|
We present a probabilistic Las Vegas algorithm for solving sufficiently generic square polynomial systems over finite fields. We achieve a nearly quadratic running time in the number of solutions, for densely represented input polynomials. We also prove a nearly linear bit complexity bound for polynomial systems with rational coefficients. Our results are obtained using the combination of the Kronecker solver and a new improved algorithm for fast multivariate modular composition.
|
Let be an effective field. This means
that we have a data structure for representing elements of
, together with algorithms for performing the
field operations. Consider a homogeneous system of polynomial equations
where . We are interested in
the exact resolution of such a system through the computation of a
parametrization of its zero-set by a so-called primitive
element; see section 4.2 for a precise definition.
Throughout this paper, we assume that are given
in standard dense representation. The total degree of each
is written
,
and we let
. Our algorithm
requires the following regularity conditions:
The intermediate ideals
are absolutely radical, which means radical over the
algebraic closure
of
, for
;
Conditions admit random coefficients. Under these conditions it
is well known that the system
admits exactly
geometrically regular solutions in the
projective space
over
. Condition
We prove new probabilistic complexity bounds for computing all the
solutions in terms of a univariate parametrization by a primitive
element. For finite fields ,
our bound is arbitrarily close to quadratic in the number of the
solutions whenever
: see
Corollary 5.5. For rational coefficients of bounded height,
we deduce a complexity bound that is nearly linear in the expected bit
size of the output of the algorithm: see Theorem 6.11.
These results improve upon the best previously known complexity bounds.
We set . In order to simplify
the presentation of complexity bounds, we use soft-Oh notation:
means that
;
see [21, chapter 25, section 7] for technical details. The
least integer larger or equal to
is written
; the largest one smaller or
equal to
is written
.
The -vector space of
polynomials of degree
is written
. The remainder of the division of
by
is written
. In order to express that an expression
is explicitly computed modulo
, we write
.
The Galois ring of characteristic and
cardinality
, written
, is defined as
with monic and irreducible modulo
.
The complexity of polynomial system solving is a central question in
effective algebraic geometry, which is still open. Except in some very
special cases, no softly linear time algorithms are known. In order to
simplify the discussion, we assume that for
.
One known favorable special case concerns systems of bounded degrees
over
that satisfy
. This
solver is probabilistic of Las Vegas type, with a well understood
probability of failure that can easily be made arbitrarily small in
practice; for details, see the recent work by Giménez and Matera
[22]. The Kronecker algorithm has a long history, lying
deep in Kronecker's work [44], but it emerged after a
series of seminal articles by Giusti, Heintz, Matera, Morais, Pardo, and
their collaborators in the nineties [24-26];
for the history of this algorithm along with references, we refer to the
introduction of [18].
Assume that the input polynomials are given as a straight-line program
of size over
without
division (see the definition in [13] for instance) and that
we use the computation tree model for our complexity measures. If the
cardinality of
is sufficiently large and the
conditions
![]() |
(1.1) |
operations in and with a uniformly bounded
probability of failure. The complexity bound (1.1)
essentially comes from [27, Theorem 1] by taking into
account that
; this is
detailed in section 5 below. If the
are all bounded, then we notice that
,
so the bound (1.1) is essentially quadratic. In the other
extreme case when
is bounded, we rather have
, so the bound (1.1)
becomes essentially cubic. In this paper, we show how to obtain a nearly
quadratic bound for this case as well.
Gröbner bases constitute another cornerstone for polynomial system
solving. Algorithms in this setting are usually deterministic, but their
complexities are highly intricate in general: they depend on both the
system and the admissible ordering attached to monomials. Complexities
of unfavourable cases are doubly exponential in ; see a summary of known families in [21,
chapter 21]. The first simply exponential bounds of the form
for zero dimensional systems are due to Giusti, Lazard and
Lakshman in the eighties [23, 45-48].
The constant hidden behind the latter “
” may be related to a feasible exponent
for linear algebra, in the sense that any two square
matrices of size
may be multiplied with
operations in their coefficient ring. Basically one
may perform Gaussian elimination on Macaulay matrices: for instance,
under
operations in
,
where
is the Macaulay bound; see [4],
or [8, chapter 26] for a proof from scratch, for instance.
At present time it is not known how to exploit the structure of Macaulay
matrices to perform Gaussian elimination in time quadratic in or even in
.
One contribution to this problem, due to Canny, Kaltofen, and Yagati [14], relies on sparse polynomial interpolation and the
Wiedemann solver. Further important results based on structured linear
algebra have been established by Mourrain, Pan, and Ruatta [56]:
they proved that the number of roots and of real roots can be computed
with
floating point operations, and that all the
(real) roots in a given box up to the error
can be computed with
floating
point operations, where
and
depend linearly on
as well as on the
conditioning of the system. Yet another important advance here is
Faugère's F5 algorithm, which suppresses useless rows in Macaulay
matrices. However, its worst case complexity
remains essentially cubic in
[3, 4]. Other methods derived or inspired by Macaulay's work have
been developed by Mourrain and Trébuchet [57-59].
For Gröbner bases of zero dimensional systems, monomial orderings
may be changed quite conveniently by means of linear algebra. This is
known as the FGLM algorithm (named after the initials of its authors,
Faugère, Gianni, Lazard and Mora [20]) and involves
a cost . Recently,
Faugère, Gaudry, Huot and Renault [19] decreased it
to
. Nevertheless one must
keep in mind that the best practical values for
are only about
, which is far
from the best known theoretical bound
,
due to Le Gall [49].
Another aspect of the present paper concerns multivariate modular
composition, that is the computation of where
,
has
degree
, and
are in
. When
, the best known complexity bound
for any field
is due to Brent and
Kung [11]. Their algorithm even yields a sub-quadratic cost
when using fast linear algebra; see [40,
p. 185]. Here the constant
is such that a
matrix over
may be multiplied
with another
rectangular matrix with
operations in
.
Huang and Pan proved that
;
see [38, Theorem 10.1]. In [35, section 3] we
extended Brent and Kung's result to the multivariate setting. A major
breakthrough for the modular composition problem is due to Kedlaya and
Umans [41, 42], in the case when
is a finite field
and more
generally a finite ring of the form
for any
integer
and
monic of
degree
. For any fixed
rational value
, they showed
that
can be computed modulo
in time
whenever the partial degrees of
are
, and
under the condition
; see [42, Theorem 7.1]. Extensions to compositions modulo triangular
sets can be found in [63].
The first technical but important contribution of this paper concerns
the refinement of Kedlaya and Umans' complexity result on modular
composition [41, 42]. In fact, in Corollary 3.6 we achieve a sharper complexity bound in terms of the
total degree of
,
that is softly linear in the bit size
of the
coefficients, and that holds without the assumption
. We also handle fast evaluations of
in a larger algebra of the form
, where
and
is monic in
of degree
: this is a particular case of composition modulo
triangular sets studied in [63], for which we also improve
upon the dependency in the coefficient size. These new results are based
on our recent advances on multivariate multi-point evaluation algorithms
[36]. At a more abstract level, our algorithm applies to
any field
over which fast multi-point
multivariate polynomial evaluation exists. We recall that fast
multivariate multi-point evaluation remains an important open problem:
no efficient algorithms are known over a general abstract field
and we are not aware of any efficient implementations
of Kedlaya and Umans' method; for some recent progress on special cases,
we refer to [33, 34].
Our main contributions concern the complexity of polynomial system
solving. We first prove a Las Vegas type probabilistic bit complexity
bound over the finite field
, for the dense representation of input
polynomials, and where
is any fixed rational
value: see Corollary 5.5. Whenever
, this bound is arbitrarily close to quadratic in
the number of the roots
of the system. This
improves upon previously known bounds. For example, the complexity bound
(1.1) simplifies to
whenever
. With input polynomials given in
dense representation, the quantity
is of the
order
. If all the
remain bounded and if
tends to
infinity, then we have
Consequently, the expected complexity of the Kronecker solver becomes
, in terms of the number of
arithmetic operations in
.
However, if the
are no longer bounded, then
may grow with
and the worst
case complexity of the solver becomes
.
Our next main result concerns the case .
We assume that
have integer coefficients of bit
size
, and we achieve
expected time
. Whenever
, this time turns out to be
nearly linear in the bit size of the output: see the precise bound in
Theorem 6.11. This complexity bound admits major
consequences for symbolic real solving, especially for algorithms based
on the Kronecker solver; see for instance [2] and
references therein. Once a Kronecker parametrization is known, it is
also possible to appeal to univariate numerical solvers to deduce
complex or real approximations of the roots of the system; see for
instance the book [55].
Let us briefly describe the structure of this paper. Section 2 is devoted to prerequisites about the complexity model and fast multi-point polynomial evaluation. The next section 3 deals with various particular cases of multivariate modular composition involved in the Kronecker solver. Section 4 concerns the algebraic data structures needed by the solver, along with general position and some of the randomness aspects. The solver is itself recalled in section 5, that ends with our main result over the finite fields. Our main complexity bound over the rational numbers is stated in the final section 6.
It may be helpful to notice that the probability analysis of our algorithms contains two main stages. On the one hand, we need to put the system in general position and prepare the main data structures. This part is handled in section 4. On the other hand, for the sake of efficiency, some subalgorithms of the Kronecker solver involve additional random parameters that must be taken outside certain proper closed subvarieties (see Propositions 5.2 and 5.3), or integer parameters that must be coprime with a certain set of bad primes (see Lemma 6.7). The probability of picking suitable random parameters of this kind will be analyzed separately.
