|
. This work has
been supported by the ANR-09-JCJC-0098-01
The technique of relaxed power series expansion provides an
efficient way to solve so called recursive equations of the form
|
Let be an effective field of constants of
characteristic zero. This means that elements in
can be encoded by data structures on a computer and that we have
algorithms for performing the field operations of
.
Let be a column vector of
indeterminate series in
. We
may also consider
as a power series
. Let
be a column
vector of expressions built up from
,
and constants in
using
ring operations, differentiation and integration (with constant term
zero). Finally, let
be a finite number of
initial conditions. Assume that the system
![]() |
(1) |
admits a unique solution . In
this paper, we are interested in the efficient computation of this
solution up to a given order
.
In the most favourable case, the equation is of
the form
where the coefficient of
in
only depends on earlier coefficients
of
, for
each
. In that case,
actually provides us with a recurrence relation for the computation of
the solution. Using the technique of relaxed power series expansions [vdH02, vdH07a], which will briefly be recalled in
section 2.4, it is then possible to compute the expansion
up till order
in time
where is the number of multiplications occurring
in
,
is the total size of
as an expression, and
denotes the complexity of relaxed multiplication of
two power series up till order
.
Here we assume that
is represented by a directed
acyclic graph, with possible common subexpressions. For large
, we have
, where
denotes the
complexity [CT65, SS71, CK91] of
multiplying two polynomials of degrees
.
If
admits sufficiently many
-th roots of unity, then we even have
and
. For
moderate
, when polynomial
multiplication is done naively or using Karatsuba's method, relaxed
multiplication is as efficient as the truncated multiplication of
polynomials at order
.
One particularly important example of an equation of the above type is the integration of a dynamical system
where is algebraic (i.e. does not
involve differentiation or integration). In that case, given the
solution
up till order
, we may consider the linearized system
up till order , where
stands for the Jacobian matrix associated to
at
. If we
have a fundamental system of solutions of
up
till order
, then one step of
Newton's method allows us to find the solution of (4) and a
new fundamental system of solutions of the linearized equation up till
order
[BK78, BCO+06].
A careful analysis shows that this leads to an algorithm of time
complexity
In [vdH10], this bound has been further improved to
under the assumptions that admits sufficiently
many
-th roots of unity and
that
.
Although the complexity (5) is asymptotically better than
(3) for very large ,
the relaxed approach often turns out to be more efficient in practice.
Indeed, Newton's method both suffers from a larger constant factor and
the fact that we profit less from the potential sparsity of the system.
In particular, if
, then the
relaxed approach is generally faster. Moreover, as long as
multiplications are done in the naive or Karatsuba model, the relaxed
approach is optimal in the sense that the computation of the solution
takes roughly the same time as its verification. Another advantage of
the relaxed approach is that it generalizes to more general functional
equations and partial differential equations.
Let us now return to our original implicit system (1). If
is a system of differentially algebraic
equations, then we may also seek to apply Newton's method. For non
degenerate systems and assuming that we have computed the solution
and a fundamental system of solutions for the
linearized equation up till order
,
one step of Newton's method yields an extension of the solutions up till
order
, for a fixed constant
. From an asymptotic point of
view, this means that the complexities (5) and (6)
remain valid, modulo multiplication of
by the
differential order of the system in these bounds.
Another approach for the resolution of (1) is to keep
differentiating the system with respect to until
it becomes equivalent to a system of the form (2). For
instance, if
is algebraic, then differentiation
of (1) yields
Consequently, if is invertible, then
provides us with an equivalent system which can be solved by one of the previous methods. Unfortunately, this method requires the computation of the Jacobian, so we do not longer exploit the potential sparsity of the original system.
Yet another recent approach [vdH09b] is to consider not yet
computed coefficients of as formal unknowns, and
solve the system of equations
for increasing
values of
. For large
, the system
usually reduces to a linear system of equations. In particular, the
coefficients of series with unknown coefficients are not polynomials but
merely linear combinations. Using the so called “substitution
product”, the multiplication of series with unknown coefficients
can be done while taking advantage of this linearity.
