say that is a head chopper for (1.1) if is an invertible matrix.

Example 3.1. With and as in Example 2.1, the identity matrix is a head chopper. Indeed, for this choice of , we obtain

(3.4)

so and is invertible. The matrix is also a head chopper, with

(3.5)

Example 3.2. Consider the equation (1.1) for and

for some formal parameter . Then we claim that

(3.6)

is a head chopper. Indeed, a straightforward computation yields

(3.7)

which shows that the leading coefficient of as a polynomial in is (formally) invertible.

3.2.Head reduction

Before studying the computation of head choppers, let us first show how they can be used for the construction of so-called “head reductions”, by generalizing the inductive construction from Example 2.3. Let be a head chopper for (1.1) and assume in addition that and . Given -subvector spaces and of , we say that a -linear map is a partial reduction for (1.1) if for all .

Let . Writing with and , we say that is an exceptional index if or . Here we understand that stands for the evaluation of at and similarly for . We write for the finite set of exceptional indices. If , then we notice that the matrix is invertible.

Any can be regarded as a polynomial in . Given , let

If and , then recall that the matrix is invertible. We may thus define the -linear mapping by

We indeed have , since

The mapping also induces a mapping that we will still denote by . Setting , the relation (3.3) yields

This shows that the mapping is a partial reduction. If and , then we have and the identity map is clearly a partial reduction as well.

Now we observe that compositions of partial reductions are again partial reductions. For each , we thus have a partial reduction

Now let be the unique mapping with for all and . Then is clearly a partial reduction as well and it admits a finite dimensional image . For any , we call the head reduction of . The following straightforward algorithm allows us to compute head reductions:

Algorithm HeadReduce

Input:

Output: the head-reduction of

repeat

if for all then return

Let be maximal with

Theorem 3.3.

The routine terminates and is correct.

Example 3.4. Let and be as in Examples 2.1, 2.2, 2.3 and 3.1. Taking the head chopper with as in (3.5), we get

for all and . In other words, coincides with from (2.10), so the head reduction procedure coincides with the reduction procedure from Example 2.3, except that we directly reduce any with to itself (we have and ). The fact that we may actually reduce elements with somewhat further is due to the fact that the coefficient of in (3.4) vanishes for . Indeed, this means that the matrix from (3.4) actually evaluates to a polynomial in at , so we may use it instead of the matrix from (3.5) as a head chopper.

Example 3.5. Let and be as in Example 3.2, with . Taking and as in (3.6) and (3.7), we obtain , , and

For , we note that

Let us show how and can be used to compute the head-chopper of , where

Applying to , we find that is maximal with , so we set

We repeat the loop for the new value of , which yields

Repeating the loop once more, we obtain

At this point, we have , so we have obtained the head reduction of the original .

Example 3.6. Let , , , and be as in Example 3.2 and . Taking

we observe that , whence any element of is head-reduced. Even for equations of bounded degree and order, this means that head reduced elements can have arbitrarily large degree.

Remark 3.7. It is straightforward to adapt HeadReduce so that it also returns the certificate with . Indeed, it suffices to start with and accumulate at the end of the main loop.

Remark 3.8. In the algorithm HeadReduce we never used the assumption that has one row. In fact, the same algorithm works for matrices with an arbitrary number of rows . This allows for the simultaneous head reduction of several elements in , something that might be interesting for the application to creative telescoping.

4.Row swept forms

The computation of head choppers essentially boils down to linear algebra. We will rely on the concept of “row swept forms”. This notion is very similar to the more traditional row echelon forms, but there are a few differences that are illustrated in Example 4.1 below.

Let be a matrix and denote the -th row of by . Assuming that , its leading index is the smallest index with . We say that is in row swept form if there exists a such that and for all . Notice that has rank in this case.

An invertible matrix such that is in row swept form will be called a row sweeper for . We may compute such a matrix using the routine RowSweeper below, which is really a variant of Gaussian elimination. Whenever we apply this routine to a matrix such that the truncated matrix with rows is in row swept form, we notice that these first rows are left invariant by the row sweeping process. In other words, the returned row sweeper is of the form . If, in addition, the matrix has rank , then is of the form .