Throughout this paper, we will be working in the Turing machine model [61] with sufficiently many tapes (seven tapes are usually sufficient in order to implement basic routines on polynomials, series, and matrices in a natural way). The Kronecker algorithm is randomized, so it also requires a special instruction to generate a random symbol in one cell within constant time.
In some cases, algebraic structures have a natural bit size (e.g.
modular integers, finite fields); in other cases the size is variable
(e.g. arrays, polynomials). In both cases elements are represented on
tapes as sequences of symbols, followed by a specific termination symbol
””. The reader
must keep in mind that heads of the machine can just move one cell left
or right at each time step. Algorithms must take care of using data in
the most contiguous way as possible, and loop counters cannot be used
for free.
The rest of the section gathers standard data types and elementary operations needed in the next sections. We freely use well known complexity bounds on polynomials and matrices from [21]; details about Turing machine implementations of these algorithms can be found in [65].
We use binary representation for integers. A modular integer in is represented by its natural preimage in
. Integers can be added in linear
time and multiplied in softly linear time
;
see [30] and historical references therein. Integer
divisions and extended gcds in bit size
also
take softly linear time [64]. We will also use truncated
-adic integers, for which we
refer the interested reader to [6] for practical
algorithms.
One dimensional arrays are sequences of elements terminated by
“”. For example
the vector
is stored as
.
For bidimensional arrays we use column-major representation. This means
that an array of size
(
rows and
columns), is
stored as the vector of its columns, i.e.
. Such arrays are allowed to contain elements
of different types and sizes. For example the matrix
over
is stored as
.
In the Turing machine model, we recall that the transposition of a
bidimensional array can be achieved in softly linear time:
be an
matrix. Let
denote the size of
for all
, and define
. Then
can be
transposed in time
.
For univariate polynomials we use dense representation, which means that
a polynomial of degree is stored as the vector
of its
coefficients from degrees
to
. Additions
and subtractions take linear time in
.
Multiplying two polynomials over finite fields may be achieved in softly
linear time [15, 29, 31].
In the present paper we consider finite rings of the form with
and
monic of degree
. Elements of
are stored as their natural preimage in
with coefficients in
.
Additions and subtractions in
take linear time,
products take time
; see for
instance [21, part II].
be as above and let
be polynomials in
. For any
sequence of points
in
we may compute
in time
.
be a Galois ring, let
be elements in
such that
is invertible modulo
for all
, and let
be vectors
in
. We may compute the
unique polynomials
in
such that
for
and
in time
.
For a polynomial we use the recursive dense
representation, by regarding
as an element
of
. In particular,
admits the same representation as its expansion
as a univariate polynomial in
. In our algorithms, the number of variables
is not part of the representation of
, so it must be supplied as a separate
parameter.
The support of
is defined as the set of monomials with non-zero coefficients and we
write
for its cardinality. Assuming that, apart
from the mandatory trailing “#” symbol, the representations
of coefficients in
do not involve other
“
” symbols (this
can always be achieved through suitable renamings
), we denote the number of “
” symbols involved in the representation
of
by
.
We notice that
. For the
remainder of this subsection, we assume that the size of elements in
is bounded by a constant
.
Example of degree
is represented by
.
The bivariate polynomial
is represented by
.
be of partial degree
in
for
.
Then
, where
.
be a non-zero polynomial of total degree
. Then
, where
.
of a non-zero polynomial
can be computed in time
, where
.
of a polynomial
can be computed in time
,
with the convention that
.
be as
above. Any partial derivative of
of total
degree
can be computed in time
, where
.
Proof. In order to compute the partial derivative in
, we run over all the
coefficients of
in time
by Lemma 2.6.
If we are interested in the derivative in with
, then we reinterpret
as an element of
and its
derivative requires time
.
be as
above, let
be of total degree
, and let
.
Then
may be computed in time
, where
.
Proof. We use Horner's method to evaluate univariate
polynomials over at
in
softly linear time. In total, the machine runs over the entire
representation of
and the time spent in
arithmetic operations is at most
.
be as above, let
, and let
be elements in
. Then
can be computed in time
where is the cardinality of the support of
in the variables
.
At several places we will use the following bound on the cardinality of
the support of a polynomial in variables and
total degree
:
![]() |
(2.1) |
For the sake of efficiency, a homogeneous polynomial in the variables
has a dedicated
representation made of
and
. In this way,
can be
recovered as
.
The problem called multivariate multi-point evaluation over an
effective ring is the following: given
and points
in
, compute
.
The first following statement concerns the cost of the naive evaluation
algorithm in terms of the total degree of
.
with
monic of
degree
be as above, let
be of total degree
. Then
we may evaluate
at a point
in time
.
In terms of the partial degrees of ,
we will need the following better complexity bounds that refine previous
work by Kedlaya and Umans [41, 42]. The first
theorem takes partial degrees into account, while the second one is in
terms of the total degree.
be a fixed rational value. Let
with
monic of degree
, let
be
of partial degree
in
for
, and let
be a sequence of points in
.
Then we may compute
in time
be a fixed rational value. Let
with
monic of degree
, let
be
of total degree
, and let
be a sequence of points in
. Then we may compute
in time
The Kronecker solver needs to perform a linear change of variables of
matrix into the input homogeneous polynomials.
Let
be such a polynomial in
of total degree
, and let
be a
matrix over
. We need to compute
A natural idea is to interpolate from its values
at
, where
is a subset of
of cardinality
. Such an interpolation is known to be
computable in time
; see
Lemma A.8 of the appendix. On the other hand the values of
at
may be obtained in
time
by [36, Proposition 5.7], for
any fixed rational value
.
However, this approach suffers from two drawbacks: it takes neither the
homogeneity of
nor the structure of
into account. At the end of Appendix A, we
show the following sharper complexity result:
be homogeneous of
degree
, and let
be an
matrix over
. If
and if a
LU-decomposition of
is given, then we may
compute
in time
.
This section is devoted to multivariate modular composition, that is the
evaluation of a multivariate polynomial in at a
point in
, where
with
monic and
. As recalled in the introduction, no fast
algorithm is known for this task over a general ring or field
. For the Kronecker solver
presented in this paper, the following particular instances of
are needed:
and
, along with tangent numbers over theses rings.
In this section, we design fast algorithms for these cases, on the basis
of Kedlaya and Umans' ideas, and new variants from [36].
If , then
is the finite field
, whereas
corresponds to the ring
of the
-adic integers
truncated to precision
.
Until the end of the section we set
![]() |
(3.1) |
and
where is a monic polynomial of degree
in
.
Let represent a cost function for the
multi-point evaluation of
-variate
polynomials over
with partial degrees
, total degree
, and for
evaluation
points. The following algorithm reduces modular composition to
multi-point evaluation.
Algorithm
Write
for the canonical preimage of in
, where
and
belong to
and have partial degrees
in
and
in
. In a similar manner write
for the preimage of in
.
Build points whose reductions modulo
are pairwise distinct.
For and for
, compute
For compute
and
.
For and for
, compute
Interpolate the polynomials and
in
,
of partial degrees
in
and
in
,
and such that
and
for all
and
.
Return reduced modulo
.
Proof. First we verify that
Therefore, for all and
we obtain
Since has partial degrees
in
and
in
, we deduce
whence the correctness of the algorithm.
The rewriting in step 1 takes linear time. The construction of the
points in step 2 costs . In
step 3 we appeal to Lemma 2.2 in order to obtain
in time , and then to obtain
for in time
.
Notice that the required data reorganizations can be done in softly
linear time thanks to Lemma 2.1.
The cost of step 4 is provided by Lemma 2.11,
and simplifies to by (2.1) and
Lemma 2.6.
The derivations in step 5 incur by Lemma 2.9. Taking into account data reorganizations via
Lemma 2.1, the total cost of step 5 is
In step 6, the interpolations amount to by Lemma
2.3. Finally step 7 takes additional time
.
Our main complexity bound for modular composition is presented in the next theorem, under the condition that sufficiently many evaluation points are at our disposal. For convenience we recall the following lemma.
be a fixed rational value, assume
that
and that
![]() |
(3.2) |
Let be of total degree
and let
be in
,
with
monic in
of
degree
. Then
can be computed in time
Proof. Let be such that
First we examine the case .
Lemma 3.2 and the fact that the function
is nondecreasing yield
so we have . By adapting
Lemma 2.12 to
,
the cost of the naive modular evaluation is
, since
.
From now we assume that holds. We set
. If
is
bounded then so is
and the above adaptation of
Lemma 2.12 may again be used to achieve a naive evaluation
cost in
. Consequently, we
may further assume that
is sufficiently large to
satisfy
![]() |
(3.3) |
We perform the multivariate Kronecker segmentation of
into
which has
variables
and partial degrees
as follows. We let
and introduce the Kronecker substitution map
We set and
.
Observe that we have
and
. From
and
we deduce that
Still by allowing to be sufficiently large we
may further assume that
![]() |
(3.4) |
On the one hand, by [36, Proposition 5.1] and (3.3),
the computation of takes time
On the other hand, for we compute
by binary powering, in time .
