|
Abstract
Let and
, where
is an effective
field and
and
are
given a suitable asymptotic ordering
.
Consider the mapping
,
where
. For
, it is natural to ask how to solve the
system
. In this paper, we
will effectively characterize
and show how to
compute a so called distinguished right inverse
of
. We will also
characterize the solution space of the homogeneous equation
.
A well-known theorem [Fab85] states that any linear
differential equation over admits a basis of
formal solutions of the form
with ,
,
and
. This theorem naturally generalizes to the case
when
is replaced by an effective algebraically
closed field of coefficients
.
If we also replace the coefficients by polynomials in
, then several algorithms exist for the
computation of a basis of solutions [Mal79, DDT82,
vH97].
There are several directions in which the above theorem may be
generalized. In [vdH97, vdH01, vdH06],
it is shown how to deal with so called transseries coefficients (a
transseries is an object which is constructed from
or
and an infinitely large variable
using exponentiation, logarithm and infinite summation).
In collaboration with M. Aschenbrenner and L.
van den Dries, we are currently working on a generalization to arbitrary
asymptotic fields (an asymptotic field is a differential field with a
total asymptotic ordering which is “naturally compatible”
with the derivation).
In this paper, we will be concerned with the generalization to the case
of linear partial differential equations. The asymptotic resolution of
systems of such equations can be decomposed into two subproblems: the
computation of analogues of the exponential parts
and the computation of the corresponding coefficients. We intend to deal
with the first subproblem in a forthcoming paper and focus on the second
subproblem in what follows.
In the case of holonomic systems of linear differential equations,
algorithms are known for the computation of formal and convergent
generalized series solutions [SST00, Chapter 2] in what the
authors call “Nilsson rings” [Nil65]. On the
other extreme, there exists a method [AC01] to find
“fractional power series solutions” to a single
p.d.e. with coefficients in .
In this paper, we will search for formal series solutions to consistent
systems of linear differential equations in variants of Nilsson rings of
the form
. One of the major
difficulties is to cope with the integrability constraints which arise
when considering more than one equation.
In fact, in the continuation of our previous work on transseries, we
will rather work with infinitely large variables
and series in
. In this
equivalent setting, our linear differential operators belong to
and we consider series in
where
. More precisely, we
assume a total asymptotic ordering
on
and consider so called grid-based series [vdH97,
vdH06] with monomials in
and
coefficients in
.
In sections 2 and 3 we first recall classical
algorithms for the computation of “standard bases”, which
are used to reduce a system of equations like
with
and
to suitable
normal forms. The first algorithm is a variant of the skew version [Cas84, Cas87, Gal85, Tak91]
of Buchberger's algorithm [Buc65, Buc85],
although we rather compute coherent autoreduced sets in the sense of
differential algebra [Ros59]. We also recall Mora's
standard cone algorithm [Mor83, MPT92].
However, we will systematically present them in the setting of
p.d.e.s with second members, so the reader might at least
want to take a look at the notations. Also, corollaries 7
and 12 characterize when a system of equations with second
members satisfies the necessary integrability constraints which ensure
the existence of a solution.
In section 4, we will start with the study of linear
p.d.e.s with constant coefficients in . It is classical that the resolution of such
equations in
is equivalent to finding the roots
of a set of polynomial equations in
.
In particular, solution sets in
correspond to
radical ideals in
. More
generally, we will show that there exists a correspondence between the
solution sets in
and arbitrary ideals in
.
An important technique that we will use is the computation of so called
“distinguished solutions” to systems of equations with
second members. More precisely, given ,
we may consider
as an operator
. Denoting
,
we will effectively construct a right-inverse
of
. This right-inverse is
unique with the property that the coefficient of any
in any
vanishes, where
denotes the set of dominant monomials of solutions
to
. Having constructed
, we will also show how the space
of solutions
to
can be
obtained from
.
In the last section 5, we will study the case of linear
p.d.e.s with coefficients in (for
effective purposes) and
(for theoretical
purposes). We will first show how to reduce systems of such equations to
suitable asymptotic normal forms. Given a system in normal form, we will
next show how to compute a distinguished right inverse in a
coefficientwise manner. We will also characterize the set
in this context and give an explicit “strong
basis” for
.
