Fast Gröbner basis computation and
polynomial reduction for generic bivariate ideals

Joris van der Hoevena, Robin Larrieub

Laboratoire d'informatique de l'École polytechnique

LIX, UMR 7161 CNRS

Campus de l'École polytechnique

1, rue Honoré d'Estienne d'Orves

Bâtiment Alan Turing, CS35003

91120 Palaiseau, France

a. Email: vdhoeven@lix.polytechnique.fr

b. Email: larrieu@lix.polytechnique.fr

Final version of December 2019

Let be two bivariate polynomials over an effective field , and let be the reduced Gröbner basis of the ideal generated by and with respect to the usual degree lexicographic order. Assuming and sufficiently generic, we design a quasi-optimal algorithm for the reduction of modulo , where “quasi-optimal” is meant in terms of the size of the input . Immediate applications are an ideal membership test and a multiplication algorithm for the quotient algebra , both in quasi-linear time. Moreover, we show that itself can be computed in quasi-linear time with respect to the output size.

Keywords:

polynomial reduction; Gröbner basis; complexity; algorithm

A.M.S. subject classification: 13P10

1.Introduction

Gröbner bases, also known as standard bases, are a powerful tool for solving systems of polynomial equations, or to compute modulo polynomial ideals. The research area dedicated to their computation is very active, and there is an abundant literature on efficient algorithms for this task. See for example [5, 6, 8] and references therein. Although this problem requires exponential space in the worst case [21], it is in fact tractable for many practical instances. For example, computer algebra systems often implement Faugère's F5 algorithm [6] that is very efficient if the system has sufficient regularity. In this case, a polynomial complexity bound (counting the number of field operations in terms of the expected output size) was established in [1].

The F5 algorithm and all other currently known fast algorithms for Gröbner basis computations rely on linear algebra, and it may seem surprising that fast FFT-based polynomial arithmetic is not used in this area. This can be seen as an illustration of how difficult it is to compute standard bases, but there is another explanation: traditionnally, Gröbner basis algorithms consider a large number of variables and the degree of the generating polynomials is kept small; but fast polynomial arithmetic works best in the opposite regime (a fixed number of variables and large degrees). Even in this setting, it is not clear how to use FFT techniques for Gröbner basis computation. As a first step, one may consider related problems, such as the reduction of multivariate polynomials. It was shown in [16] that reduction can be done in quasi-linear time with respect to the size of the equation . However, this equation is in general much larger than the intrinsic complexity of the problem, given by the size of and the degree of the ideal (which is linked to the size of the generating polynomials). Recent work in the bivariate setting [17] gave an asymptotically optimal reduction algorithm for a particular class of Gröbner bases. This algorithm relies on a terse representation of in space, where stands for the “soft Oh” notation (that hides poly-logarithmic factors) [11]. Assuming that this representation has been precomputed, the extended reduction can be performed in time , instead of the previous where .

Instead of making regularity assumptions on the Gröbner basis itself, one may focus on the generating polynomials. If the ideal is defined by generic polynomials given in total degree, then the Gröbner basis presents a particular structure, as studied for example in [10, 22]. This situation is often used as a benchmark for polynomial system solving: see the PoSSo problem [7]. In this paper, we restrict ourselves to the bivariate case, as studied for example in [20]. In what follows, are generic polynomials of degree respectively. We denote by the ideal they generate, and we consider its Gröbner basis with respect to the graded lexicographic order. The computation of such a basis is classical, but the hypotheses from [17] are not satisfied. However, we show in this paper that a similar terse representation does exist in this case. Therefore, reduction in quasi-linear time remains possible, and it even turns out that the suitable representation can be computed in time from the input . Combining these two algorithms, we obtain an ideal membership test in quasi-linear time , without any precomputation. Similarly, there is a quasi-linear multiplication algorithm for (again without precomputation). Finally, we show that the reduced Gröbner basis can be computed in quasi-linear time with respect to the output size.