In this paper, we will present a variant of the approach of [vdH09b]. Roughly speaking, we reconsider the series with unknown coefficients as vectors of partially unknown series. Technically speaking, this is done via the concept of anticipators, which will be introduced in section 3. Using this technique, and under mild assumptions, we show in section 4.1 how to rewrite the original system of equations into a new system which is both recursive and not much larger in size. We may then apply a standard relaxed algorithm for its resolution. This leads to slightly sharper complexity bounds than those from [vdH09b], which will be presented in section 4.2.
The main interest of the new technique though is the fact that it allows for the direct resolution of implicit equations by existing software for recursive equations. Moreover, recursive equations naturally occur in the area of reliable computing [Moo66, AH83, Neu90, MKC09, Rum10]. In this setting, our power series are replaced by so called Taylor models [MB96, MB04], which systematically require certified bounds for the remainders on polydisks. In section 5, we will show how our results apply in this case. In particular, the new algorithms are an asymptotically efficient device for the certified resolution of implicit equations and integration of dynamical systems on implicitly defined varieties.
Remark -adic numbers
instead of power series [BL11]. They also implemented their
ideas in the
Assume that we have fixed a set of function
symbols, together with an arity
for
each
. Then a dag
over
is a rooted finite directed acyclic
-labeled graph, such that each
-labeled node has
successors, together with an ordering on the successors.
For instance,
is a typical dag for the expression ,
with
,
and
. We will denote by
the number of nodes of a dag
(also called its size) and by
its number of
multiplications (also called its multiplicative size). For our example
dag
, we thus have
and
. We will
denote by
the set of dags over
.
More generally, we may consider multivariate dags with an arbitrary number of roots, which again come with ordering. For instance,
is a bivariate dag which represents a vector of two expressions and
. We
will denote by
the number of roots of a
multivariate dag
, which we
will also call its dimension. We will write
, where
stands for the
subdag whose root is the
-th
root of
. We will denote by
the set of multivariate dags over
of dimension
and
.
Consider a linear operator .
We say that
is a coefficientwise
operator, if there exist fixed constants
such
that
for all . For every
, the operator
defined by
is an example of a coefficientwise operator. The truncation operators
, which are defined by
constitute another family of examples. We will denote by and
the sets of all operators of
the form
with
resp.
with
. Finally, we define the coefficientwise operators
and
by
and we notice that is the inverse of
on
.
Let be
“indeterminate series” in
.
We will sometimes consider
as a series with
formal coefficients
Let be a set of coefficientwise linear
operators. In what follows, we will take
and denote by the set of dags over
. Similarly, we set
and
. Dags in
,
and
will respectively be called algebraic, differential
and integral. Notice that polynomials in
are regarded as dags of size
,
independently of their degree; this is motivated by the fact that
coefficient extraction is trivial for explicit polynomials.
Clearly, any dag can be considered as a function
. Given a small symbolic
perturbation
, we may expand
as a Taylor series in
and truncation at order yields
We claim that can be regarded as a dag in
. For instance, if
then
In general, the claim is easily shown by induction over .
We claim that any dag can be regarded as a
series in
such that each coefficient is a dag over
Indeed, by induction over the size of ,
we first define the valuation
of
by
We next define the coefficients by another
induction over the size of
.
If
, then we take
. Otherwise, we take
As a result of the claim, we emphasize that only
depends on the coefficients
up till order
of
.
Remark can be replaced by any algebraically equivalent
formula, as long as
only depends on
and
. Assuming
the concept of relaxed power series, to be introduced below, this means
that we compute
using the formula
where is computed using a relaxed multiplication
algorithm.
Let us briefly recall the technique of relaxed power series
computations, which is explained in more detail in [vdH02].
In this computational model, a power series is
regarded as a stream of coefficients
.
When performing an operation
on power series it
is required that the coefficient
of the result
is output as soon as sufficiently many coefficients of the inputs are
known, so that the computation of
does not
depend on the further coefficients. For instance, in the case of a
multiplication
, we require
that
is output as soon as
and
are known. In particular, we may use the
naive formula
for the computation of
.
The additional constraint on the time when coefficients should be output
admits the important advantage that the inputs may depend on the output,
provided that we add a small delay. For instance, the exponential of a power series
may be
computed in a relaxed way using the formula
Indeed, when using the naive formula for products, the coefficient is given by
and the right-hand side only depends on the previously computed
coefficients .
The main drawback of the relaxed approach is that we cannot directly use
fast algorithms on polynomials for computations with power series. For
instance, assuming that has sufficiently many
-th roots of unity and that
field operations in
can be done in time
, two polynomials of degrees
can be multiplied in time
, using FFT multiplication [CT65].