Remark
Consider the “monomial monoid” ,
whose elements are of the form
,
with
. A total ordering
on
is called a monomial
ordering, if it is compatible with the multiplication,
i.e.
. It is
classical [Rob85] that any such an ordering is non uniquely
determined by a finite sequence of vectors
and
![]() |
(1) |
Here denotes the scalar product. Clearly, the
relation (1) allows to extend
to
and even
.
Moreover, this extension is unique so as to preserve the compatibility
with the multiplication.
We say that is admissible if
for all
. In
that case,
extends the (partial) divisibility
ordering
on
.
In particular, from Dickson's lemma, it follows that
is well-ordered. Given a subset
,
we will denote by
the final segment of
generated by
for the divisibility
relation. We recall that each final segment is finitely generated.
Let be a constant field of characteristic zero.
Given a monomial ordering on
,
a non-zero polynomial
and a monomial
, we denote by
the coefficient of
in
. We also denote by
the
highest monomial for
occurring in
and by
the corresponding
coefficient. We call
the dominant
monomial of
,
its dominant coefficient and
its dominant term. The relation
naturally extends to
by
. We denote by
the
equivalence relation associated to
,
so that
. Similarly, we write
if
,
which is equivalent to
.
Let be a differential field with derivations
and field of constants
. Given formal variables
, we denote by
the
differential algebra of differential polynomials in
over
. Any
admits a unique decomposition
where each is homogeneous of degree
. We denote by
the
space of homogeneous polynomials of degree
and
.
Given and a tuple
,
the substitution of
for
(
) in
yields a new tuple of differential polynomials in
, called the composition of
and
, and denoted by
. If
and
, then
. In particular,
is an
algebra for
and
is a
subalgebra of
whenever
and
are subalgebras of
with
. If
, we will denote by
the
unique elements of
, such
that
.
Example
then
Example
are differential operators, then
In other words, is isomorphic to
.
In the remainder of this paper, we will only study differential
polynomials with second members with
and
as above (and often
). Formally speaking, the monomial
monoid
for
is isomorphic
to
. Consequently, the sets
and
are isomorphic as
vector spaces (but not necessarily as algebras, except when
). This isomorphism induces natural definitions
of
,
and
for
and of
,
,
and
on
. These definitions naturally extend to
, by taking
for all
. For instance,
, if
.
In our context of linear differential polynomials with second members, a
differential ideal of is a
-subvector space which is stable under
(i.e. left composition with
). Moreover, if
,
then we require that
. Any
tuple
naturally generates a differential ideal
. If
, then
.
When seeing
as a system of equations
, where
belongs to any differential
-algebra
, then these equations are
equivalent to
. In
particular, we say that a second system
is
equivalent to
, if
.
In the sequel, it will be convenient to extend notation for sets to
tuples. For instance, and
if
is smallest with
and
if no such
exists.
Assume that we fixed an admissible monomial ordering on
and denote
. Let
with
, so that
for some
. The partial
reduction
of
w.r.t.
is defined by
![]() |
(2) |
Given a system , let
. A normal form for
modulo
is an
, with
or
, and such that
for certain (where
)
and
with
for all
. In that case, we write
. We say that
is autoreduced if
for all
. By using partial reductions of
w.r.t. members of
as
long as possible, one obtains a normal form with
:
Algorithm
Input: and
Output: a normal form of
modulo
while do
return
Given , let
be such that
and
.
Setting
,
and
, the
-polynomial of
and
is defined by
![]() |
(3) |
By construction, be have . We
also notice that
, whenever
. We say that a system
is coherent if
for
all
. A coherent and
autoreduced system
will also be called a
standard basis. Given an arbitrary system
, the following classical algorithm computes a
standard basis which is equivalent to
.
Algorithm
Input:
Output: a standard basis which is equivalent to
while
if then
return
while do
if then
return
Remark , then under the natural
isomorphism of
with
, the notions of partial reduction and
-polynomials correspond to reduction and
-polynomials in Buchberger's
algorithm (up to details: Buchberger rather takes
. Also, he not only reduces the dominant term of
in
,
but all terms). Consequently, Buchberger's algorithm for computing a
Gröbner basis [Buc65, Buc85] corresponds
to the above algorithm for computing a coherent autoreduced set.