Notations and terminology. We assume that the reader is familiar with the theory of Gröbner basis and refer to [11, 2] for basic expositions. We denote the set of monomials in variables by . A monomial ordering on is a total ordering that is compatible with multiplication. Given a polynomial in variables , its support is the set of monomials with . If , then admits a maximal element for that is called its leading monomial and that we denote by . If , then we say that is a term in . Given a tuple of polynomials in , we say that is reduced with respect to if contains no monomial that is a multiple of the leading monomial of one of the .

Unless stated otherwise, we will always work in the bivariate setting when , and use and as our main indeterminates instead of and . In particular, . Moreover, we only consider the usual degree lexicographic order with , that is

Acknowledgements. We thank Vincent Neiger for his remark that simplified Algorithm 6.

2.Presentation of the setting

We consider an ideal generated by two generic polynomials of total degree and we assume . Here the adjective “generic” should be understood as “no accidental cancellation occurs during the computation”. This is typically the case if are chosen at random: assuming has sufficiently many elements, the probability of an accidental cancellation is small. In this generic bivariate setting, it is well known that the reduced Gröbner basis of with respect to can essentially be computed as follows:

Remark 1. For simplicity we neglected technical details to actually have the reduced Gröbner basis. First, one should set , to ensure that no term in is divisible by the leading term of (which may happen if ). Moreover, elements of the reduced basis must be monic; this means a division by the leading coefficient is missing in the given formulas.

We will show next how to construct another (non-reduced) Gröbner basis with even simpler recurrence relations. Based on these recurrence relations, we will design a variant of the algorithm from [17] that computes the reduction modulo in quasi-linear time.

Definition 2. For a polynomial of total degree , we define its dominant diagonal by .

We have the trivial properties that and . For generic and , the diagonals and are also generic. Then the remainders during the Euclidean algorithm follow a normal sequence (that is the degrees decrease by exactly 1 at each step).

Remark 3. When we say “ are generic”, we actually make two assumptions. The first one is that the remainders follow a normal sequence as we just mentioned. The second one is simply that the coefficient of in is nonzero, so that is the leading monomial of .

Let us consider the sequence defined as:

Let us first notice that the term corresponds to the S-polynomial of and , as in the classical Buchberger algorithm [3]. We observe next that , and for , , so that the diagonals are the successive remainders in the Euclidean algorithm (hence the quotients of degree 1). From these two facts, we deduce that the have the same leading monomials as the , so is a Gröbner basis of with respect to . Of course, any has the same normal form with respect to and .

With this construction, it is easy to deduce higher order recurrence relations for each (with ). As we will see in section 4, the polynomials and have a size that can be controlled as a function of . More precisely, they are homogeneous polynomials of degree roughly .

3.Algorithmic prerequisites

In this section, we quickly review some basic complexities for fundamental operations on polynomials over a field . Notice that results presented in this section are not specific to the bivariate case. Running times will always be measured in terms of the required number of field operations in .

3.1.Polynomial multiplication

We denote by the cost of multiplying two dense univariate polynomials of degree in . Over general fields, one may take [24, 23, 4]

In the case of fields of positive characteristic, one may even take , where denotes the iterated logarithm [13, 14]. We make the customary assumptions that is increasing and that , with the usual implications, such as .

For multivariate polynomials, the cost of multiplication depends on the geometry of the support. The classical example involves dense “block” polynomials in of degree in each variable . This case can be reduced to multiplication of univariate polynomials of degree using the well known technique of Kronecker substitution [11]. More generally, for polynomials such that the support of the product is included in an initial segment with elements, it is possible to compute the product in time [18]. Here an initial segment of is a (finite) subset such that for any monomial , all its divisors are again in .

For the purpose of this paper, we need to consider dense polynomials in whose supports are contained in sets of the form . Modulo the change of variables , such a polynomial can be rewritten as , where the support of is an initial segment with the same size as . For a product of two polynomials of this type with a support of size , this means that the product can again be computed in time .