Given the truncations
and
at order
of power series
, we may thus compute the truncated product
in time
as well. This is much
faster than the naive
relaxed multiplication
algorithm for the computation of
.
However, the formula for
when using FFT
multiplication depends on all input coefficients
and
, so the fast algorithm
is not relaxed. Fortunately, efficient relaxed multiplication algorithms
do exist:
be the time complexity for the
multiplication of polynomials of degrees
in
. Then there exists a
relaxed multiplication algorithm for series in
of time complexity
.
admits a primitive
-th root of unity for all
, then there exists a relaxed multiplication
algorithm of time complexity
.
In practice, the existence of a
-th
root of unity with
suffices for multiplication
up to order
.
In what follows, we will denote by the
complexity of relaxed multiplication up till order
. Let us now consider a general equation of the
form
where an
-dimensional
dag. We say that (7) is a recursive equation, if
each coefficient
only depends on earlier
coefficients
of
.
That is,
for all
.
In order to solve (7) up till order
, we then need to perform
relaxed multiplications at order
and
coefficientwise operations
or
at order
.
This yields the following complexity bound:
When solving an implicit equation in using a
relaxed algorithm, the coefficients
are computed
only gradually. During the resolution process, it might happen that we
wish to evaluate dags at higher orders than the number of known
coefficients of
. That is,
given
and
,
we might need
, even though
only
are known. In that case, we have a problem,
but we may still do the best we can, and compute
instead of
.
This motivates the introduction of the -th
order anticipator
of
by
where
On the one hand, we will show in this section that
can be computed simultaneously by a dag
of
multiplicative size
and total size
. On the other hand, we will show that
is essentially a linear perturbation of
, which can be computed explicitly.
Let us show how to compute a dag for .
The following rules are straightforward:
As to multiplication, for ,
we have
Consequently,
for some polynomial with
and
(in particular,
). Notice also that
Now assume that and
are
known. Then we may simultaneously compute
in the
following way:
This computation involves one series product and
additions and scalar multiplications. For large
, we may further reduce the cost to
since the computation of
really comes down to
the computation of two truncated power series products
and
at order
.
In summary, we obtain
Since only depends on
, we notice that
In general, for and
, we may expand
Let denote the
-th
basis element of
, so that
for all
.
When considering
as a column vector, it follows
by linearity that
where is a row matrix with entries
Notice that depends on
, but
does not.
Let us investigate the functions more closely.
If
is algebraic, then we have
whence
In particular, is actually constant. If
is differential, of differential order
(this means that
is the maximal
number of
-nodes on a path
from the root of
to a leaf), then, considering
as a differential polynomial in
, we have
whence
![]() |
is a polynomial of degree at most .
Similarly, if
is algebraic in
, where
,
then
whence
Consequently, there exists a polynomial of
degree
with
for all .
For more general integral dags ,
it can be checked by induction over the size of
that
is still a rational function in
, which remains bounded for
, and whose denominator has integer
coefficients. Similarly, if
,
then
is a rational function in
, whose denominator has integer coefficients.
Assume now that we want to solve a system of power series equations
![]() |
(14) |
where is a vector of dags and
a finite number of initial conditions. For definiteness, it is also
important that (14) admits a unique solution
. This will be guaranteed by an even stronger
technical assumption to be detailed below. Roughly speaking, for a given
index
, we will
assume that each coefficient
with
can be determined as a function of the previous
coefficients
using only the equations
.
Let . For each
and
, we
introduce the
matrix
the block matrix
the ,
and
block column vectors
and the column vector
In view of (10), the equations then
translate into
We will say that (14) is -predictive
or predictive of index
,
if, for all
, there exist
and
matrices
and
, such that
In that case, we have
whence
provides us with an explicit formula for .
Now let
and
be the
operators on vectors of power series
with the
property that
and
.
Then we may rewrite (16) into
This is the desired recursive equation for .
Remark
Let us first consider the case of an algebraic dag . In that case, the matrix
does not depend on
and its coefficients are
explicitly given by (11). We may now determine
and
matrices
and
with
using Gaussian elimination in order, and whenever such matrices exist.
The equation -predictive, if and only if this is indeed
possible.