We may now compute
as
Since (3.4) ensures
the condition (3.2) allows us to combine Proposition 3.1 and Theorem 2.13. This yields the complexity bound
![]() |
|||
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
|||
![]() |
![]() |
||
![]() |
(using (3.4)) | ||
![]() |
![]() |
for a suitable constant . If
, then this bound further
simplifies to
Otherwise we have which implies that
. In this case, the condition (3.2) allows us to combine Proposition 3.1 and
Theorem 2.14, in order to obtain
in
time
which simplifies to thanks to
.
Overall we have proved a complexity bound .
Taking into account that
,
from (3.3), it follows that
,
which concludes the proof.
In order to generalize Theorem 3.3 to small residue fields, namely whenever inequality (3.2) does not hold, we extend the cardinality of the ground ring. We will rely on the following construction.
be a Galois ring, and let
be an integer that is coprime with
. Then we may compute a monic irreducible
polynomial
of degree
in time
with any fixed arbitrarily low
probability of failure, or in expected time
with a Las Vegas algorithm. Then we may build the Galois ring
as
,
such that the following isomorphism holds:
for some polynomials and
in
. The polynomials
,
and
can be computed in time
. Each application of
and
takes time
.
Proof. The polynomial can be
obtained by using Ben-Or's randomized algorithm [21,
Theorem 14.42]. Since
and
are coprime, a natural candidate for
is the
composed product of
and
(usually written
),
![]() |
(3.5) |
It is well known that is irreducible of degree
. Let
, that can be computed in time
. We construct
as the
monic part of the resultant
in
and we write
for the corresponding subresultant
of degree
in
.
For subresultants over Galois rings, we need to compute the necessary
minors of the Sylvester matrix by means of Berkowitz' algorithm [5]. In this manner
,
, and
can be obtained in time
.
It is well known that is invertible modulo
, so we compute
and
in time
.
Let us consider an element represented by
, and
where . Equivalently, we have
which corresponds to the usual Chinese remaindering formula associated
to the factorization (3.5) of .
We observe that for all roots
of
. Consequently, for all
roots
of
with
, we have
. It follows that
actually
represents
. It can be
computed in time
.
In the other direction, given a preimage of
, we obtain
as
, in time
.
Remark in specific cases; see [9,
28, 51]. For simplicity we have preferred to
restrict to general well known algorithms, because the complexity bound
of Lemma 3.4 is to be used only when
is rather small.
The following corollary aims at completing Theorem 3.3 for small residue fields. It will not be used in the sequel because the extension of the residue field will instead be handled at the top level of the Kronecker solver.
be a fixed rational
value, let
be as in (3.1), and
assume that
. Let
be of total degree
and let
be in
,
with
monic in
of
degree
. Then
can be computed using a probabilistic Las Vegas
algorithm in expected time
Proof. Notice that the total degree of may be obtained in time
by Lemmas
2.6, 2.8 and inequality (2.1). It
remains to handle the case when (3.2) does not hold in
Theorem 3.3. In other words, assume that
We first compute the smallest integer such that
that is
and we set
so is coprime to
and
We next apply Lemma 3.4, from which we borrow the notation:
of degree
is obtained in
expected time
; the
computation of
,
, and
requires time
.
With this construction in hand we set
and
and we proceed as follows for the modular composition problem:
We cast into
in time
;
We cast into
in time
;
We cast into
in time
;
By Theorem 3.3, we can evaluate
in
in time
We cast into
in time
.
Adding up, we obtain the claimed complexity bound—up to replacing
by
from the outset.
Consider a polynomial system over an effective
field
, which satisfies the
regularity assumptions
Under our regularity assumptions, and letting , the coordinate ring
has
dimension
and degree
by
the Bézout theorem. In particular the system
admits
distinct solutions in
, which are all regular. After applying a
generic affine change of coordinates, the Kronecker solver first solves
the system
then
…,
and so on until
. We first
study how to perform this change of variables. Next we explain how
positive dimensional solution sets of the intermediate systems
are represented.
An homogeneous ideal of
is said to be in Noether position when
and
is an integral ring extension of
, where
. Let
be the column vector
with entries
. If
is an invertible
matrix over
, then we write
If is generated by a proper regular sequence
, then we say that a matrix
puts the intermediate ideals
in simultaneous Noether position if
is
in Noether position for all
.
Let us now examine the probability that this happens for an upper
triangular matrix
with ones on the diagonal and
random entries above the diagonal.
satisfy
of
of cardinality
, consider the matrix
where the are taken at random in
. Then the probability that
puts
into simultaneous
Noether position is at least
.
Proof. We use the incremental method described in
the proof of [18, Theorem 1.9] as follows (for historical
references, we also refer to [18]). The values are taken such that the coefficient of
in
is non-zero. For such values the ideal
is in Noether position.
Then we consider the determinant of the
multiplication endomorphism by
in
: it belongs to
and has
degree
. The values
such that the coefficient of
in
is non-zero put
into
Noether position.
By induction we consider the determinant of the
multiplication endomorphism by
in
: it belongs to
and has
degree
. The values
such that the coefficient of
in
is non-zero put
into
Noether position. By taking into account the fact that
, the Schwartz–Zippel lemma [66,
69] leads to a probability of success at least
Let be an absolutely radical ideal of
such that the coordinate ring
is zero dimensional—here we consider the affine case when
is not necessarily homogeneous. A polynomial
is said to be a primitive element for
if the projections of its powers generate
as a
-vector
space. Let
represent the degree of
(i.e. its dimension as a
-vector space).
of
, the
probability that a random vector
induces a
primitive element
of
is
.
Proof. The minimal polynomial
of
in
actually belongs
to
when the
are regarded
as parameters, and the total degree in the
is
; see for instance [18,
Corollary 3.4]. The points
to be avoided are the
zeros of the discriminant of
,
seen as a polynomial of total degree
in
. We conclude by the aforementioned
Schwartz–Zippel lemma.
For a primitive element there exist unique
polynomials
such that
,
is monic and separable,
, and
The polynomials are called the
parametrization of
by
. Equivalently, we may define
for
, whence
These polynomials are called the Kronecker
parametrization of
by
; such a parametrization is uniquely determined
by
.
satisfy
are simultaneously in Noether position for
. Given a finite subset
of
, the probability that a
vector
induces a primitive element
common to all
for
is
.
Proof. Each has dimension
as a
-vector
space [18, section 2.4 and the relationship to the
geometric definition of the degree of an algebraic variety in Corollary
3.10]. By the previous lemma and the fact that
, the probability that
is
primitive for all
is at least
Let satisfy
for . Thanks to Noether
positions, the system
has all its roots in the
affine space. More generally given a point
all
the roots of
are in the affine space. A point
is called a lifting point if
is absolutely radical.
of
, the
probability that a vector
is a simultaneous
lifting point for
is
.
Proof. Without loss of generality we may assume that
is algebraically closed, so we may consider a
primitive element
that is common to all
, thanks to Lemma 4.3.
The minimal polynomial
of
in
is homogeneous in
of
degree
, and is monic and
separable in
. Any point
such that the specialization of
at
is separable ensures that
is absolutely radical; see for instance [18, Proposition
3.6]. The probability of picking such a point at random in
is at least
by the
Schwartz–Zippel lemma. The probability that
satisfies the requested property is thus at least
again by using .
Assume that we are given a polynomial system
that satisfies
:
satisfy
be a
finite subset of
. If we
take
in
at random, as
well as
in
,
and if we let
then after replacing
by
.
Proof. We first pick up a matrix
at random as in Lemma 4.1 in order to get simultaneous
Noether positions. Then comes the choice of the lifting point, for which
we use the strategy presented in the proof of Lemma 4.4.
At the end of the resolution, unless the execution has failed, we expect
to obtain an invertible matrix
over
and polynomials
such that
is separable of degree
, the
have degree
, and
On the other hand, being absolutely radical, its
roots are all regular,
is separable, and the
value of the Jacobian matrix of
in
at
is invertible modulo
; see for instance [18,
section 4.2].
One natural question at the end of a probabilistic resolution process is
to determine whether the computed solution set is correct or not. In the
case of sufficiently generic systems, namely satisfying
and
regular solutions have been
found. This idea is formalized in the following proposition, which turns
a probabilistic resolution algorithm into a Las Vegas one.
in
, an
invertible
matrix
, a linear form
,
and polynomials
in
that satisfy the following conditions:
is separable of degree
, the
have degree
,
,
for all
,
the Jacobian matrix in of
is invertible at
modulo
.
Then, is in Noether
position, and
form a parametrization of
by
.
Let be a fixed rational value, and assume
that
and if
then the above conditions can be checked in time .
Proof. The assumptions ensure that the system admits
distinct isolated
regular solutions in the affine subspace defined by
. By Bézout's theorem, this system has
therefore no solution at infinity, namely in the hyperplane defined by
.
For , let
denote the variety of zeros of
.
Let us first show that
for which
and
. In other words
would have a (non-empty) equidimensional component
of dimension
and a possibly empty component
of dimension
.
Heintz' version of the Bézout theorem [32, Theorem
1] would imply that
whence . Therefore the number
of isolated roots of
would be
which contradicts the fact that admits
distinct isolated regular solutions.
Let be in
.
We consider a non-zero homogeneous polynomial
in
. We write
into
with
primitive as a
polynomial of
and
a
monomial in
. In particular
is monic in
and
vanishes at all the points of
, so it belongs to
.
This shows that the class of
in
is integral over
when seen as an element of
. We deduce that
is in Noether position; for instance by using [18,
Theorem 1.12].