3.2.Relaxed multiplication

For the above polynomial multiplication algorithms, we assume that the input polynomials are entirely given from the outset. In specific settings, the input polynomials may be only partially known at some point, and it can be interesting to anticipate the computation of the partial output. This is particularly true when working with (truncated) formal power series instead of polynomials, where it is common that the coefficients are given as a stream.

In this so-called “relaxed (or online) computation model”, the coefficient of a product of two series must be output as soon as and are known. This model has the advantage that subsequent coefficients and are allowed to depend on the result . This often allows us to solve equations involving power series by rewriting them into recursive equations of the form , with the property that the coefficient only depends on earlier coefficients for all . For instance, in order to invert a power series of the form with , we may take . Similarly, if has characteristic zero, then the exponential of a power series with can be computed by taking .

From a complexity point of view, let denote the cost of the relaxed multiplication of two polynomials of degree . The relaxed model prevents us from directly using fast “zealous” multiplication algorithms from the previous section that are typically based on FFT-multiplication. Fortunately, it was shown in [15, 9] that

(1)

This relaxed multiplication algorithm admits the advantage that it may use any zealous multiplication as a black box. Through the direct use of FFT-based techniques, the following bound has also been established in [19]:

In the sequel, we will only use a suitable multivariate generalization of the algorithm from [15, 9], so we will always assume that

In particular, we have .

3.3.Polynomial reduction

Let us now consider a Gröbner basis of an ideal in , or, more generally, an auto-reduced tuple of polynomials in . Then for any , we may compute a relation

such that is reduced with respect to . We call an extended reduction of with respect to .

The computation of such an extended reduction is a good example of a problem that can be solved efficiently using relaxed multiplication and recursive equations. For a multivariate polynomial with dense support of any of the types discussed in section 3.1 , let denote a bound for the size of its support. With as in ( 1 ), it has been shown

1. The results from [16] actually apply for more general types of supports, but this will not be needed in this paper.

1 in [ 16 ] that the quotients and the remainder can be computed in time

(2)

This implies in particular that the extended reduction can be computed in quasi-linear time in the size of the equation . However, as pointed out in the introduction, this equation is in general much larger than the input polynomial .

Extended reductions are far from being unique (only is unique, and only if is a Gröbner basis). The algorithm from [16] for the computation of an extended reduction relies on a selection strategy that uniquely determines the quotients. More precisely, for every monomial , we define the set ; then we need a rule to chose a particular index (assuming is non-empty). The initial formulation [16] used the simplest such strategy by taking , but the complexity bound (2) holds for any selection strategy. Now the total size of all quotients may be much larger than the size of for a general selection strategy. One of the key ingredients of the fast reduction algorithm in this paper is the careful design of a “dichotomic selection strategy” that enables us to control the degrees of the quotients.

Remark 4. The notion of selection strategy is somewhat similar to the concept of involutive division introduced for the theory of involutive bases [12], although our definition is more permissive.

4.Concise representation for Gröbner bases

The ideal has a degree , and the reduced Gröbner basis takes space . This overhead compared to can be reduced: it was shown in [17] that special bases called vanilla Gröbner bases admit a terse representation in space that allow for efficient reduction. Obtaining such a terse representation can be expensive, even if the Gröbner basis is known, but this can be seen as a precomputation. However, for the graded lexicographic order, neither nor are vanilla, so the results from [17] do not apply. Nevertheless, we show in this section that does admit a concise representation which is analogous to the concept of terse representation. This representation is compatible with a new reduction algorithm detailed in section 5, and it turns out that it can even be computed in time from the input .

4.1.Definition

As detailled in section 2, the basis is constructed such that there are recurrence relations of the form for each (with ). Equivalently, it is more convenient to write this in matrix notation

(3)

(where we set to avoid case distinction). This matrix notation has the advantage that the can be computed from one another using . Moreover, the size of the coefficients of the can be controlled as a function of .

Definition 5. For each , let be the successive quotients in the euclidean algorithm for the dominant diagonals and (as in section 2). For each with , define the matrix by

Proposition 6. Let the matrices be as in Definition 5. For all with , we then have

Also, for all with .