Coherent autoreduced sets were first introduced by Rosenfeld [Ros59]
and they are similar (although more effective) to the characteristic
sets introduced by Ritt [Rit50]. We opted for Hironaka's
name standard bases here [Hir64] in view of the
generalization in the next section.
Consider a standard basis .
Then the reduction of each
-polynomial
with
to zero yields a
relation
with for all
.
This relation may be rewritten as
![]() |
(4) |
with . We call (4)
the critical relation for the pair
. Notice that we may regard the set of all critical
relations as a tuple
.
Lemma be a standard basis. Then the
generate the space of all
with
. In other words, given
with
, there exists a
with
.
Proof. Assume for contradiction that there exists a
relation which is not generated by the
. We may choose
such that
is minimal, as well as the number of
with
.
Since
, there must be at
least two indices
and
with
. Using the fact that
divides
,
let
be such that
,
and
.
By construction,
and
, so
and
. For all
,
we also have
, so
. It follows that the relation
is smaller than the original relation
in the sense of the minimality hypothesis. This
contradiction completes the proof.
Consider a system of linear differential
polynomials in
. Given a
tuple
, we say that
is compatible with
,
if for every relation
with
, we have
.
The set of such tuples forms a subvector space of
, which we denote by
.
Corollary with
and
is a standard basis if and only if
is a standard basis and
is compatible with
.
Proof. Assume that is a
standard basis. Consider
with
and
. Then
if
and
.
It follows that
is a standard basis with
critical relation
for all
. Given a relation
,
lemma 5 now implies
for a certain
. We conclude that
, whence
.
Assume now that is a standard basis and that
is compatible with
.
Then
is autoreduced, since
for all
. Furthermore, for
all
, the relation
implies
. But
the relation
precisely proves that
reduces to zero modulo
.
Hence
is coherent.
Corollary and
, we have
if and only if
.
Let and
.
A canonical form for
modulo
is an
with
for certain and
with
for all
,
and such that
for each term
occurring in
. It is easy to
modify
so that it computes a canonical form
of
modulo
with
:
Algorithm
Input: and
Output: a canonical form of
modulo
while do
Choose highest for
return
Lemma be a standard basis. Then we have
where and
.
Proof. Assume for contradiction that
is such that
. Replacing
by
, we
may assume without loss of generality that
is a
canonical form w.r.t.
.
Now choose
with
such
that
is minimal, in the same sense as in the
proof of lemma 5. Since
,
we have
, so there must be at
least two indices
and
with
. Setting
with the notations from the proof of lemma 5,
then yields a more minimal representation for
. This contradiction proves
that
for all
.
In classical polynomial elimination theory, the use of non-admissible
monomial orderings allows for the computation in localized rings and
completions, such as rings of power series. However, additional care is
needed in order to ensure termination. For instance, the naive reduction
of modulo
would yield an
infinite sequence
. The
tangent cone algorithm [Mor83, MPT92] allows
for the computation of standard bases in the case of localizations of
polynomial rings.
In this section, we will present the tangent cone algorithm in the
differential setting. In all what follows, is a
differential field with constant field
.
Geometrically speaking, elements of
or
localizations of
can still be thought of as
operators. For instance,
naturally operates on
.
Let and let
be a monomial ordering on
.
Given
, we define
,
and
as in (3). As a special case,
is given by (2) if
. Now let
be the opposite
ordering of
. Given
, we denote the dominant monomial
of
for
by
for and we define
,
,
,
,
etc. in a similar way. We will also write
for the element of
with
. If
is admissible,
and
for all
, then we notice that
(i.e.
for the natural extension of
the ordering
to
).
Moreover, if
, then
.
In the sequel, we will assume that the vectors
which determine
using (1) are all
in
. In that case
is called a tangent cone ordering. Notice that it
is possible to consider more general tangent cone orderings [MPT92],
but we have chosen to keep the exposition as simple as possible. Given
, let
Given with
,
we denote
Notice that and
(for a
dummy
). Now let
and
be such that
and
. Then we have
and we define the
-th
ecart of
by
.
We call
the ecart of
and recall that
is well-ordered by the
lexicographical ordering. The definition extends to the case when
by taking
for all
.