-predictive
equation
is algebraic. Then we may compute
terms of its unique solution
in time
Proof. By what precedes, the operators and
in (17) are
really the constant matrices from (18). By lemma 6,
the size of the righthand side of (17) as a dag is
therefore bounded by
and its multiplicative size
is exactly
. The result thus
follows from proposition 5.
Assume now that . Then we
claim that there exists an algorithms for checking
-predictivity and constructing a general
formula for the corresponding matrices
and
. Indeed, we recall from section 3.2 that
is the evaluation at
of a matrix
with entries in
and denominators in
. We may thus use Gaussian elimination in order to
compute
and
matrices
and
with entries in
and
whenever such matrices exist. For those which
are not positive integer roots of one of the denominators of the entries
of
, we now have
and
. For each
of the finite number of integer roots
,
we may directly compute the matrices
and
by Gaussian elimination over
, whenever such matrices exist.
-predictive
equation
be the maximal degree of an entry of
. Then we may compute
terms of the solution
to
Proof. The computation of
(and the finite number of exceptional
for which
is a root of one of the denominators) is a
precomputation. The determination of every next
can be done in time
, via a
diagonal translation of
and evaluation of the
rational functions which are the entries of the
bottom
submatrix. Now assume that we maintain
upper and lower triangular matrices
,
and a permutation matrix
at each stage such that
.
Then the determination of
,
and
as a function of
and
can be done in time
using naive linear algebra. The determination of
and
from
,
and
can again be done in time
.
Consequently, the cost of applying the operators
and
during the relaxed resolution of (17)
at order
is bounded by
. The cost of the evaluation of the remaining dag is
bounded by
, as in the
algebraic case.
If , then power series
solutions to recursive or implicit equations often converge in a small
disk around the origin. For instance, if
is an
algebraic dag and
, then the
recursive equation
admits a convergent solution, by Cauchy-Kovalevskaya's theorem. Given a
point in a sufficiently small disk where
converges, we might like to compute an approximation
of
,
together with a bound
for the error:
. Interval arithmetic [Moo66,
AH83, Neu90, MKC09, Rum10]
provides a classical framework for deriving such enclosures in a
systematic way. We will rather rely on ball arithmetic [vdH09a],
which is more suitable for complex and multiple precision arithmetic.
Let us briefly recall the principles behind ball arithmetic. Given a
normed vector space , we will
denote by
or
the set of
closed balls with centers in
and radii in
. Given such a ball
, we will denote its center by
and its radius by
.
Conversely, given
and
, we will denote by
the
closed ball with center
and radius
.
A continuous operation is said to lift
into an operation
on balls, which is usually
also denoted by
, if the
inclusion property
is satisfied for any and
. For instance, if
is a
Banach algebra, then we may take
Similar formulas can be given for division and elementary functions.
It is convenient to extend the notion of a ball to more general radius
types, which only carry a partial ordering. This allows us for instance
to regard a vector of balls as a
“vectorial ball” with center
and
radius
. If
, then we write
if and
only if
for all
.
A similar remark holds for matrices and power series with ball
coefficients.
In concrete machine computations, numbers are usually approximated by
floating point numbers with a finite precision. Let
be the set of floating point numbers at a given working precision, which
we will assume fixed. It is customary to include the infinities
in
as well. The IEEE754
standard [ANS08] specifies how to perform basic arithmetic
with floating point numbers in a predictable way, by specifying a
rounding mode
(down, up and nearest). A multiple
precision implementation of this standard is available in the
, we will denote by
its approximation using floating pointing arithmetic with
rounding mode
. This notation
extends to the case when
and
are replaced by their complexifications
and
.
Let and
or
and
. We will
denote by
or
the set of
closed balls in
with centers in
and radii in
. In this case,
we will also allow for balls with an infinite radius. A continuous
operation
is again said to lift to an
operation
on balls if (19) holds
for any
and
.
The formulas for the ring operations may now be adapted to
where ,
and
are reliable bounds for the rounding errors
induced by the corresponding floating point operations on the centers;
see [vdH09a] for more details.
In order to ease the remainder of our exposition, we will avoid
technicalities related to rounding problems, and compute with
“idealized” balls with centers in
and radii in
. For those who
are familiar with rounding errors, it should not be difficult though to
adapt our results to more realistic machine computations.