Now consider the polynomials such that

With the convention that the zero polynomial is homogeneous of any degree, we have

  • For all , is homogeneous of degree and is homogeneous of degree .

  • is homogeneous of degree and is homogeneous of degree .

  • For all , is homogeneous of degree and is homogeneous of degree .

Proof. This is an immediate induction on using .

As in [17], we use a dichotomic selection strategy to control the degrees of the quotients in the extended reduction, together with rewriting rules to evaluate the combination . Then it is sufficient to know enough head terms of each to compute the reduction. Finally, we observe that if and are known with sufficient precision, then and can be computed with the same precision using the relation (3) above. More formally, we define the precision as follows:

Definition 7. Given a polynomial , we define its upper truncation with precision as the polynomial such that

In other words, and have the same terms of degree at least , but all terms of of degree less than are zero.

With the dichotomic selection strategy that we plan to use, we have for , so it is a priori sufficient to compute with this precision. However, the rewriting rules between the involve two consecutive elements, so that and need to be known with the same precision.

Definition 8. The concise representation of consists of the following data:

  • The sequence of truncated elements , where

    • for , ;

    • for all other , , with the notation ;

  • The collection of rewriting matrices for each where the notation is defined as follows: with the matrices as in Definition 5, we set

    with and .

As expected, the concise representation requires quasi-linear space with respect to the degree of the ideal.

Proposition 9. The concise representation requires space .

Proof. It is easy to see that has degree at most in the variable , and at most in the variable and in total degree. Then for , each non-truncated element takes space. Similarly, for , each truncated element requires space. For each , there are roughly indices such that , so all elements together require space.

There are elements in , each consisting of four homogeneous polynomials of degree roughly (by Proposition 6), except for that has two polynomial entries of degree roughly and two entries of degree roughly . Hence requires space and the collection of all takes space.

4.2.Computation

The definition of the concise representation as above is constructive, but the order of the computation must be carefully chosen to achieve the expected quasi-linear complexity. First, by exploiting the recurrence relations , it is easy to compute from using the following auxilliary function:

Algorithm 1

Input: as in Definition 8.

Output: as in Definition 8.

Set .

For each :

Set .

If :

Set .

Return .

Lemma 10. Algorithm 1 is correct and takes time .

Proof. Recall that for all with . In particular, taking shows that if . Concerning the last element of (that is, if is even, otherwise), define

If , then is odd, that is and , so that is indeed the product of the last two elements of . Conversely if , then and so that and have the same last element.

The complexity bound is obtained with the same argument as for Proposition 9.

The algorithm to compute the concise representation can be decomposed in three steps. First a euclidean algorithm gives recurrence relations of order 1. A second elementary step is to deduce higher order relations by the above algorithm. Finally, one has to compute the truncated basis elements . Starting with those of highest precision avoids computing unnecessary terms, so that quasi-linear complexity can be achieved.

Algorithm 2

Input: , two generic bivariate polynomials of total degrees and with .

Output: , the concise representation of a Gröbner basis of with respect to .

Set and .

Set and .

For each :

Set and .
// Fail if the quotient has degree

If , then set , otherwise set .

Set .

Set and .

For each :

Compute from using Algorithm 1.

Compute .
// use

For each :

For each with , set
// Use

Return and .

Theorem 11. Algorithm 2 is correct and takes time

Proof. The reduction can be done in a relaxed way at a cost of . The first loop is clearly correct and each step requires operations. Alternatively, all the successive quotients can be computed with a fast Euclidean algorithm (see for example [11]) at a cost of operations.

In the last loop on , since the precision decreases at each step and since there is no accidental cancellation, the invariant

holds. Indeed, regarding upper truncations, it is clear that as soon as . Then we have (handle the case separately)

which proves the correctness. Let us now evaluate the complexity of this loop. For each index such that , we have to estimate the support of and (where ). For , and are completely known, with a support of size . In all other cases and are upper truncations, with a support of size . Consequently, each iteration of the loop requires operations.