Given , some easy properties
of the ecart are
![]() |
(5) |
Moreover, if , then
where the inequality is strict whenever .
It follows that
![]() |
(6) |
In particular, if , then
![]() |
(7) |
The following lemma will guarantee the termination of the tangent cone algorithm.
Lemma and
be such that for all
, we have
for some
.
Whenever for some
, then
.
Then the sequence is finite.
Proof. Assume for contradiction that there exist
infinitely many with
. By Dickson's lemma, we may find two such indices
with
and
. But then
which contradicts our assumption (b). It follows that is strictly decreasing for sufficiently large
. We conclude by the fact that
is well-ordered.
Given , a
normal form for
modulo
is an
, with
or
, and such that
for certain and
with
and
for all
. Notice that this notion extends the previous
notion of normal forms, since
if
is admissible. In our new context, we may use the
following algorithm to compute a normal form:
Algorithm
Input: and
Output: a normal form of
modulo
while do
Take with
such that
is minimal
return
Indeed, the sequence of successive values of
during the algorithm fulfills the conditions of
lemma 9, so this sequence is finite. Moreover, using
induction, it is easily checked that there exist
and
and with
and
for all
.
So the last term of the sequence is indeed a normal form for
modulo
.
Defining the notions of autoreduced systems, coherent systems and
standard bases as in section 2.3, the same algorithm may be used to compute an equivalent standard basis
for a given system. Given a standard basis
and
, we have a relation
with and
for all
. As before, we may rewrite this
relation as a critical relation of the form
.
In order to generalize lemma 5, let
be the set of series
with well-ordered support
. If
is admissible, then
coincides with
. If
is admissible then
elements of
are power series in
applied to
. The set
is naturally stable under composition. We denote
.
Lemma be a standard basis. Then the
generate the space of all
with
.
Proof. We have to construct
with
, where
corresponds to the critical relation
.
For each
, let
. Writing
,
let us construct
by transfinite induction over
. Given an ordinal
, the induction hypothesis is as
follows:
has been constructed for all
in a final segment
of
for
.
for all
.
Denoting and
,
we have
for all
and
.
If or
is a limit
ordinal, then we may take
.
If
and
,
then we are done. So assume that
and
. Let
and
let
be minimal such that
. Let
,
with
let and take
for all
with
.
By construction,
for all . Since
, it follows that
as
well. This proves the last induction hypothesis. By transfinite
induction, we conclude that there exists an
with
, whence
.
Consider a system of linear differential
polynomials in
. Assume also
that
naturally operates on a subring
of
(for instance, we may take
). Given a tuple
, we say that
is
compatible with
, if for
every relation
with
, we have
.
The set of such tuples forms a (strong) subvector space
of
. The following
consequences of the above lemma is proved in a similar way as
corollaries 6 and 7.
Corollary
Corollary
Let and
.
A canonical form for
modulo
is an
with
for certain and
with
and
for all
, and such that
for
each term
occurring in
. Although we have no algorithm to compute canonical
forms, like in section 2.5, the existence of canonical
forms can be proved using a similar transfinite induction as in the
proof of lemma 10. Using another transfinite induction,
lemma 8 also generalizes to the current setting:
Lemma
In this section, we consider systems of linear
partial differential equations in one unknown
with coefficients in a field of constants
of
characteristic zero. We will consider the resolution of such systems in
the algebras
where (Kronecker symbol). We will first consider
homogeneous linear differential equations, but we will also study linear
differential equations with second members. In the latter case, we will
allow the second members to belong to
or
. Throughout this section
stands for an admissible tangent cone ordering on
.
In this section, we will only consider linear p.d.e.s
without second members. Let be a homogeneous
linear differential polynomial. We may represent
as
where is a polynomial in
. Inversely, each polynomial
gives rise to a homogeneous linear differential polynomial
. Denoting
,
we have
for all and in particular
Let denote the set of all
with
. We have
for all , where
denotes the
-vector
space generated by
. Given
, we will denote
.
More generally, given a set of homogeneous
linear differential polynomials, a subset
of
, a subset
of
and a subset
of
, we denote
and
Because of the natural isomorphisms
all algebraic geometry properties of the correspondences and
induce analogue properties for
the correspondences
and
. In particular, Hilbert's Nullstellensatz implies
Theorem be a coherent and autoreduced system with
. If
is
algebraically closed, then
admits a solution
.