If we are computing with analytic functions on a disk, or multivariate
analytic functions on a polydisk, then Taylor models [MB96,
MB04] provide a suitable functional analogue for ball
arithmetic. We will use a multivariate setup with
as our coordinates and a polydisk
for a fixed
, where we use vector
notation. Taylor models come in different blends, depending on whether
we use a global error bound on
or individual
bounds for the coefficients of the polynomial approximation. Individual
bounds are sharper (especially if we truncate up to an small order such
that the remainder is not that small), but more expensive to compute.
Our general setup will cover all possible blends of Taylor models.
We first need some more definitions and notations. Assume that is given the natural partial ordering. Let
denote the
-th
canonical basis vector of
,
so that
and
for
. For every
, we will write
.
A subset
is called an initial segment,
if for any
and
with
, we have
. In that case, we write
and
. In what follows, we
assume that
and
are
fixed initial segments of
with
. For instance, we may take
and
or
or
.
Recall that or
.
Given a series
, we will
write
for its support. Given a subset
and a subset
,
we write
and
.
If
is analytic on
,
then we denote its sup norm by
A Taylor model is a tuple ,
where
,
and
are as above,
and
. We will write
for the set of such Taylor models. Given
and
, we will also denote
and
.
Given an analytic function
on
, we write
,
if there exists a decomposition
with and
for all
. Optionally, we may also require
that
for all
.
Given two Taylor models
, we
will say that
is included in
, and we write
if
for any
.
This holds in particular if
,
in which case we say that
is strongly
included in
and write
.
Addition, subtraction and scalar multiplication are defined in a natural
way on Taylor models. For multiplication, we need a projection with
for all
and
if
.
One way to construct such a mapping is as follows. For
, we must take
.
For
, let
be largest such that
. Then
we recursively define
. Given
, we now define their product
by
Using the observation that ,
this product satisfies the inclusion property that
for any analytic functions
and
on
. Finally, let
be a coefficientwise operator on
such that there exist constants
with
for any which converges on
. Then we say that
is
bounded and we may lift
into an
operator on Taylor models by taking
Indeed, the formula (20) implies the inclusion property
for any
.
Remark , then there exists a more algebraic
alternative for the definition of the product
. Let
be the finite set of
minimal elements of
and consider the ideal
in
generated by the relations
, for
. Then
is naturally
isomorphic to
as a vector space and we may
transport the product of
to
. However, this more algebraic definition does
not seem to be more convenient from the computational point of view.
Remark has the advantage that it is
easy to check, contrary to general inclusion of Taylor models. An
intermediate relation
which can still be checked
can be defined by recursion over
and
: if
,
then we take
if and only if
. Otherwise, let
be
largest for a total ordering on the monoid
(e.g. the lexicographical ordering). Let
,
,
and
.
If
, then we take
if and only if
.
Otherwise, we first compute the smallest
with
. Setting
, we now take
if and
only if
.
Most of the definitions and results of sections 2.1 and 2.4 generalize in a straightforward way to the multivariate
case. Only the complexity results need additional attention, since needs to be replaced by a suitable multivariate
analogue; see [vdH02, LS03, vdH05,
vdHL10, vdHS10] for various results in this
direction. Let us take
with for all
and
, and where
is set of bounded coefficientwise operators on
. Given
,
consider the equation
This equation is said to be recursive if
only depends on
with
, for every
.
In that case, the equation admits a unique solution, whose coefficients
may be computed using a relaxed algorithm.
Let ,
,
,
,
and
be as in the previous section. Consider a polydisk
with
.
Then any Taylor model
can naturally be
reinterpreted as a Taylor model in
,
which we will denote by
.
and let
be a vector of Taylor models with the property that
. Then the unique solution
to
and
.
Proof. Given power series
and
, we will say that
is a majorant for
, and write
,
if
for all
.
Let
be an arbitrary element of
and consider the sequence
.
By induction, we have
. In
particular, there exists a bound
with
for all
. For
any
, it follows that
Since (21) is recursive, the sequence
tends coefficientwise to the unique power series solution
of (21). More precisely, for any
and
such that
, we have
.
Therefore, given
, the
valuation of
is at least
. Consequently,
In particular,
We conclude that for
, whence
converges to
on the polydisk
.
Since
for all
,
we also get
, by
continuity.
Remark only holds
on slightly smaller disks. In would be interesting to investigate under
which conditions we have
.