Now let us show that R is satisfied. Consider an
irreducible component
of
for some index
and then a point
in
, which is non-empty in
the projective setting. This point
is a solution
of
, and as such, the value
of the Jacobian matrix of
at
has full rank by hypothesis. Therefore, the Jacobian matrix of
has full rank over a Zariski dense subvariety of
. It follows that the primary
component associated to
is necessarily prime. On
the other hand, R
implies that
is unmixed by [53, chapter 7, section 17]; in other words,
has no embedded primary ideal. Consequently
is absolutely radical.
Assume now that , and let us
examine complexities. Testing the separability of
boils down to computing its gcd with
,
which takes time
. We compute
in time
by Proposition
2.15. Then we compute the Jacobian matrix
in
of
in time
by Lemma 2.9. The evaluations of all the
and of
at
in
can be done in time
by means of Theorem 3.3—recall that these homogeneous
polynomials to be evaluated are already represented by their
specialization at
. Finally
the determinant of the latter value of
is
obtained without division thanks to Berkowitz' algorithm [5]
in time
. Testing the
invertibility of this determinant simply requires additional time
.
Let be an effective field and let
be homogeneous polynomials in
. Throughout this section, we assume that
conditions
, we assume that a
parametrization of the ideal
by a primitive element is given, so that
We will build primitive elements for all intermediate ideals from a
unique tuple . In order to
obtain a similar parametrization for
,
we apply the two following steps.
The parametrization of is lifted into a
parametrization of
of the form
where is a monic polynomial of degree
in
,
and the
are in
. It turns out that
actually belong to
and have total degrees
; see for instance [18, Corollary 3.4]. We will compute them by a variant of
the Newton–Hensel lifting strategy over
.
Geometrically speaking represents the
affine curve
. The
intersection of this curve with the hypersurface
corresponds to
. The
minimal polynomial of
modulo
will be computed as a resultant, and the rest of the
parametrization of
is completed by means
of a suitable deformation technique. We define
to be such that this parametrization is of the form
The lifting step essentially relies on a multivariate variant of the
Hensel lemma (or, equivalently, on a multivariate power series analogue
of Newton's method). The Jacobian matrix of in
is written
The identity matrix of size is written
.
Algorithm
Initialize ,
,
,…,
.
Initialize with the inverse of
modulo
.
While do the following:
update
and over
;
set ;
compute
over ;
compute ;
update computed over
;
for from
to
,
update computed over
.
For from
to
, compute
over
.
Return truncated to order
and then regarded in
.
be a fixed rational value. If
and if
then it takes time .
Proof. This algorithm directly comes from [27, section 4]. We do not repeat the correctness proof but focus on the complexity analysis in the Turing model for the dense representation of polynomials.
The partial evaluations of at
are achieved in time
by Lemma 2.10
and (2.1)—recall that the homogeneous polynomials are
actually represented by their specialization at
and their total degree. All the partial derivatives needed for the
Jacobian matrix can be obtained in time
by Lemma
2.9.
In step 2, we compute in time
by Theorem 3.3. Then the inversion of
is performed by Berkowitz' algorithm in time
.
In step 3 we regard and the entries of
as polynomials in
:
their evaluations at
take time
by Theorem 3.3. The remaining computations in steps 3 until
5 take time
.
For the intersection step we follow the method designed in [27,
section 6]. We first compute a deformation of the parametrization of
by the primitive element
where the are new variables satisfying
for all
. Let
and let
be a shorthand
for
. The extension
of
to
is
parametrized by
as follows
where is monic in
of
degree
, and the
have degree
in
, for
.
Of course,
respectively coincide with
modulo
. In
addition we know that the total degrees of
in
and
is
; see [18, section 3].
Algorithm
Compute .
Compute .
For from
to
do the following:
compute ;
compute ;
compute .
Return truncated to
and regarded in
.
values of
in
,
Algorithm 5.2 is correct. If
, then it takes time
.
Proof. This algorithm corresponds to [27,
Algorithm 8]. Its correctness is explained in [27, section
6.4]. We adapt the complexity analysis to our context. In steps 2 and
3.a the inverse of modulo
is only required modulo
. If
we were performing this modular inversion over
then we would have to handle rational functions of degrees growing with
. In order to avoid this
coefficient swell we use truncated power series instead of rational
functions. In fact, we compute the inverse of
modulo
for
specialized
to
, then we use classical
Hensel lifting to recover
.
This method works fine whenever the resultant of
and
does not vanish at
. This resultant has degree
in
.
Algorithm 5.2 needs to shift by
in the input and by
in
the output, which can be done in time
using a
divide and conquer algorithm; see for instance [7, Lemma
7]. Then
is obtained in time
. The rest of the algorithm performs polynomial
products of cost
.
If is a polynomial in
of
total degree
in
and
, and assuming
, then we write
for the polynomial in obtained after replacing
by
.
This substitution can be performed efficiently as follows:
Compute with a softly linear number of
operations in
;
Compute . This
substitution is applied to each homogeneous component of
seen in
.
In fact if
is the homogeneous component of
degree
then we are led to compute
, which basically reduces to
computing
, and that
corresponds to a univariate polynomial shift. Such a shift is known
to take softly linear time, as mentioned above.
If then this substitution in
takes time
. We are now ready
to complete the intersection step.
Algorithm
Compute .
Compute and
.
For from
to
compute:
;
.
Let .
Compute with
in
.
Return truncated modulo
and seen as polynomials in
.
be a finite subset
of
. If we take
in
and then
in
at random, then Algorithm 5.3
works correctly as specified with probability
. Let
be a fixed rational
value, and assume that
. If
, and if
then Algorithm 5.3 takes time .
Proof. The algorithm is borrowed from [27, Algorithm 7]. Again we only focus on the complexity and probability analyses in our context.
First we need to compute along with the
corresponding Bézout relation that yields
. We wish to apply a fast subresultant algorithm,
namely either the one of [21, chapter 11] or the one of [51]. For this purpose, we need to ensure the following
properties:
the leading coefficient (in degree )
of
is non-zero modulo
,
the leading coefficient, in degree ,
of
is non-zero modulo
,
the non-zero subresultant coefficients from degree
to
are non-zero modulo
.
In fact, these coefficients are the only quantities that need to be inverted during the execution of the aforementioned fast subresultant algorithms.
We claim that these properties are met for sufficiently generic values
of the . In order to prove
and quantify this claim, we introduce
where the are new independent variables, and set
. The extension
of
to
is
parametrized by
as follows
where
actually belong to
,
is monic, separable, and of degree
in
,
have degree
in
,
have total degree
in
and
,
have total degree
in
.
For the proofs of these facts we refer the reader to [18,
section 3]. It follows that are the respective
specializations of
at
. We further introduce
By the specialization property, the subresultants of
and
, seen as polynomials of
respective degrees
and
in
, are the specializations
of those of
and
at
. Such a subresultant coefficient
of
and
is a specific
minor of the Sylvester matrix of
and
, of size at most
. Therefore such a subresultant coefficient is
a (non-reduced) rational function with denominator
and numerator in
of total degree
in
.
Consequently the product of all these non-zero numerators yields a
non-zero polynomial
of total degree
such that implies that the
non-zero subresultant coefficients of
and
are invertible; in particular the resultant is
invertible and so are the leading coefficients of
and
.
The fast computation of raises the same issue,
and requires the same kind of probability analysis. We thus introduce
which belongs to , has degree
in
,
in
,
and total degree
in
. In this way
is the specialization of at
.
A subresultant coefficient of and
, seen as polynomials of respective degrees
and
,
is a (non-reduced) rational function with denominator
and numerator in
of total degree
in
.
Consequently there exists a non-zero polynomial
of total degree
such that implies that all the non-zero
subresultant coefficients of
and
and the leading coefficients of
and
are invertible.
As in the proof of Lemma 4.2, there exists a non-zero
polynomial of total degree
such that
implies that
is a primitive element for
.
This ensures the correctness of the output.
Now assume that . This
happens with probability
whenever is taken at random in
.
The subresultant coefficients of and
have degree
in
, so except for
values
of
, we can compute
modulo
in time
, by means of the fast subresultant algorithm.
A subresultant coefficient of
and
has degree
in
, so except for
values
of
,
can be computed by means of the fast subresultant algorithm modulo
in time
.
Consequently a suitable value for
is found at
random in
with probability of success
All the substitutions take
time according to the above discussion. All the shifts by
and
in
totalize time
.
The steps 3.b take time in total. In step 4 we
first specialize the variables
to
in time
, after
which the rest of the evaluation of
is decomposed into evaluations in
for
. More
precisely, for the evaluation in
we ignore the
variables
for
and we may
thus appeal to Theorem 3.3 in order to achieve time
Putting together the propositions of this and the previous sections, we obtain the following complexity bound.
be a fixed rational value, let
and let
be homogeneous
polynomials in
of total degrees
such that:
Then an invertible matrix ,
a primitive element
and a Kronecker
parametrization of
can be computed in time
, with a probability of
failure that is bounded by
.
Proof. This result is a consequence of the Kronecker
solver that was recalled at the beginning of this section. First we pick
up an invertible matrix of size
with entries at random in
as described in Lemma
4.5, in order to ensure that after replacing
by
conditions
Computing takes time
by
Proposition 2.15.