5.Fast reduction algorithm

In this section, we present a variant of the algorithm from [17] that performs the reduction in quasi-linear time using the concise representation of . The general idea is the same: reduce modulo the truncated basis to compute the quotients faster, and to compute the remainder, use the recurrence relations to maintain sufficient precision. However, crucial differences with the setting of vanilla Gröbner bases from [17] make the new reduction algorithm more cumbersome.

5.1.General idea

To understand the difficulties we have to overcome in this section, let us recall how reduction algorithms work. Say that is to be reduced modulo ; then the most common way is to successively reduce the terms of in decreasing order with respect to the relevant monomial ordering. This is in particular how the relaxed algorithm from section 3.3 works (see [16] for the details). If we decompose the set of all monomials into slices of constant degree, we see that the different slices are handled one after the other (starting with those of highest degrees).

In the setting of vanilla Gröbner bases from [17], these slices are roughly parallel to the Gröbner stairs. This implies that the quotients with respect to the truncated basis are also valid with respect to the full basis (under the assumption that the degrees of the quotients are controlled, and that each is known with the appropriate precision). Then the reduction algorithm decomposes in two phases:

  1. Compute the quotients by reducing modulo . Here the speedup comes from the fact that the equation

    is much smaller (in terms of the number of coefficients) than the equation

    (4)
  2. Rewrite equation (4) to compute the remainder . First, find new quotients such that

    then compute as

    (5)

As we said earlier, Gröbner bases in the setting of this paper are not vanilla. In particular, the slices of constant degree are not parallel to the Gröbner stairs: the former have a slope of , while the latter have a slope of . For this reason, even with the bound on the degree of the , there are terms of that are not in normal form with respect to . This implies that the quotients with respect to are not valid with respect to . To maintain sufficient precision to compute the remainder, the two phases must be merged, and equation (4) is rewritten “on the fly” during the reduction algorithm. More precisely, as soon as the quotient is known, the product is replaced by some , where are known with precision larger than .

5.2.Dichotomic selection strategy

To control the degrees of the quotients, we use a dichotomic selection strategy as presented in Figure 1. The idea is to reduce each monomial preferably against one end of the Gröbner basis ( or ), or the where has the highest valuation. To describe the reduction algorithm, it is convenient to introduce the function (depending on the leading terms of the ) defined as follows:

Notice that this definition is non-ambiguous: there is only one index with maximal such that . Indeed, if have the same valuation , then there is some with and . Moreover, if and both divide , then so does .

Intuitively, if , then the term is reduced against and leads to the term in . The case corresponds to monomials that are already in normal form with respect to .

Figure 1. The dichotomic selection strategy (): monomials falling in each area are reduced against the corresponding basis element.

Lemma 12. Let . If , then .

Proof. Let , and denote . Then we observe that : if not, then would divide , whereas . A similar reasoning with (or , whenever ) shows that .

5.3.Reduction algorithm

Recall that Definition 5 gives recurrence relations between the : for any indices with , we have

(6)

This allows for substitutions in the expression that are analogous to those used in the reduction algorithm from [17]. In our setting, the substitutions used in the reduction algorithm are based on the following result:

Lemma 13. Let be a Gröbner basis with recurrence relations defined by the matrices as in Definition 5 (for all indices such that ). Given quotients define

Then we have

(7)

Assume now that is such that have degree less than and . Then:

  • If , then have degree less than and can be computed in time .

  • If , then , and they can be computed in time .

Proof. Equation (7) is an immediate consequence of the recurrence relations (6) between the . Similarly, the bounds on are consequences of the degree bounds from Proposition 6.

We can now adapt the extended reduction algorithm from [16] to perform these replacements during the computation.

Algorithm 3

Input: , where is a bivariate polynomial of degree , and is the concise representation of a Gröbner basis (as in Definition 8).

Output: , the extended reduction of with respect to .

Set

Set
// new quotients after substitutions as in Lemma 13
Set and
// : active indices for relaxed multiplications

For , set .