Algorithm RowSweeper

Input: a matrix

Output: a row sweeper for

,

for from to do

if for all and then return

Let be minimal such that for some

Swap the -th and -th rows of and

for from to do

,

return

Example 4.1. Given the matrix

the algorithm RowSweeper produces the row sweeper

Since the two first rows of were already in row swept form, this matrix is indeed of the form . The resulting row swept form for is

The more traditional row echelon and reduced row echelon forms insist on moving the rows for which is minimal to the top, so the first two rows are not left invariant. The different “normal” forms that we obtain for our example matrix are shown below:

5.Computing head choppers

5.1.Transforming head choppers

In Example 3.1 we already pointed out that head choppers are generally not unique. Let us now study some transformations that allow us to produce new head choppers from known ones; this will provide us with useful insights for the general construction of head choppers. For any , we define the operator on by

Proposition 5.1. For all , we have

Proof. Setting , and , we have

In other words, .

Proposition 5.2. Assume that and that is invertible. Then

  1. is a head chopper for (1.1) if and only if is a head chopper for (1.1).

  2. is a head chopper for (1.1) if and only if is a head chopper for (1.1).

Proof. Assume that is a head chopper for (1.1). Setting and , we have and is invertible. Similarly, setting and , we have , whence is invertible. The opposite directions follow by taking and in the roles of and .

Given a head chopper for (1.1) and , let be minimal such that or . Then is also maximal with the property that and . From Proposition 5.2(a) it follows that is again a head chopper for (1.1). Without loss of generality, this allows us to make the additional assumption that at the beginning of subsection 3.2.

5.2.Head annihilators

In order to compute head choppers by induction, it will be convenient to introduce a partial variant of this concept. First of all, we notice that the equations (3.13.3) and Proposition 5.1 generalize to the case when , where is arbitrary. Notice also that , where . Given and , let

It is easy to see that both and are -modules.

Now consider a matrix with rows ordered by increasing degree

Let , let be the matrix with rows , and let be maximal such that . We say that is a -head annihilator for (1.1) if the following conditions are satisfied:

HA1

The rows of form a basis for the -module ;

HA2

The matrix is invertible;

HA3

The first rows of are -linearly independent.

The matrix is obviously a -head annihilator with . If , then we notice that HA3 implies that is a head chopper for (1.1). We also have the following variant of Proposition 5.2(a):

Proposition 5.3. For any , we have

Moreover, is a -head annihilator if and only if is a -head annihilator.

Using a constant linear transforation as in Proposition 5.2(b), we may now achieve the following:

Proposition 5.4. Let be a -head annihilator for (1.1). Let and be as in HA1HA3 and denote . Then there exists an invertible matrix of the form

such that the last rows of vanish and such that is a -head annihilator for (1.1).

Proof. Let

be the row sweeper for as computed by the algorithm RowSweeper from section 4. By construction, for all , and the last rows of vanish. We claim that for all . Indeed, if , then this would imply that , which contradicts HA2. From our claim, it follows that and is maximal with the property that . Since the first rows of and coincide, the first rows of are -linearly independent. This shows that HA3 is satisfied for . As to HA2, let be the invertible matrix with

Then we notice that , whence is invertible. The rows of clearly form a basis for , since is invertible.

As long as is not invertible, we finally use the following simple but non-constant linear transformation in order to improve the rank of :

Proposition 5.5. Let be a -head annihilator for (1.1). Let , let , and assume that the last rows of vanish. Let be the matrix with rows

Then is a -head annihilator for (1.1).

Proof. We have for all and for all . In particular, we have and is maximal with the property that . Setting , we also observe that for all . Since and the last rows of vanish, the first rows of both and are -linearly independent. In other words, HA3 is satisfied for . As to HA2, we observe that , whence is invertible.

Let us finally show that forms a basis for the -module . So let . Then , so for some row matrix . Setting , we have , whence . Since the first rows of are -linearly independent and the last rows of vanish, we get for all . Let be the row vector with for and for . By what precedes, we have and . Now we have for and for . In other words, , as desired.

5.3.Computing head choppers

Propositions 5.4 and 5.5 allow us to compute -head annihilators for (1.1) with arbitrarily large . Assuming that we have in HA3 for sufficiently large , this yields the following algorithm for the computation of a head chopper for (1.1):

Algorithm HeadChopper

Input: and

Ouput: a head chopper for (1.1)

repeat

if is invertible then return

,

Example 5.6. Before we prove the correctness of HeadChopper, let us show how it works for and as in Example 3.2. We enter the loop with

so that is a -head annihilator for (1.1). During the first iteration of the loop, we set

and then

Propositions 5.4 and 5.5 imply that the new matrix is a -head annihilator for (1.1). The second iteration of the main loop yields

after which we set ,

At this point, is a -head annihilator for (1.1). The third iteration of the main loop yields

and then , ,

Applying to both and , we find the head chopper from Example 3.2.