Recall that stands for an admissible
tangent cone ordering on
.
Consider a standard basis
for
. We may regard
as an
operator from
into
,
whose image is in
. We denote
by
the set of monomials
, such that
for all
. The aim of this section is to
construct a right inverse
of
, which is “distinguished” in the
sense that
for all
and
.
The relation on
induces
a relation
on
by
Whenever are such that
and
, it follows that
Indeed, if , then
for at least one
with
.
Proposition for
and
, let
be such that
is maximal for
. Then
does not depend
on the choice of
and
.
Proof. We will first show that
whenever
is another index with
. Let
and
be such that
and consider the
associated critical relation
with for all
.
Since
is compatible with
, it follows that
For each , we have
It follows that
![]() |
(8) |
Hence
It follows that
Now implies
,
so we conclude that
.
It remains to be proved that ,
i.e.
for all
. If
,
then
and for all
with
, we have
. By strong linearity, it follows that
. Furthermore
implies
, whence
. If
,
then the relation (8) remains valid. Moreover, if
is such that
,
then
and
Since , it follows that
, whence
. By construction, we therefore have
.
Given , let
be the term as in proposition 15. Now consider the sequence
defined by
and
.
This sequence is finite, since
and
is a well-ordering on
.
Consequently,
is a solution to
with
for all
.
We set
and call
the
distinguished right inverse of
.
Let . Since
for all
, it follows that
. Consequently
is a solution to
with
. Inversely,
implies
, since otherwise
for some
and
. We claim that the
form a basis for the solution space
of
in
. Indeed,
given an arbitrary solution
,
consider the sequence defined by
and
as long as
.
This sequence is necessarily finite, since
and
is well-ordered. Hence,
. We call
the
distinguished basis of
.
We notice that , where
, so that
decomposes into an isomorphism on
with left
inverse
and the zero map on
. We also notice that the distinguished right
inverse
is uniquely determined by the fact that
for all
and
. Indeed, assume that
and
and
for all
. Then
for
and
for all
. It follows that
.
Let us now consider an arbitrary system .
Using the tangent cone algorithm,
may be
rewritten into an equivalent system
which is a
standard basis. Then the sets
and
are independent from the particular choice of
, since
is precisely
the set of elements which cannot occur as dominant monomials of elements
in
, by lemma 13.
Consequently, the construction of the distinguished right-inverse and
the distinguished basis
do not depend on the
choice of
, and we may define
,
, etc.
Let us now consider a general system as an
operator
. Then
acts “by spectral components”
. More precisely, given
, let
be the unique operator
such that
for all . Considering
as an operator in
,
we obtain
from
by
substituting
for each
. Given
with , it follows that
Hence, denoting by the solution space of
for
, the
solution space of
for
is
given by
Denoting by the distinguished inverse of
as an operator on
,
the mapping
is a right-inverse of .
Moreover,
is unique with the property that
, where
Remark on
to
in any way which preserves spectral components
(i.e. if
, then
for all
),
the space
coincides with the set of all
such that
for all
; see the next section.
Theorem be the set of differential ideals of
and let
the set of subsets of
which occur as zero-sets of systems
. Then the correspondences
are mutually inverse bijections.
Proof. Let and
be two differential ideals with the same set of solutions
. Then the differential ideal
generated by
and
still
has the same set of solutions. Assuming for contradiction that
, the set
strictly contains
or
, say
.
Now consider the differential ideal
.
By theorem 14, there exists an
with
. Since
and
(here
stands for
, where
is any system which generates
),
it follows that
. But then
and
.
Remark (which is
necessarily radical and even prime) and “linear differentially
algebraic” zero-sets in
.
Via the isomorphism
,
arbitrary ideals of
are therefore also in a
geometric correspondence with zero-sets of linear differential
operators. This provides a geometrical reason why the existence of
Ritt-Rosenfeld-Buchberger-type algorithms for the computation with
ideals, and not merely radical ideals, is important.
The study of the linear p.d.e.s with
coefficients in is equivalent to the study of
equations with coefficients in
modulo the
substitutions
,
and multiplication with a suitable
. Since the ordinary partial derivatives preserve
the “valuation” in
,
it will be more convenient to work with coefficients in
.