For :

// keep only the indices with

For

Let be the term of in , computed using relaxed multiplications.

Let .

If , then update , else update .

For all such that :
// See Remark 14

If , then update and and

If and , then:

Set .

Set .

Update .

Update

Set .

Return

Remark 14. Recall that has degree and has degree for . Therefore, the loop on such that is trivial: the set of such contains at most 1 element, except when , in which case there are 2 elements and .

Theorem 15. Algorithm 3 is correct and runs in time

Proof. Let us first explain why the relaxed strategy can indeed be used. We regard the quotients as streams of coefficients. These coefficients are produced by the updates and consumed in the relaxed evaluation of the products . Since our reduction process is essentially based on the same recursive functional equation as in [16], the production of coefficients always occurs before their consumption.

We will show that Algorithm 3 computes the same result as the traditional relaxed reduction algorithm from [16], because the term that is considered at each step is the same in both cases. More precisely, we must show that for each , at the start of each iteration of the loop on , we have

(8)

where is the non-truncated analogous of .

Let us start with the description of a few invariants at the start of the main loop on To avoid case distinctions, let be the function on defined by

Then the following invariants hold:

  1. .

  2. .

  3. and are at most for all even .

  4. For some , we have .

  5. With as above, if or or ( with even), then .

Invariant 1 is immediate. Invariants 2 and 3 follow from Lemma 13, using by Lemma 12. For invariant 4, we recall that for all . Finally, invariant 5 is an immediate consequence of the definition of the index .

Let us now prove the main claim (8). Notice first that if and , then . Recall also that for . This means for . Since by definition

we deduce that

To complete the proof (by invariant 1), we show that

For the first identity, we contend that and have the same terms of degree because of invariants 3 and 5. This is clear if , or if since and . Assume therefore that and . By invariant 5, the index verifies , hence was removed from during a previous iteration of the loop on . Since for all , this actually happened during the previous iteration (with instead of ). It follows that . By definition of and invariant 3, the polynomials and have the same terms of degree at least .

The second identity follows immediately from invariant 2. The last identity follows from the implication , so that and have the same terms of degree .

For the complexity, relaxed multiplications are used to compute the coefficients of the , whose support is a subset of the support of . Then the relaxed multiplications take time

It remains to evaluate the cost of the zealous multiplications during the rewriting steps. For each index with even, and for each , there is an update of at a cost of , followed by an evaluation of the products and at a cost of operations. This totals for operations. Summing for all , we get a total cost of for all rewriting steps. This fits in the announced complexity bound, by our assumption that .

Remark 16. The reduction algorithm can be optimized by delaying the substitutions until more quotients are known. This allows to decrease by a logarithmic factor the number of products , but leads to various technical complications. A priori, this does not change the asymptotic complexity; however, with a bound for relaxed multiplication better than , these zealous products would become predominant without this optimization.

6.Applications

Under some regularity assumptions, we provided a quasi-linear algorithm for polynomial reduction, but unlike in [17], it does not rely on expensive precomputations. This leads to significant improvements in the asymptotic complexity for various problems. To illustrate the gain, let us assume to simplify that , and neglect logarithmic factors. Then, ideal membership test and modular multiplication are essentially quadratic in . Also, computing the reduced Gröbner basis has cubic complexity. In all these examples, the bound is intrinsically optimal, and corresponds to a speed-up by a factor compared to the best previously known algorithms.

6.1.Ideal membership

From any fast algorithms for Gröbner basis computation and (multivariate) polynomial reduction, it is immediate to construct an ideal membership test:

Algorithm 4

Input: , bivariate polynomials of degrees , , and with and generic.

Output: true if , false otherwise.

Let be the concise representation of the Gröbner basis of with respect to , computed using Algorithm 2.

Let be an extended reduction of modulo , computed using Algorithm 3.

Return true if , false otherwise.

Theorem 17. Algorithm 4 is correct and takes time .