On the other hand the probability that are
suitable for all the intersection steps is at least
by Proposition 5.3. The probability that a random value of
is suitable for all stages of the solver is at
least
Overall the probability of failure is bounded by . Notice that the constant
is not intended to be optimal. The total complexity follows by combining
Propositions 5.1, 5.2 and 5.3.
be a fixed rational
value, let
and let
be
homogeneous polynomials in
of total degrees
such that
Then the Kronecker algorithm computes an integer , the finite field
, the images of
into
, an invertible matrix
with entries in
,
a primitive element
and a Kronecker
parametrization of
over
in time
with any arbitrarily small fixed
probability of failure, or in expected time
with a Las Vegas algorithm.
Proof. The case
is already handled in Theorem 5.4, and corresponds to . So we focus on the opposite case
, for which
. We compute the first integer
such that
that is
Then we set
so is coprime to
and we
have
. Now we appeal to Lemma
3.4: we construct
in time
with probability of failure
;
then the casts between
and
can be achieved in time
.
With this construction at hand we proceed as follows:
We cast into
in time
;
We call the Kronecker solver over .
In view of Theorem 5.4, the resolution over
takes time
with a probability
of failure
.
Using Proposition 4.6, the correctness of the result can be
verified in time as well. We are thus able to
reach any arbitrarily small fixed probability of failure by repeating
the resolution sufficiently many times. Altogether, this leads to a Las
Vegas algorithm that satisfies the claimed expected complexity
bound—of course,
needs to be replaced by
from the outset to conclude the proof.
Remark are
, then it is far more efficient to
compute:
a permutation of
,
, linear forms
,
homogeneous polynomials of total degrees
,
such that for
and
for
. In
this way we are reduced to solve a “smaller” system with all
degrees
. This simplification
in the input system can be achieved in time
by
means of Proposition 2.15. We leave out the details for the
sake of conciseness.
From now we turn to solving polynomial systems
with coefficients in
, still
under assumptions
, and we appeal to the algorithm designed in [27, section 4.6]: we first perform the resolution modulo a
suitable prime number
, and
then lift the parametrization of the solution set over the
-adic numbers by means of Newton–Hensel
lifting. Roughly speaking, once the bit size exceeds twice the maximum
bit size of the integers of the Kronecker parametrization over
, then this parametrization can be
reconstructed from its
-adic
expansion.
In order to formalize this approach, we first need to estimate the bit
size of the Kronecker parametrizations in terms of the input size. Then
we design a probabilistic algorithm that generates suitable values for
with high probability.
For a scalar, a vector, a matrix, or a polynomial
over
we write
for the
maximal norm of its entries or coefficients. Then the height of
, written
, is defined as
Until the end of the text we set
In section 5.1 we have described an algorithm for lifting
power series in a Kronecker parametrization. In the following paragraph
we adapt the method to lift -adic
integers instead. The algorithm originates from [27,
section 4.6]. The Jacobian matrix of
in
is still written
Algorithm
Initialize ,
,
,…,
.
Initialize with the inverse of
modulo
.
While do the following:
update and
, over
;
set ;
compute over
;
compute over
;
replace computed over
;
for from
to
,
replace computed over
.
For from
to
, compute
over
.
Return .
be a fixed rational value. If
then Algorithm 6.1 takes time , where
.
Proof. This algorithm directly comes from [27, section 4.6]. We do not repeat the proof of the correctness but focus directly on its bit complexity for the dense representation of input polynomials.
Before all we may compute the partial derivatives of the in time
by Lemma 2.9
to any precision
. In step 2,
we compute
in time
by Theorem 3.3. Then the inversion of
modulo
is performed by Berkowitz' algorithm in
time
.
In step 3 the evaluations of and
at
at a given precision
take time
,
again in view of Theorem 3.3. Since we keep doubling
until we reach the precision
, the total cost is bounded by
. The rest of the computations in steps 3 and 4
take time
.
Let be a zero dimensional subvariety of
. “The” Chow
form of
is the polynomial
that satisfies
where the are taken in
. The Chow form is defined up to a scalar factor in
, and is homogeneous of total
degree
.
From now we assume that is the variety of
, where
is
an invertible
integer matrix such that
is in Noether position. In this setting,
is included in the affine subspace
defined by
, and we have
. We then define the normalized Chow
form
of
as the
-multiple of
having the coefficient of
equal to
. The normalized Chow form belongs to
.
In order to bound bit sizes of Kronecker parametrizations of , we first need to bound the bit
size of the coefficients of
.
For this purpose we follow a construction due to Philippon [62],
and introduce
where denotes the
-Mahler
measure of
defined as
where is the usual Haar measure of mass 1 over
the sphere
So is to be understood as a real
-dimensional integral.
On the other hand, the usual Mahler measure of
is defined as an
-dimensional
complex contour integral
The relationship between these two measures is given in [52, Theorem 4]:
which implies
![]() |
(6.1) |
Then we appeal to [43, Corollary 2.10, p. 553] which
ensures the existence of a rational number such
that
has all its coefficients in
and
![]() |
(6.2) |
To relate the Mahler measure to the actual bit size of
we use [43, Lemma 1.1, p. 529]:
![]() |
(6.3) |
Combining the three latter inequalities leads to
Let be a Kronecker parametrization of
by the primitive element
taken in
. It relates to the Chow form
of
as follows:
Since belongs to
,
all the coefficients of
and of
for
also belong to
,
and their bit size can be bounded as follows. By
and inequality (2.1), the Chow form
admits at most
terms, whence
where is a shorthand for
. Now we estimate how the coefficients of
increase because of the change of variables
.
Proof. The proof is immediate by using and inequality (2.1).
Proof. This is a well known unconditional variant of Stirling's formula:
be an
matrix over
such
that
is in Noether position. Then we have
In addition, for a Kronecker parametrization
of
, and with
as above, then we have
Proof. From Lemmas 6.2 and 6.3 we obtain
Consequently,
Combining this bound with (6.4)–(6.7)
and using , we deduce
Finally we use to bound
by
, which concludes the
proof.
Given an input polynomial system satisfying
, we are led
to characterize prime numbers
for which the
system admits
isolated simple roots when
considered modulo
. We will
rely on the a posteriori verification criterion from
Proposition 4.6, in order to ensure that a resolution
modulo
can be lifted over the rational numbers.
Proof. The inequality is an equality for . For
we
have
whence by induction on
.
be a
matrix over the real numbers, then we have
, where
represents
the
-th column vector of
.
be homogeneous polynomials in
which satisfy
such that
and for any that does not divide
the sequence
satisfies
.
Proof. By Lemma 4.5 there exists an
invertible matrix
with
integer entries such that
and
is in Noether position. On the other hand by Lemma 4.2
there exists a primitive element
for
with integer coefficients and such that
. Let
represent the
corresponding Kronecker parametrization, and let
and
be as above. Let
be
a non-zero coefficient in
of a term of degree
for
,
let
be the Jacobian matrix of
with respect to
, and
By construction, ,
and thus
are positive integers.
Now if
does not divide
, then
for
, and the system
admits
isolated regular roots modulo
. Consequently
satisfy
R
, R
, and
by Proposition 4.6.
To conclude the proof it remains to bound .
First, we may further rewrite a bound for the quantity
introduced in Lemma 6.4, by making use of
:
By applying Lemma 6.6 on the Sylvester matrix of and
, we obtain
whence
As to , we deduce from Lemma
6.4 that
On the one hand, we have
hence
On the other hand, Lemma 6.5 gives us
From Lemma 6.4 and ,
it follows that
Let denote the group of permutations of
. Then we have
Again by Hadamard bound we deduce that
whence
The conclusion follows from adding right-hand sides of (6.8),
(6.9), and .
Remark
for
. Let
be an invertible
matrix over
and let
be a linear form in
. Let
be a prime number,
and let
be polynomials in
such that:
is monic of degree
and
for
;
is separable modulo
;
;
is a root of
modulo
;
The value of the Jacobian matrix of in
at
is invertible
modulo
.
Then satisfy
),
is a primitive element for
,
and
coincide modulo
with the Kronecker parametrization of
over
by
.
Proof. By Proposition 4.6, the residues
modulo of
form a
Kronecker parametrization of
.
We may use Algorithm 6.1 from a theoretical point of view
in order to prove the existence of univariate polynomials
over the
-adic
numbers
, satisfying the
following properties:
is monic of degree
and
for
;
have nonnegative valuations in
and coincide with
modulo
;
;
is a root of
modulo
;
The value of the Jacobian matrix of in
at
is invertible modulo
.
Consequently, Proposition 4.6 ensures that
is a Kronecker parametrization of
over
by the primitive element
.
By uniqueness, it coincides with the image over
of the Kronecker parametrization of
over
by
. Using
again Proposition 4.6 with the parametrization over the
rationals concludes the proof.
Now we address the question of obtaining a suitable prime number.
and
such that
, computes a
prime
in the range
, which does not divide
, with probability of success
, and in time
.
Proof. This is a consequence of [21,
Theorem 18.10, part (i)]: we just appeal to the deterministic primality
test designed in [1] in order to ensure that is always prime in return. This test works in polynomial
time.
The idea behind the Kronecker solver over is as
follows. For a suitable prime
,
we use the Kronecker algorithm to solve the input system over
. If the algorithm finishes with
success, then we appeal to the a posteriori verification
criterion of Lemma 6.9. If the criterion passes, then we
may lift the Kronecker parametrization over the
-adic integers to arbitrary precision
by using Algorithm 6.1, and we know that the
-adic parametrization is the
image of the rational one. Consequently, as soon as
is sufficiently large, we can reconstruct the parametrization over
.