5.4.Correctness and termination

Proposition 5.7. Let . Consider the value of at the beginning of the loop and after iterations. Then is a -head annihilator.

Proof. We first observe that throughout the algorithm. Let us now prove the proposition by induction over . The proposition clearly holds for . Assuming that the proposition holds for a given , let us show that it again holds at the next iteration. Consider the values of and at the beginning of the loop and after iterations. Let be maximal such that . From the induction hypothesis, it follows that the first rows of are -linearly independent, whence the matrix is of the form

Now Proposition 5.4 implies that is still a -head annihilator. Since the last rows of vanish, Proposition 5.5 also implies that is a -head annihilator. This completes the induction. Notice also that is maximal with the property that .

Proposition 5.8. Assume (for contradiction in Theorem 5.9 below) that the algorithm HeadChopper does not terminate for some given input . Then there exists a non-zero row matrix with . In particular, .

Proof. Assume that HeadChopper does not terminate. Let be the value of at the beginning of the main loop after iterations. Also let and be the values of and as computed during the -th iteration.

Let be maximal such that . Using the observation made at the end of the above proof, we have , so there exist an index and with for all . Furthermore,

and

Moreover, for , the row sweeper is even of the form

By induction on , we observe that . For , we also have for all , again by induction. Consequently, for all , which means that the sequence formally converges to a limit in . By construction, the first rows of are zero, its last rows have rank , and . We conclude by taking to be the last row of .

Theorem 5.9. The algorithm HeadChopper terminates and returns a head chopper for (1.1).

Proof. We already observed that throughout the algorithm. If the algorithm terminates, then it follows that is indeed a head chopper for (1.1). Assume for contradiction that the algorithm does not terminate and let be such that . Let be a fundamental system of solutions to the equation (1.1), where is some differential field extension of with constant field . From we deduce that , whence . More generally, whence and for all . Since the space has dimension over , it follows that there exists a polynomial of degree at most in such that and . Since is a fundamental system of solutions, we have . This contradicts the existence of an element with .

Remark 5.10. In [18, section 6], we proved a polynomial degree bound for the computed head chopper. Sharper bounds have been proven in [4, 19] and the complexity of reduction-based creative telescoping has been further investigated in [19].

Note that one should not confuse the existence of polynomial degree bounds for head choppers with the absence of such bounds for exceptional indices. Indeed, Example 3.6 shows how to obtain arbitrarily high exceptional indices for equations of bounded degree and order. Yet, the degrees of the corresponding head choppers are also bounded, as shown in Example 3.2.

Remark 5.11. As stated in the introduction, the construction of head choppers bears some similarities with Abramov's EG-eliminations [1]. Let be an indeterminate and let be the shift operator. Then EG-eliminations can be used to compute normal forms for linear difference operators in . The rank of the leading (or trailing coefficient) of the normal form is equal to the rank of the original operator. Abramov achieves such normal form computations by transforming the problem into a big linear algebra problem over . Our algorithm for the computation of head choppers is different in two ways: the operator is not a shift operator and we work directly over .

6.Tail reduction

Head reduction essentially allows us to reduce the valuation in of elements in via the subtraction of elements in . Tail reduction aims at reducing the valuation in in a similar way for any in the algebraic closure of . It is well known that holonomy is preserved under compositions with rational functions. Modulo suitable changes of variables, this allows us to compute tail reductions using the same algorithms as in the case of head reduction. For completeness, we will detail in this section how this works.

6.1.Tail choppers

More precisely, let , and . We may regard as a Laurent polynomial in with matrix coefficients :

(6.1)

If , then we denote its valuation in by . Setting

(6.2)

the equation (1.1) implies

(6.3)

for any matrix . The matrix can also be regarded as a Laurent polynomial with matrix coefficients . We say that is a tail chopper at for (1.1) if is an invertible matrix. In fact, it suffices to consider tail choppers at the origin:

Lemma 6.1. Let , where . Define , and . Then is a tail chopper at for (1.1) if and only if is a tail chopper at for .

Proof. Setting , we have . Consequently, and .

There is also a direct link between head choppers and tail choppers at via the change of variables .

Lemma 6.2. Let . Setting , we define , and . Then is a tail chopper at for (1.1) if and only if is a head chopper for .

Proof. Setting , we have

Consequently, and .