Assume that we have fixed an admissible ordering
on
, determined by
. Assume also that we have fixed a
total ordering on
which gives
the structure of a totally ordered
-vector
space. Then we also have a natural ordering
on
:
A subset of
is said to
be grid-based if there exist
with
and
.
Given a ring of coefficients
the set of series
with grid-based support
forms an
-algebra [vdH97,
vdH06]. We denote this algebra by
and its elements are called grid-based series. This still goes
through for coefficients in
,
since such operators act by spectral components. In this section, we
will consider systems of linear p.d.e.s in
and study their solutions in
.
The admissible orderings on
and
on
may
be combined into a total admissible ordering
on
using
Consequently, an element can also be regarded as
a series
with anti-well-ordered support in
(the support is not necessarily grid-based, although
we might have required this). Similarly, elements in
can be seen as series with monomials in
.
The ordering
is extended to
by understanding that
for all
. We will use
in order
to emphasis when a notation should be understood with respect to the
relation
.
Consider a system such that
for all
. Given
with
and
, let
,
,
and
We say that is a standard basis for
if for each
there exists
a critical relation
![]() |
(9) |
where is such that
for
all
.
Given with
(or
with
), let us
denote
(resp.
).
Lemma be a standard basis and let
be such that
. Then
is a standard basis and
.
Proof. Since for all
, the system
is autoreduced. For all
, the
relation (9) implies
so is a standard basis for the relations
. Now consider a relation
. Then we have
for some
. Now
implies
. We
conclude that
, so
.
Lemma be a standard basis and
.
Then
is again a standard basis.
Proof. Any satisfy the
relations
Hence, any critical relation for
induces a critical relation
for
. So we may take
.
Given an arbitrary system ,
an equivalent standard basis can be “computed” by a variant
of Hironaka's infinite division “algorithm”. If the
dependency of
in
is only
polynomial, then a fully effective method can be devised, by adapting
the algorithms from section 2.3.
In this subsection and in this subsection only, let ,
,
and
.
The set
is formally isomorphic (as a vector
space) to
by sending each
to
and
to
. Moreover, the ordering
on
corresponds to a tangent cone ordering on
. Consequently, the
definition of ecart in section 3.2 transposes to elements
in
.
Unfortunately, we do not necessarily have for
and
(for instance
). Nevertheless, this relation does
hold if
. For this reason, we
adapt the definition of partial reduction by setting
for all with
.
Because of the twist, we again have
We also notice that coincides with the usual
partial reduction “up to lower order terms”, since
. We obtain the following version
of
:
Algorithm
Input: and
Output: an “asymptotic normal form” of
modulo
while do
Take with
such that
is minimal
return
The termination of the modified version of NF is proved in the same way
as before. Again, the successive values of
in the algorithm verify relations
for certain ,
and
with
and
for all
.
Example
and
. Then
Hence divides
,
from the asymptotic point of view.
In a similar way, we may define the twisted -polynomial of
by
Given a system , the
corresponding algorithm
now computes an
equivalent system
, such that
for all
we have a relation
where ,
and
are such that
and
for all
.
But
admits
as inverse in
, which leads to the relation
![]() |
(10) |
Moreover, each induces an element
with . When rewriting (10) in terms of
and
, we obtain a critical relation for
and
in the sense of section 5.1. Modulo this normalization of the result, SB therefore
computes a skew standard basis.
Let again ,
and
. Let
. Using the isomorphism
, we observe that
,
, etc. are
well-defined. Given
it is also convenient to
extend the notation
by setting
if and only if
.
Lemma be a standard basis for
and
let
be such that
.
Then there exists a
with
and
. In particular, if
, then there exists a
with
.
Proof. Let with
be such that
.
Then
is stable under composition. For each
with
, let
us show how to construct
,
such that
satisfies
. We use weak induction over
.
So let and assume that
has been constructed for all
.
Let
. Since
for all
with
,
and
, we have
. Setting
,
we have
, so
for some
. Taking
, it follows that
.
By induction, we conclude that is well-defined
and we have
for all
, so
.
Corollary and
, we have
if and only if
.
Proof. Similar to the proofs of corollaries 6
and 7.