Let us recall that a rational number ,
with
and
,
is uniquely determined by its
-adic
expansion to precision
whenever
and
; see for instance [68, section 4]. Using the fast Euclidean algorithm from [21, Theorem 5.26] this reconstruction takes
. The choices of
and
the precision
needed by the solver are made
precise in the following algorithm. We recall that
is a fixed positive rational number, thought to be small.
Algorithm
For each from
to
, compute the maximum
bit size
of the coefficients of
.
Compute and a positive integer
such that
.
Compute such that
Compute a prime number at random in
via Lemma 6.10.
Call the Kronecker algorithm underlying Theorem 5.4
with the images of
in
. If the
resolution fails then go to step 2.
Let be the data returned by the
Kronecker algorithm in step 3. If
,
or if
is not separable, or if
, or if one of the
does not vanish at
modulo
, or if the
value of the Jacobian matrix of
in
at
is not invertible
modulo
, then go to
step 2.
Let (resp.
)
be the canonical preimage of
(resp. of
) with entries in
. Let
be the first integer such that
Compute over
, and use Algorithm 6.1 to
obtain the Kronecker parametrization
by
of
over
.
By rational reconstruction, compute the unique polynomials in
that coincide with
modulo
.
Notice that is
whenever
all the
have their coefficients in
. Therefore in the following complexity
estimates we shall write
in order to make
complexity bounds valid in this particular case.
Proof. By Lemma 6.10 the prime determined in step 2 does not divide the integer
defined in Lemma 6.7 with probability
. In other words, the
sequence
satisfy properties
. Thanks to
, Theorem 5.4 ensures that the
Kronecker solver modulo
succeeds with
probability
. It follows that
the expected number of times that the algorithm repeats steps 2 to 4 is
bounded.
Once the algorithm passes step 4, Lemma 6.9 ensures that
the parametrization modulo may be lifted to any
arbitrary precision, and that the parametrization over
may be recovered whenever the precision is sufficiently large. According
to the preceding discussion on rational reconstruction, we need that
with defined in Lemma 6.4; it
satisfies
with and
being
. This concludes the proof of
correctness.
The computation of needs time
by Lemmas 2.6 and 2.8. Then
,
and
can be computed in time
,
which is negligible since
The cost of step 2 is by Lemma 6.10,
which is again negligible.
By Theorem 5.4 step 3 takes time
since . Proposition 4.6
then provides us with the cost
for step 4. Computing over
in step 5 takes
by Proposition 2.15.
The cost of the
-adic Hensel
lifting is given in Proposition 6.1:
where . We verify that
Step 6 requires additional time .
The version of the Kronecker solver summarized in Algorithm 6.2
can be optimized by using ideas previously developed for the case of
input polynomials given by straight-line programs [27]; see
implementation details in the
A first optimization consists in minimizing the height of the
parametrization over the rationals. For this purpose it is worth finding
a matrix and a primitive element of small
height. This search may be achieved efficiently modulo
, and we refer the reader to [50]
for the description of the underlying algorithms.
A second optimization concerns the early termination of the lifting
stage over the -adic
integers. In fact one may try to perform rational reconstruction before
the bound on the precision is actually reached. If all the
reconstructions succeed, then it suffices to verify that the Kronecker
parametrization actually describes the solution set. A probabilistic
method to do this consists in evaluating the equations at the
parametrization modulo an independent random prime number
with a bit size similar to the one of
. This verification may also be done
deterministically over the rationals, but at a higher cost.
For an input system in variables of degrees
, the quantity
is of the order
.
Therefore with
, Algorithm 6.2 involves primes
of bit size close
to 52 bits. In fact, the bounds used in the design of Algorithm 6.2
have been simplified and overestimated for the sake of the presentation.
Sharper bounds should be considered in practice, in order to ensure that
fits into 32 bits for small and moderate
systems, and 64 bits for larger ones.
Let us finally mention that the present bound for the bit size of the Kronecker parametrization over the rationals turns out to be essentially sharp for generic systems, according to [54, Theorem 3.1]. Consequently the complexity reached within Theorem 6.11 is nearly linear in the “expected” size of the output, from the asymptotic point of view. This fact is illustrated by the following example.
Example
where and
.
Let
be a root of
and set
so that is clearly a zero of
. For
,
we observe that
It follows that is also a zero of
. The value of the Jacobian matrix of
in
at
is
By expanding with respect to the last column, we
obtain
Consequently is invertible. From
Proposition 4.6 thus ensures that
satisfies R
and R
,
that it is in Noether position, and that it is parametrized by the
primitive element
, as
follows:
The Kronecker solver was already known to be competitive both in theory and practice for polynomial systems represented by straight-line programs; see for instance timings in [2, 27]. Under suitable genericity assumptions and for suitable ground fields, we have shown in this paper that this solver also leads to improved probabilistic asymptotic complexity bounds for the dense representation of input polynomials.
Our main results concern finite fields and rational numbers, but our methods can in principle be extended to any ground field over which fast multivariate multi-modular composition is available. Unfortunately, it is still a major open problem to design an algorithm for modular composition over arbitrary ground fields.
Another major problem concerns the practical efficiency of algorithms for modular composition. This holds in particular for the recent fast algorithms for finite fields designed by Kedlaya and Umans [42], and revisited in [36]; we refer the reader to the concluding section of [36] for a quantitative discussion. Consequently we do not expect our Theorem 3.3 to be of practical use, and we cannot claim our new main complexity bounds for polynomial system solving to be relevant in practice.
Nevertheless, it is important to point out that Theorem 3.3 uses modular composition as a blackbox. Whenever faster algorithms for this task do become available, they can directly be used in our algorithms. As one possible research direction, it is worthwhile to investigate the replacement of generic modular compositions by specific ones, such as those studied in [33, 34] and which feature more promising performances.
The Kronecker solver admits different extensions for when genericity
assumptions
Yet another important question for systems over the rational numbers concerns the complexity of finding numerical approximations of the roots. In generic cases we have proved in Theorem 6.11 how to compute Kronecker representations of solution sets in quasi-optimal time. This implies that numerical roots may also be computed in quasi-linear time but whenever the requested precision is “sufficiently large”, by means of fast univariate polynomial solvers. It would be interesting to quantify the latter “sufficiently large”, and to compare to the state of the art of numerical polynomial solvers.
This appendix is devoted to subjecting a multivariate polynomial to a linear change of variables. More precisely,
given
and an
matrix
over a commutative ring
,
then we wish to compute
The fast algorithms that we propose below do not seem to be available in the literature. They are suitable for any coefficient ring with sufficiently many elements, and they are also well suited for homogeneous polynomials.
In this subsection we focus on the algebraic model (computation trees
for instance), we let be an effective
commutative ring, and
is a cost function such
that two polynomials in
may be multiplied with
cost
. The evaluation of a
multivariate polynomial at points in a block of points
, where
is a finite
subset of
, is usually
achieved by the successive use of fast univariate evaluations, as
recalled in the following lemma.
, let
be of partial degree
in
for
,
and let
be a subset of
of cardinality
. Then, all
the values of
at
can
be computed with
arithmetic operations in
.
Proof. We interpret as a
univariate polynomial in
,
We evaluate at
recursively. Then, for each
,
we evaluate
at all the points of
, with a total cost
. Denoting by
the cost of
the algorithm in terms of operations in
,
we thus obtain
By induction over , it
follows that
which implies the claimed bound.
The next lemma, also well known, concerns the corresponding interpolation problem.
, let
be pairwise distinct points in
such that
is invertible whenever
, let
be a family of
values in
for
running
over
. The unique
polynomial
of partial degrees
and such that
for all
can be computed with
arithmetic operations in
, including inversions.
Proof. Again we interpret as
a univariate polynomial in
,
![]() |
(A.1) |
For all we interpolate the values
with
operations in
. We then recursively interpolate
and form
as in (A.1).
The total cost is obtained as in the proof of the previous lemma.
The aim of the following proposition is the fast evaluation of at a set of points of the form
, for any matrix
and any
vector
.
,
let
be of partial degree
in
for
,
let
be a subset of
of
cardinality
such that
is invertible whenever
,
let
be a
matrix over
, and let
. Let
be the column
vector with entries
. If an
LU-decomposition of
is given, then
and
can be computed with
arithmetic operations in
, including inversions.
Proof. We write .
We first assume that
is upper triangular, and we
partition
into
where and
.
We compute
for using
operations in
. For
, we then evaluate
at
by induction. The base case
takes constant time
.
Consequently, for any
, the
total number of operations in
is
, by the same argument as in the proof of Lemma
A.1. We recover
with
operations in
by Lemma A.2.
If is lower triangular then we may revert of the
variables in
and the columns of
in order to reduce to the upper triangular case. Alternatively, we may
adapt the latter decomposition of the set of points, as follows:
where and
.
So we compute
and evaluate at
by
induction, for
.
Finally if is general, then it suffices to use
the given LU-decomposition, where
is lower
triangular with
on the diagonal, and
is upper triangular. In fact we have
, so we compute
and
then
and
.