Finally, the matrix is a tail chopper at almost all points :

Lemma 6.3. Let be such that . Then is a tail chopper for (1.1) at .

Proof. If and , then (6.2) becomes for . In particular, and is invertible in .

6.2.Computing tail choppers

Now consider a monic square-free polynomial and assume that we wish to compute a tail chopper for (1.1) at a root of in . First of all, we have to decide how to conduct computations in . If is irreducible, then we may simply work in the field instead of and take to be the residue class of , so that becomes a generic formal root of . In general, factoring over may be hard, so we cannot assume to be irreducible. Instead, we rely on the well known technique of dynamic evaluation [13].

For convenience of the reader, let us recall that dynamic evaluation amounts to performing all computations as if were irreducible and were a field with an algorithm for division. Whenever we wish to divide by a non-zero element (with ) that is not invertible, then provides us with a non-trivial factor of . In that case, we launch an exception and redo all computations with or in the role of .

So let be a formal root of and define , , and . Let be a head chopper for the equation , as computed using the algorithm from section 5.3. Then is a tail chopper at by Lemmas 6.1 and 6.2.

6.3.Tail reduction

Let be a tail chopper for (1.1) at . Let , , and be as above, so that , is a head chopper for the equation . In particular, rewriting linear combinations with as linear combinations with , we may head reduce as described in section 3.2. Let be such that . Then we may rewrite as an element of . We call the tail reduction of at and write .

Let be the finite set of exceptional indices for the above head reduction and . Setting and , it can be checked that the following algorithm computes the tail reduction at :

Algorithm TailReduce

Input:

Output: the tail reduction of at

repeat

if for all then return

Let be minimal with

7.Global reduction

7.1.Gluing together the head and tail reductions

Let us now study how head and tail reductions can be glued together into a global confined reduction on . More generally, we consider the case when , where is a monic square-free polynomial such that divides for some .

We assume that we have computed a head chopper for (1.1) and tail choppers for (1.1) at each the roots of in . In particular, we may compute the corresponding head and tail reductions

Given an element of the Galois group of over , we may also assume without loss of generality that the tail choppers were chosen such that for all (note that this is automatically the case when using the technique of dynamic evaluation from section 6.2).

Partial fraction decomposition yields -linear mappings

and

with

for all . This allows us to define a global reduction of by

(7.1)

Theorem 7.1. The formula (7.1) defines a confined reduction on .

Proof. Let be an automorphism of over . Then naturally extends to by setting for all . Given and , we have . By our assumption that , it follows that

Summing over all , we get . Since this equality holds for all automorphisms , we conclude that . Similarly, given with , we have for all automorphisms , whence . This shows that (7.1) defines a reduction on . For any in the image of the restriction of to , we finally observe that , and are uniformly bounded, by construction. In other words, the reduction is confined.

7.2.Machine computations

For actual implementations, one may perform the computations in extension fields , where is an irreducible factor of (or simply a square-free factor, while relying on dynamic evaluation as in section 6.2). Let be the roots of such an irreducible factor and assume that we wish to compute for . Instead of computing each separately, one may use the formula

where is the canonical root of in and for all .

Example 7.2. Let be a fixed parameter. Consider the function

which satisfies the differential equation

This equation admits two singularities at , where . Any non-zero element of is a tail chopper at . Taking as our tail chopper, we have

for all and . Given

its tail reduction at is therefore recursively defined by

Now assume that we wish to compute the tail reduction

of

with respect to both roots and of . We have

The above computation holds when considering as a root of the polynomial in the algebraic closure of . Exactly the same computation can therefore be used for the other root of this polynomial. The computation also holds for a generic root in the algebraic extension and we obtain

7.3.Normalizing the reduction

Given the confined reduction from section 7.1, let us now give a general procedure how to turn it into a normal confined reduction . For this purpose, we assume that we know a -subvector space of with the property that for any with , we have .

Remark 7.3. It can be shown that there exist integers , and such that we can take

For equations (1.1) of bounded degree and size, Example 3.6 shows that can become arbitrarily large, whence so can the dimension of . For this reason, normal confined reductions can be computationally expensive, so it is usually preferable to rely on non-normalized reductions. One way to compute , and was detailed in [18, sections 6 and 7], but better approaches have been proposed since [4, 19].

Now let and let be a supplement of in so that . We may compute bases of and using straightforward linear algebra. The canonical -linear projections and with are also computable. We claim that we may take for every .

Proposition 7.4. The mapping defines a computable normal confined reduction on .