This is in particular the case if
and
for all
.
Indeed, by continuity with respect to
,
this stronger condition implies that
for some
small
. We also notice that
ball arithmetic can be slightly “inflated” by adding a tiny
constant to the radius at every operation. Now assume that
for this kind of inflated ball arithmetic. For the usual
arithmetic and each
, we then
either have
, or the
expression which computes
is identically equal
to
.
With the notations from the previous section, assume in addition that
. Let
and let
be a power series with ball
coefficients. Then the restriction
can naturally
be reinterpreted as a Taylor model in
,
which we will denote by
. For
what follows, we recall that
is computed in
, whereas
is computed in
.
Proof. The lemma follows by an easy induction
over the size of . Let us for
instance treat the case when
.
Then
for all .
and a
Taylor model
, such that
.
Proof. Let be the exact
solution to (21). Let
,
and
Consider the Taylor model
By lemma 14, we have
for sufficiently small .
Remark with
, we first compute an enclosure
for
by solving (21)
in
using a relaxed algorithm. Then the problem
reduces to the computation of a bound
with
for
. This
really comes down to determining a fixed point for the mapping
. We refer to [vdH07b,
vdH09a] for details on how to do this.
We also notice that the coefficients of which
are not in
do not depend on
. For large expansion orders, it may be wise to
implement Taylor models in such a way that these coefficients need not
to be recomputed for different values of
.
In [vdH07b, vdH09a], we describe how to do
this in the univariate case.
Let us now return to the system (14) of implicit equations
and assume that is algebraic in
. We will use a domain
of Taylor models with
,
and
.
Using the theory of sections 3 and 4.1, we may
construct a recursive equation for the solution
of (14). By (17), this equation has the form
for certain matrices and
whose entries are coefficientwise operators. In order to apply theorems
12 and 15 we need to show that these operators
and
are bounded.
Let us first show that is bounded. Let
be a convergent series on
.
For any
and
,
we have
We may therefore take for
and
.
Now consider one of the entries of
or
. By
choosing
sufficiently large, we may assume
without loss of generality that
is given by the
evaluation of a rational function in
at
for all
.
Moreover, the coefficients of the matrix (15) are given by
(13), whence they are bounded for
. Consequently,
also remains
bounded for
, and there
exists an absolutely convergent expansion
Hence, we may regard as an infinite sum
Taking sufficiently large, we may also ensure
that (22) converges for
.
Consequently, it suffices to take
for all . We may use crude
bounds for the remaining
with
. Combining theorems 9, 12
and 15, we now obtain the following theorem.
-predictive
system of equations
and a Taylor model
, such that its unique solution
is convergent on
and
.
In the Taylor model setting, it is interesting to study the solution
under small perturbations of the initial
conditions. This can be done by switching to a multivariate context as
in the previous subsections. In the system (14), we now
replace the time
by
. We also replace the initial conditions
by Taylor models
in
which only depend on
,
and such that
for all
. Although the above theory mostly generalizes in a
straightforward way, there is a small quirk: when a matrix of the type
(15) does not have full rank, then the rank of a small
perturbation of it is generally strictly higher. Consequently,
-predictivity is not preserved
under small perturbations.
Nevertheless, if we restrict our attention to -predictive systems, for which the matrix (15)
does have full rank, then small perturbations of the system remain
-predictive and the generalization
of theorem 17 provides us with the complete flow of the
implicit equation at any order. Such
-predictive
systems are very common in practice and a well known example is the
pendulum, whose equations are given by
More generally, dynamical systems on implicitly defined varieties are
-predictive systems. Notice
that the classical implicit function theorem also falls into this
category.
Of course, if the solution is known in a certain
disk
(as well as some of its derivatives, if
necessary), then we may use the value of
at a
given point
as a new initial condition at
and compute an analytic continuation of the solution
around
with the same method. In general
-predictivity is preserved
throughout this process, whence our method yields an efficient method
for high precision integration of dynamical systems on implicitly
defined varieties. Predictive systems of a higher index usually degrade
into
-predictive systems
after one analytic continuation. If not, then there is usually an
algebraic reason for that, in which case the system can be rewritten
into a “simpler system” using differential algebra [Rit50,
Kol73, Bou94]. Unfortunately, although the
order (or ranking) of the new system is lower, its size as a dag usually
explodes.