In the next lemma the same technique is adapted to homogeneous polynomials.
be homogeneous of degree
, let
be
a
matrix over
,
and let
be a subset of
of cardinality
such that
is invertible whenever
. If
an LU-decomposition of
is given, then
can be computed with
arithmetic operations in
.
Proof. Assume first that is
lower triangular and let
. We
are led to compose
with
by means of Proposition A.3. If is
upper triangular then it suffices to revert the variables
in
, and the
columns of
, in order to
reduce to the lower triangular case. Alternatively, we may set
and compose
with
in order to obtain . Finally,
for any
, it suffices to use
the given LU-decomposition.
be homogeneous of
degree
, let
be an
matrix over
, and let
be a subset
of
of cardinality
such
that
is invertible whenever
. If an LU-decomposition of
is given, then
can be computed with
arithmetic operations in
.
Proof. The total number of coefficients in is
by inequality (2.1).
We decompose
![]() |
(A.2) |
where is made of the terms of
which are multiple of
, then
is made of the terms of
which are multiple of
,…,
and finally
is made of the terms of
which are multiple of
(that is a
-multiple of a power of
). In this way, we are led to
compute
for
,
with
of degree
;
this requires
operations in
, by Lemma A.4. Then
can be recovered with further
operations.
Remark , for instance in geometric progressions, then
multipoint evaluations and interpolations in one variable and in degree
over
cost
by means of [10], that saves a factor of
in the above complexity estimates.
For the purpose of the present paper, we need to adapt the results of
the previous subsection to the case when is the
Galois Ring
, and in the
context of Turing machines. In the next lemmas we use the lexicographic
order on
, written
, defined by
In terms of Turing machines, we need the following variants of Lemmas A.1 and A.2.
,
let
be of partial degree
in
for
,
and let
be values in
. Then, the values
for
running over
in the
lexicographic order
can be computed in
time
Proof. The proof follows the one of Lemma A.1
while taking data reorganizations into account. More precisely, using
one matrix transposition, we reorganize the
values of the
after the recursive calls into the
sequence of
for running over
in the
lexicographic order
. Then,
after the multi-point evaluations of
,
we need to transpose the
array made of the
values of
, in order to
ensure the lexicographic ordering in the output. The cost of these
transpositions is
by Lemma 2.1,
which is negligible.
and
. Let
be pairwise distinct
values in
such that
is
invertible modulo
for all
, and let
be a family
of values in
for
running over
in the lexicographic order
. The unique polynomial
of partial degree
in
for
, and
such that
for all
in
, can be computed in
time
Proof. The proof follows the one of Lemma A.2,
by doing the data reorganizations in the opposite direction from the one
in the proof of Lemma A.7.
From now, for convenience, we discard the case . In this way, whenever
, we may use
.
and
. Let
be of partial
degree
in
for
, and let
be a
matrix over
.
If an LU-decomposition of
is given, then
can be computed in time
.
Proof. We first generate a subset
of
of cardinality
in
time
; this ensures the
invertibility of
for
. The proof then follows the one of Proposition A.3 while taking data reorganizations into account. When
is upper triangular, the computation of
requires the multi-point evaluation of
regarded in
: we may simply
appeal to the fast univariate algorithm because it only involves
additions, subtractions and products by elements of
over the ground ring
.
Consequently
may be obtained in time
, by Lemma 2.5. In
addition, the
array of values of the
must be transposed at the end, in order to guarantee the
lexicographic ordering necessary to interpolate
.
When is lower triangular, the data
reorganization costs essentially the same, except that the computation
of
takes time
by Lemmas
2.11 and 2.5.
Before achieving the proof of Proposition 2.15, we further need the following lemma in order to change the representation of a homogeneous polynomial.
be a homogeneous polynomial of
degree
in
,
represented as before by
and
, and let
.
Then, for any
we can compute
in time
.
Proof. For simplicity the proof is done for , but it extends in a
coefficientwise manner to any
.
A sparse representation of
is made of a sequence
of pairs of coefficients and vector exponents. More precisely, if
then a sparse representation of it is the sequence of
the pairs
, for all the
non-zero coefficients
. The
bit size of a vector exponent is
,
and therefore the bit size of a sparse representation of
is
by (2.1).
In order to prove the lemma, we first convert , given in dense representation, into a sparse
representation. When
the sparse representation
of
may be obtained in time
. Otherwise
and we
regard
in
,
and recursively compute the sparse representation of
for
. These representations
may naturally be glued together into a sparse representation of
, in time
, by adding the exponent of
into each exponent vector. A straightforward induction leads to a total
time
for the change of representation of
. Then the sparse representation of
may be deduced with additional time
by appending the exponent of
needed for homogenization.
Second, from the latter sparse representation of
we may simply discard the exponents of
and
multiply the coefficients with the corresponding powers of
, in order to obtain a sparse representation of
in time
.
Finally it remains to construct the dense representation of from its sparse representation. To this aim we sort the
sparse representation in increasing lexicographic order on the exponent
vectors in time
. We next
compute the dense representation by induction over
. Writing
the sparse representations of are computed by
induction, after removal of the powers of
.
The induction ends when
, in
which case the conversion to dense representation requires time
. In total, the dense
representation of
can be computed in time
.
Proof of Proposition 2.15. We follow the proofs of Lemma A.4 and Proposition A.5, still while taking into account the cost of data reorganizations.
In the proof of Lemma A.4, the cost of obtaining and
is given by Lemma A.10,
that is
.
In the proof of Proposition A.5 we first need to compute
the decomposition (A.2) of .
The polynomial
is represented by
and . Consequently
may be easily obtained in time
. Then the rest of the decomposition
is obtained from
,
recursively. The total cost for obtaining all the
is therefore bounded by
.
For any , any
, and any
,
the computations of
and of
take time
since their supports have cardinality
by (2.1).
Finally, from
we obtain the representation of as
using additional time . The
cost of the data reorganizations in the proof of Proposition A.5
is negligible.
M. Agrawal, N. Kayal, and N. Saxena. PRIMES is in P. Ann. Math., pages 781–793, 2004.
B. Bank, M. Giusti, J. Heintz, G. Lecerf, G. Matera, and P. Solernó. Degeneracy loci and polynomial equation solving. Found. Comput. Math., 15(1):159–184, 2015.
M. Bardet. Étude des systèmes algébriques surdéterminés. Applications aux codes correcteurs et à la cryptographie. PhD thesis, Université Pierre et Marie Curie - Paris VI, 2004. https://tel.archives-ouvertes.fr/tel-00449609.
M. Bardet, J.-C. Faugère, and B. Salvy. On the
complexity of the Gröbner basis
algorithm. J. Symbolic Comput., 70:49–70, 2015.
S. J. Berkowitz. On computing the determinant in small parallel time using a small number of processors. Inform. Process. Lett., 18:147–150, 1984.
J. Berthomieu, J. van der Hoeven, and G. Lecerf. Relaxed algorithms for p-adic numbers. J. Théor. Nombres Bordeaux, 23(3), 2011.
J. Berthomieu, G. Lecerf, and G. Quintin. Polynomial root finding over local rings and application to error correcting codes. Appl. Alg. Eng. Comm. Comp., 24(6):413–443, 2013.
A. Bostan, F. Chyzak, M. Giusti, R. Lebreton, G. Lecerf, B. Salvy, and É Schost. Algorithmes Efficaces en Calcul Formel. Frédéric Chyzak (self-published), Palaiseau, 2017. Electronic version available from https://hal.archives-ouvertes.fr/AECF.
A. Bostan, Ph. Flajolet, B. Salvy, and É. Schost. Fast computation of special resultants. J. Symbolic Comput., 41(1):1–29, 2006.
A. Bostan and É. Schost. Polynomial evaluation and interpolation on special sets of points. J. Complexity, 21(4):420–446, 2005.
R. P. Brent and H. T. Kung. Fast algorithms for manipulating formal power series. J. ACM, 25(4):581–595, 1978.
W. D. Brownawell. Bounds for the degrees in the Nullstellensatz. Annal. of Math., 126(3):577–591, 1987.
P. Bürgisser, M. Clausen, and M. A. Shokrollahi. Algebraic complexity theory, volume 315 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, 1997.
J. F. Canny, E. Kaltofen, and L. Yagati. Solving systems of nonlinear polynomial equations faster. In Proceedings of the ACM-SIGSAM 1989 International Symposium on Symbolic and Algebraic Computation, ISSAC '89, pages 121–128, New York, NY, USA, 1989. ACM.
D. G. Cantor and E. Kaltofen. On fast multiplication of polynomials over arbitrary algebras. Acta Infor., 28:693–701, 1991.
J.-M. Couveignes and R. Lercier. Fast construction of irreducible polynomials over finite fields. Israel J. Math., 194(1):77–105, 2013.
C. D'Andrea, A. Ostafe, I. E. Shparlinski, and M. Sombra. Reduction modulo primes of systems of polynomial equations and algebraic dynamical systems. Trans. Amer. Math. Soc., 371(2):1169–1198, 2019.
C. Durvye and G. Lecerf. A concise proof of the Kronecker polynomial system solver from scratch. Expo. Math., 26(2):101–139, 2008.
J.-C. Faugère, P. Gaudry, L. Huot, and G. Renault. Sub-cubic change of ordering for Gröbner basis: A probabilistic approach. In Proceedings of the 39th International Symposium on Symbolic and Algebraic Computation, ISSAC '14, pages 170–177, New York, NY, USA, 2014. ACM.