6.2.Multiplication in the quotient algebra

We designed a practical representation of the quotient algebra that does not need more space (up to logarithmic factors) than the algebra itself, while still allowing for efficient computation. The main difference with the terse representation from [17] is that said representation is easy to compute, so that multiplication in can be done in quasi-linear time, including the cost for the precomputation:

Algorithm 5

Input: , bivariate polynomials, with generic of degrees and of degree at most (typically in normal form).

Output: in normal form.

Let be the concise representation of the Gröbner basis of with respect to , computed using Algorithm 2.

Compute using any (zealous) multiplication algorithm.

Let be an extended reduction of modulo , computed using Algorithm 3.

Return .

Theorem 18. Algorithm 5 is correct and takes time .

Remark 19. If we assume or simply , then the algebra has dimension and the above bound can be rewritten . In the general case, we have , and one might wish to discard the term to achieve complexity . This term includes the computation of during the call to Algorithm 2 (which can be seen as precomputation), the multiplication , and the term in the complexity of the reduction (since here ). If and are given in normal form with respect to , then has in fact degree in the variable and only in . In this case, can be computed in time , and a refined analysis of the reduction algorithm should reduce the term to by observing that the degree in remains .

6.3.Reduced Gröbner basis

Since we can reduce polynomials in quasi-linear time, we deduce a new method to compute the reduced Gröbner basis: first compute the non-reduced basis, together with additional information to allow the efficient reduction (which can be done fast); then reduce each element with respect to the others.

Algorithm 6

Input: , generic bivariate polynomials of total degrees and with .

Output: the reduced Gröbner basis of with respect to .

Let be the concise representation of , computed using Algorithm 2.

For all :

Set .

Let be an extended reduction of modulo , computed using Algorithm 3.

Set .

Return .

Theorem 20. Algorithm 6 is correct and takes time .

Proof. Clearly is in the ideal and has the same leading monomial as . Moreover, is monic and none of its terms is divisible by the leading term of any , . This proves is indeed the reduced Gröbner basis of with respect to .

For the complexity, the call to Algorithm 2 takes time . Then, for each , the reduction requires operations. The assumption now yields the desired bound.

Remark 21. The input has size , and the output needs , while Algorithm 6 runs in time . Our complexity bound is therefore quasi-optimal only if . We expect that a truly quasi-linear bound can still be obtained in general through a refined analysis (as for Remark 19).

6.4.Perspectives

As we noted in different remarks, the announced bounds could be slightly improved at the expense of some technicalities. For example, the optimization mentioned in Remark 16 would lead to a tighter bound when considering even faster relaxed multiplication (as in [19]). Also, as noted in Remarks 19 and 21, a refined analysis is required when to give a truly quasi-linear complexity. Apart from this, several more theoretically challenging extensions can be considered.

First we notice that Algorithm 2 fails if the Euclidean algorithm raises a quotient with degree larger than 1. Generically, this should not happen, but this restriction can be limiting in practice, especially in the case of small finite fields. It is then natural to ask how to handle the case of non-normal degree sequences. We conjecture that our algorithm extends (and remains quasi-linear) to the case where the quotients have a bounded degree, with only a logarithmic number of them being larger than 1.

It would also be interesting to extend our ideas to the case of variables. Whishful thinking suggests that this might possible for fixed , but with a dependency polynomial in (as the ratio between the volumes of a pyramid and a parallelepiped based on the same vectors). Extending the dichotomic selection strategy to higher dimension is rather straightforward, but the question of the successive substitutions is more subtle.

Bibliography

[1]

Magali Bardet, Jean-Charles Faugère, and Bruno Salvy. On the complexity of the F5 Gröbner basis algorithm. Journal of Symbolic Computation, pages 1–24, sep 2014.

[2]

Thomas Becker and Volker Weispfenning. Gröbner bases: a computational approach to commutative algebra, volume 141 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1993.

[3]

Bruno Buchberger. Ein Algorithmus zum Auffinden der Basiselemente des Restklassenrings nach einem nulldimensionalen Polynomideal. PhD thesis, Universitat Innsbruck, Austria, 1965.