Proof. The mapping is clearly a computable confined reduction on . It remains to be shown that for all . Now , so and there exists a with . Since , it follows that and . In other words, and .

Acknowledgments. We would like to thank Pierre Lairez for a helpful remark. We also thank the second referee for pointing us to [21] and for further helpful remarks and suggestions.

Bibliography

[1]

S. A. Abramov. EG-eliminations. Journal of Difference Equations and Applications, 5(4–5):393–433, 1999.

[2]

A. Bostan, S. Chen, F. Chyzak, and Z. Li. Complexity of creative telescoping for bivariate rational functions. In Proc. ISSAC '10, pages 203–210. New York, NY, USA, 2010. ACM.

[3]

A. Bostan, S. Chen, F. Chyzak, Z. Li, and G. Xin. Hermite reduction and creative telescoping for hyperexponential functions. In Proc. ISSAC'13, pages 77–84. ACM, 2013.

[4]

A. Bostan, F. Chyzak, P. Lairez, and B. Salvy. Generalized Hermite reduction, creative telescoping and definite integration of differentially finite functions. In Proc. ISSAC '18, pages 95–102. New York, 2018. ACM.

[5]

A. Bostan, L. Dumont, and B. Salvy. Efficient algorithms for mixed creative telescoping. In Proc. ISSAC'16, pages 127–134. ACM, 2016.

[6]

A. Bostan, P. Lairez, and B. Salvy. Creative telescoping for rational functions using the Griffiths-Dwork method. In Proc. ISSAC'13, pages 93–100. ACM, 2013.

[7]

S. Chen. Some applications of differential-difference algebra to creative telescoping. PhD thesis, École Polytechnique, 2011.

[8]

S. Chen, M. van Hoeij, M. Kauers, and C. Koutschan. Reduction-based creative telescoping for Fuchsian D-finite functions. JSC, 85:108–127, 2018.

[9]

S. Chen, H. Huang, M. Kauers, and Z. Li. A modified Abramov-Petkovsˇek reduction and creative telescoping for hypergeometric terms. In Proc. ISSAC'15, pages 117–124. ACM, 2015.

[10]

S. Chen and M. Kauers. Some open problems related to creative telescoping. J. of Systems Science and Complexity, 30(1):154–172, 2017.

[11]

S. Chen, M. Kauers, and C. Koutschan. Reduction-based creative telescoping for algebraic functions. In Proc. ISSAC '16, pages 175–182. New York, NY, USA, 2016. ACM.

[12]

F. Chyzak. The ABC of Creative Telescoping — Algorithms, Bounds, Complexity. Habilitation, École polytechnique, 2014.

[13]

J. Della Dora, C. Dicrescenzo, and D. Duval. A new method for computing in algebraic number fields. In G. Goos and J. Hartmanis, editors, Eurocal'85 (2), volume 174 of Lect. Notes in Comp. Science, pages 321–326. Springer, 1985.

[14]

L. Dumont. Efficient algorithms for the symbolic computation of some contour integrals depending on one parameter. PhD thesis, École Polytechnique, 2016.

[15]

K. Geddes, H. Le, and Z. Li. Differential rational normal forms and a reduction algorithm for hyperexponential functions. In Proc. ISSAC'04, pages 183–190. ACM, 2004.

[16]

C. Hermite. Sur l'intégration des fractions rationnelles. Ann. Sci. École Norm. Sup. Série 2, 1:215–218, 1972.

[17]

J. van der Hoeven. Relax, but don't be too lazy. JSC, 34:479–542, 2002.

[18]

J. van der Hoeven. Constructing reductions for creative telescoping. Technical Report, HAL, 2017. http://hal.archives-ouvertes.fr/hal-01435877.

[19]

J. van der Hoeven. Creative telescoping using reductions. Technical Report, HAL, 2018. http://hal.archives-ouvertes.fr/hal-01773137.

[20]

H. Huang. New bounds for hypergeometric creative telescoping. In Proc. ISSAC'16, pages 279–286. ACM, 2016.

[21]

P. Monsky. Finiteness of de Rham cohomology. American Journal of Mathematics, 94(1):237–245, 1972.

[22]

M. Ostrogradsky. De l'integration des fractions rationelles. Bull. de la Classe Physico- Mathématique de l'Académie Imperiale des Sciences de Saint-Petersburg, IV:147–168, 1845.

[23]

D. Zeilberger. The method of creative telescoping. JSC, 11(3):195–204, 1991.