J.-C. Faugère, P. Gianni, D. Lazard, and T. Mora. Efficient computation of zero-dimensional Gröbner bases by change of ordering. J. Symbolic Comput., 16(4):329–344, 1993.
J. von zur Gathen and J. Gerhard. Modern computer algebra. Cambridge University Press, New York, 3rd edition, 2013.
N. Giménez and G. Matera. On the bit complexity of polynomial system solving. J. Complexity, 51:20–67, 2019.
M. Giusti. Some effectivity problems in polynomial ideal theory. In J. Fitch, editor, EUROSAM 84: International Symposium on Symbolic and Algebraic Computation Cambridge, England, July 9–11, 1984, pages 159–171, Berlin, Heidelberg, 1984. Springer Berlin Heidelberg.
M. Giusti, K. Hägele, J. Heintz, J. L. Montaña, J. E. Morais, and L. M. Pardo. Lower bounds for Diophantine approximations. J. Pure Appl. Algebra, 117/118:277–317, 1997.
M. Giusti, J. Heintz, J. E. Morais, J. Morgenstern, and L. M. Pardo. Straight-line programs in geometric elimination theory. J. Pure Appl. Algebra, 124(1-3):101–146, 1998.
M. Giusti, J. Heintz, J. E. Morais, and L. M. Pardo. When polynomial equation systems can be “solved” fast? In Applied algebra, algebraic algorithms and error-correcting codes (Paris, 1995), volume 948 of Lecture Notes in Comput. Sci., pages 205–231. Springer-Verlag, 1995.
M. Giusti, G. Lecerf, and B. Salvy. A Gröbner free alternative for polynomial system solving. J. complexity, 17(1):154–211, 2001.
B. Grenet, J. van der Hoeven, and G. Lecerf. Deterministic root finding over finite fields using Graeffe transforms. Appl. Alg. Eng. Comm. Comp., 27(3):237–257, 2016.
D. Harvey and J. van der Hoeven. Faster polynomial multiplication over finite fields using cyclotomic coefficient rings. J. Complexity, 54:101404, 2019.
D. Harvey and J. van der Hoeven. Integer
multiplication in time .
Technical report, HAL, 2019. http://hal.archives-ouvertes.fr/hal-02070778.
D. Harvey and J. van der Hoeven. Polynomial
multiplication over finite fields in time . Technical report, HAL, 2019. http://hal.archives-ouvertes.fr/hal-02070816.
J. Heintz. Definability and fast quantifier elimination in algebraically closed fields. Theor. Comput. Sci., 24(3):239–277, 1983.
J. van der Hoeven and G. Lecerf. Modular composition via complex roots. Technical report, CNRS & École polytechnique, 2017. http://hal.archives-ouvertes.fr/hal-01455731.
J. van der Hoeven and G. Lecerf. Modular composition via factorization. J. Complexity, 48:36–68, 2018.
J. van der Hoeven and G. Lecerf. Accelerated tower arithmetic. J. Complexity, 55:101402, 2019.
J. van der Hoeven and G. Lecerf. Fast multivariate multi-point evaluation revisited. J. Complexity, 56:101405, 2020.
J. van der Hoeven, G. Lecerf, B. Mourrain, et al. Mathemagix, From 2002. http://www.mathemagix.org.
Xiaohan Huang and V. Y. Pan. Fast rectangular matrix multiplication and applications. J. Complexity, 14(2):257–299, 1998.
G. Jeronimo and J. Sabia. Effective equidimensional decomposition of affine varieties. J. Pure Appl. Algebra, 169(2–3):229–248, 2002.
E. Kaltofen and V. Shoup. Fast polynomial factorization over high algebraic extensions of finite fields. In Proceedings of the 1997 International Symposium on Symbolic and Algebraic Computation, ISSAC '97, pages 184–188, New York, NY, USA, 1997. ACM.
K. S. Kedlaya and C. Umans. Fast modular composition in any characteristic. In FOCS'08: IEEE Conference on Foundations of Computer Science, pages 146–155, Washington, DC, USA, 2008. IEEE Computer Society.
K. S. Kedlaya and C. Umans. Fast polynomial factorization and modular composition. SIAM J. Comput., 40(6):1767–1802, 2011.
T. Krick, L. M. Pardo, and M. Sombra. Sharp estimates for the arithmetic Nullstellensatz. Duke Math. J., 109(3):521–598, 2001.
L. Kronecker. Grundzüge einer arithmetischen Theorie der algebraischen Grössen. J.reine angew. Math., 92:1–122, 1882.
Y. N. Lakshman. On the complexity of computing a Gröbner basis for the radical of a zero dimensional ideal. In Proceedings of the Twenty-second Annual ACM Symposium on Theory of Computing, STOC '90, pages 555–563, New York, NY, USA, 1990. ACM.
Y. N. Lakshman. A single exponential bound on the complexity of computing Gröbner bases of zero dimensional ideals. In T. Mora and C. Traverso, editors, Effective Methods in Algebraic Geometry, pages 227–234, Boston, MA, 1991. Birkhäuser Boston.
Y. N. Lakshman and D. Lazard. On the complexity of zero-dimensional algebraic systems. In T. Mora and C. Traverso, editors, Effective Methods in Algebraic Geometry, pages 217–225, Boston, MA, 1991. Birkhäuser Boston.
D. Lazard. Gröbner bases, Gaussian elimination and resolution of systems of algebraic equations. In J. A. Hulzen, editor, Computer Algebra: EUROCAL'83, European Computer Algebra Conference London, England, March 28–30, 1983 Proceedings, pages 146–156. Springer Berlin Heidelberg, 1983.
F. Le Gall. Powers of tensors and fast matrix multiplication. In K. Nabeshima, editor, ISSAC'14: International Symposium on Symbolic and Algebraic Computation, pages 296–303, New York, NY, USA, 2014. ACM.
G. Lecerf. Computing the equidimensional decomposition of an algebraic closed set by means of lifting fibers. J. Complexity, 19(4):564–596, 2003.
G. Lecerf. On the complexity of the Lickteig–Roy subresultant algorithm. J. Symbolic Comput., 92:243–268, 2019.
P. Lelong. Mesure de Mahler et calcul de constantes
universelles pour les polynomes de
variables. Math. Ann., 299(1):673–695, 1994.
H. Matsumura. Commutative ring theory, volume 8 of Cambridge Studies in Advanced Mathematics. Cambridge university press, 1989.
D. McKinnon. An arithmetic analogue of Bezout's theorem. Compos. Math., 126(2):147–155, 2001.
J. M. McNamee and V. Y. Pan. Numerical Methods for Roots of Polynomials, Part II, volume 16 of Studies in Computational Mathematics. Elsevier, 2013.
B. Mourrain, V. Y. Pan, and O. Ruatta. Accelerated solution of multivariate polynomial systems of equations. SIAM J. Comput., 32(2):435–454, 2003.
B. Mourrain and Ph. Trébuchet. Solving projective complete intersection faster. In Proceedings of the 2000 International Symposium on Symbolic and Algebraic Computation, ISSAC '00, pages 234–241, New York, NY, USA, 2000. ACM.
B. Mourrain and Ph. Trébuchet. Generalized normal forms and polynomial system solving. In Proceedings of the 2005 International Symposium on Symbolic and Algebraic Computation, ISSAC '05, pages 253–260, New York, NY, USA, 2005. ACM.
B. Mourrain and Ph. Trébuchet. Border basis representation of a general quotient algebra. In Proceedings of the 37th International Symposium on Symbolic and Algebraic Computation, ISSAC '12, pages 265–272, New York, NY, USA, 2012. ACM.
A. K. Narayanan. Fast computation of isomorphisms between finite fields using elliptic curves. In L. Budaghyan and F. Rodríguez-Henríquez, editors, Arithmetic of Finite Fields. 7th International Workshop, WAIFI 2018, Bergen, Norway, June 14–16, 2018, Revised Selected Papers, volume 11321 of Lecture Notes in Comput. Sci., pages 74–91. Springer, Cham, 2018.
C. H. Papadimitriou. Computational Complexity. Addison-Wesley, 1994.
P. Philippon. Sur des hauteurs alternatives. I. Math. Ann., 289(1):255–283, 1991.
A. Poteaux and É. Schost. On the complexity of computing with zero-dimensional triangular sets. J. Symbolic Comput., 50:110–138, 2013.
A. Schönhage. Schnelle Berechnung von Kettenbruchentwicklungen. Acta Informatica, 1(2):139–144, 1971.
A. Schönhage, A. F. W. Grotefeld, and E. Vetter. Fast algorithms: A multitape Turing machine implementation. B. I. Wissenschaftsverlag, Mannheim, 1994.
J. T. Schwartz. Fast probabilistic algorithms for verification of polynomial identities. J. ACM, 27(4):701–717, 1980.
V. Shoup. New algorithms for finding irreducible polynomials over finite fields. Math. Comp., 54(189):435–447, 1990.
P. S. Wang. A p-adic algorithm for univariate partial fractions. In Proceedings of the Fourth ACM Symposium on Symbolic and Algebraic Computation, SYMSAC '81, pages 212–217, New York, NY, USA, 1981. ACM.
R. Zippel. Probabilistic algorithms for sparse polynomials. In Proceedings EUROSAM' 79, number 72 in Lect. Notes Comput. Sci., pages 216–226. Springer-Verlag, 1979.