[4]

David G Cantor and Erich Kaltofen. On fast multiplication of polynomials over arbitrary algebras. Acta Informatica, 28(7):693–701, 1991.

[5]

Jean-Charles Faugère. A new efficient algorithm for computing Gröbner bases (F4). Journal of Pure and Applied Algebra, 139(1–3):61–88, 1999.

[6]

Jean-Charles Faugère. A new efficient algorithm for computing Gröbner bases without reduction to zero (F5). In Proceedings of the 2002 international symposium on Symbolic and algebraic computation, ISSAC '02, pages 75–83. New York, NY, USA, 2002. ACM.

[7]

Jean-Charles Faugère, Pierrick Gaudry, Louise Huot, and Guénaël Renault. Polynomial systems solving by fast linear algebra. ArXiv preprint arXiv:1304.6039, 2013.

[8]

Jean-Charles Faugère, Patrizia Gianni, Daniel Lazard, and Teo Mora. Efficient computation of zero-dimensional Gröbner bases by change of ordering. Journal of Symbolic Computation, 16(4):329–344, 1993.

[9]

M. J. Fischer and L. J. Stockmeyer. Fast on-line integer multiplication. Proc. 5th ACM Symposium on Theory of Computing, 9:67–72, 1974.

[10]

Ralf Fröberg and Joachim Hollman. Hilbert series for ideals generated by generic forms. Journal of Symbolic Computation, 17(2):149 – 157, 1994.

[11]

J. von zur Gathen and J. Gerhard. Modern Computer Algebra. Cambridge University Press, 3rd edition, 2013.

[12]

Vladimir P. Gerdt and Yuri A. Blinkov. Involutive bases of polynomial ideals. Mathematics and Computers in Simulation, 45(5):519–541, 1998.

[13]

D. Harvey and J. van der Hoeven. Faster integer and polynomial multiplication using cyclotomic coefficient rings. Technical Report, ArXiv, 2017. http://arxiv.org/abs/1712.03693.

[14]

David Harvey, Joris van der Hoeven, and Grégoire Lecerf. Faster polynomial multiplication over finite fields. Technical Report, ArXiv, 2014. http://arxiv.org/abs/1407.3361.

[15]

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

[16]

J. van der Hoeven. On the complexity of polynomial reduction. In I. Kotsireas and E. Martínez-Moro, editors, Proc. Applications of Computer Algebra 2015, volume 198 of Springer Proceedings in Mathematics and Statistics, pages 447–458. Cham, 2015. Springer.

[17]

J. van der Hoeven and R. Larrieu. Fast reduction of bivariate polynomials with respect to sufficiently regular Gröbner bases. Technical Report, HAL, 2018. http://hal.archives-ouvertes.fr/hal-01702547.

[18]

J. van der Hoeven and É. Schost. Multi-point evaluation in higher dimensions. AAECC, 24(1):37–52, 2013.

[19]

Joris van der Hoeven. Faster relaxed multiplication. In Proc. ISSAC '14, pages 405–412. Kobe, Japan, Jul 2014.

[20]

Romain Lebreton, Eric Schost, and Esmaeil Mehrabi. On the complexity of solving bivariate systems: the case of non-singular solutions. In ISSAC: International Symposium on Symbolic and Algebraic Computation, pages 251–258. Boston, United States, Jun 2013.

[21]

Ernst Mayr. Membership in polynomial ideals over q is exponential space complete. STACS 89, pages 400–406, 1989.

[22]

Guillermo Moreno-Socías. Degrevlex gröbner bases of generic complete intersections. Journal of Pure and Applied Algebra, 180(3):263 – 283, 2003.

[23]

A. Schönhage. Schnelle Multiplikation von Polynomen über Körpern der Charakteristik 2. Acta Infor., 7:395–398, 1977.

[24]

A. Schönhage and V. Strassen. Schnelle Multiplikation großer Zahlen. Computing, 7:281–292, 1971.