G. Alefeld and J. Herzberger. Introduction to interval analysis. Academic Press, New York, 1983.
ANSI/IEEE. IEEE standard for binary floating-point arithmetic. Technical report, ANSI/IEEE, New York, 2008. ANSI-IEEE Standard 754-2008. Revision of IEEE 754-1985, approved on June 12, 2008 by IEEE Standards Board.
A. Bostan, F. Chyzak, F. Ollivier, B. Salvy, É. Schost, and A. Sedoglavic. Fast computation of power series solutions of systems of differential equation. preprint, april 2006. submitted, 13 pages.
R.P. Brent and H.T. Kung. Fast algorithms for manipulating formal power series. Journal of the ACM, 25:581–595, 1978.
J. Berthomieu and R. Lebreton. Relaxed -adic hensel lifting for algebraic
systems. Work in preparation, 2011.
F. Boulier. Étude et implantation de quelques algorithmes en algèbre différentielle. PhD thesis, University of Lille I, 1994.
D.G. Cantor and E. Kaltofen. On fast multiplication of polynomials over arbitrary algebras. Acta Informatica, 28:693–701, 1991.
J.W. Cooley and J.W. Tukey. An algorithm for the machine calculation of complex Fourier series. Math. Computat., 19:297–301, 1965.
G. Hanrot, V. Lefèvre, K. Ryde, and P. Zimmermann. MPFR, a C library for multiple-precision floating-point computations with exact rounding. http://www.mpfr.org, 2000.
E.R. Kolchin. Differential algebra and algebraic groups. Academic Press, New York, 1973.
G. Lecerf and É. Schost. Fast multivariate power series multiplication in characteristic zero. SADIO Electronic Journal on Informatics and Operations Research, 5(1):1–10, September 2003.
K. Makino and M. Berz. Remainder differential algebras and their applications. In M. Berz, C. Bischof, G. Corliss, and A. Griewank, editors, Computational differentiation: techniques, applications and tools, pages 63–74, SIAM, Philadelphia, 1996.
K. Makino and M. Berz. Suppression of the wrapping effect by Taylor model-based validated integrators. Technical Report MSU Report MSUHEP 40910, Michigan State University, 2004.
R.E. Moore, R.B. Kearfott, and M.J. Cloud. Introduction to Interval Analysis. SIAM Press, 2009.
R.E. Moore. Interval Analysis. Prentice Hall, Englewood Cliffs, N.J., 1966.
A. Neumaier. Interval methods for systems of equations. Cambridge university press, Cambridge, 1990.
J.F. Ritt. Differential algebra. Amer. Math. Soc., New York, 1950.
S.M. Rump. Verification methods: Rigorous results using floating-point arithmetic. Acta Numerica, 19:287–449, 2010.
A. Schönhage and V. Strassen. Schnelle Multiplikation grosser Zahlen. Computing, 7:281–292, 1971.
J. van der Hoeven. Lazy multiplication of formal power series. In W. W. Küchlin, editor, Proc. ISSAC '97, pages 17–20, Maui, Hawaii, July 1997.
J. van der Hoeven. Relax, but don't be too lazy. JSC, 34:479–542, 2002.
J. van der Hoeven. Notes on the Truncated Fourier Transform. Technical Report 2005-5, Université Paris-Sud, Orsay, France, 2005.
J. van der Hoeven. New algorithms for relaxed multiplication. JSC, 42(8):792–802, 2007.
J. van der Hoeven. On effective analytic continuation. MCS, 1(1):111–175, 2007.
J. van der Hoeven. Ball arithmetic. Technical report, HAL, 2009. http://hal.archives-ouvertes.fr/hal-00432152/fr/.
J. van der Hoeven. Relaxed resolution of implicit equations. Technical report, HAL, 2009. http://hal.archives-ouvertes.fr/hal-00441977/fr/.
J. van der Hoeven. Newton's method and FFT trading. JSC, 45(8):857–878, 2010.
J. van der Hoeven and G. Lecerf. On the bit-complexity of sparse polynomial multiplication. Technical report, HAL, 2010. http://hal.archives-ouvertes.fr/hal-00476223/fr/.
J. van der Hoeven and É. Schost. Multi-point evaluation in higher dimensions. Technical report, HAL, 2010. http://hal.archives-ouvertes.fr/hal-00477658/fr/.