Corollary is a standard basis for
and let
be non-zero. Then
.
Proof. Let be such that
. Modulo division of
by
, we may
assume without loss of generality that
.
Let
be as in the above lemma, so that
. In fact,
, since
implies
. We conclude that
.
Consider a standard basis for
. Given
,
we may regard
as an operator on
. We denote
and write for the distinguished right inverse of
.
Proposition be a standard basis for
.
Then
admits a unique right inverse
such that
for all
.
Proof. Let with
and
be such that
and
. For any
with
, it
follows that
. Let us show by
well-ordered induction over
how to construct
such that
for
.
Given , we assume that
has been constructed for all
with
. Denoting
, we also assume that
for all
with
.
By construction, we first observe that
,
whence
. Now we take
, which is well-defined by lemmas
20 and 19. Setting
, it follows that
.
For all
with
,
we also have
. We infer that
. By induction, we obtain a
series
with
and
for all
.
We conclude that
. The
uniqueness is proved as usual.
Proposition be a standard basis for
.
For each
, let
. Then
for all solutions to
.
Proof. Setting
we have
Now for all
,
by the distinguished property of
and the fact
that
. Consequently,
and
. But
this is only possible if
.
Let us now consider an arbitrary system and let
be an equivalent standard basis. By corollary 24, we notice that the differential ideal
does not depend on the particular choice of
, and similarly for the twisted differential ideals
. Consequently, the spaces
,
and
the operator
are independent of the particular
choice of
. We may therefore
define the distinguished right inverse
of
by
.
Putting everything together from the effective point of view, we have:
Theorem and
, computes the asymptotic expansion of
.
Proof. Using the algorithm from section 5.2,
we start by computing an equivalent standard basis
for
and make the corresponding change
for
. We next
test whether
is compatible with
using corollary 23. If so, and assuming that
, we determine the dominant term
of
and compute the dominant term
of
using the method from
section 4.2. Setting
and continuing
the same procedure with
instead of
, we obtain the asymptotic expansion of
.
Remark , where
.
Remark and
not merely the first
terms (as done by the above
algorithm).
F. Aroca and J. Cano. Formal solutions to linear pdes and convex polyhedra. JSC, 32:717–737, 2001.
F. Boulier. Étude et implantation de quelques algorithmes en algèbre différentielle. PhD thesis, University of Lille I, 1994.
B. Buchberger. Ein Algorithmus zum auffinden der Basiselemente des Restklassenringes nach einem null-dimensionalen Polynomideal. PhD thesis, University of Innsbruck, 1965.
B. Buchberger. Multidimensional Systems Theory, chapter Gröbner bases: an algorithmic method in polynomial ideal theory, pages 184–232. Reidel, 1985. Chapter 6.
F. Castro. Théorème de division pour les opérateurs différientiels et calcul des multiplicités. PhD thesis, Université Paris-VII, 1984.
F. Castro. Calculs effectifs pour les idéaux d'opérateurs différentiels. In Travaux en cours, volume 24, pages 1–19. Hermann, Paris, 1987.
J. Della Dora, C. Discrescenzo, and E. Tournier. An algorithm to obtain formal solutions of a linear homogeneous differential equation at an irregular singular point. In Eurocal '82, volume 174 of Lect. Notes in Comp. Science, pages 273–280, Marseille (France), 1982. Springer Verlag.
E. Fabry. Sur les intégrales des équations différentielles linéaires à coefficients rationnels. PhD thesis, Paris, 1885.
A. Galligo. Some algorithmic questions on ideals of differential operators. In Proc. EUROCAL '85, volume 204 of Lecture Notes in Computer Science, pages 413–421. Springer-Verlag, 1985.
M. Granger, T. Oaku, and N. Takayama. Tangent cone algorithm for homogenenized differential operators. JSC, 39:417–431, 2005.
H. Hironaka. Resolution of singularities of an algebraic variety over a field of characteristic zero. Annals of Math., 79:109–326, 1964.
B. Malgrange. Sur la réduction formelle d'équations différentielles à singularités irrégulières. Manuscript, 1979.
T. Mora. An algorithm to compute the equations of tangent cones. In Proc. EUROCAL '83, number 144 in Lect. Notes in Computer Sc., pages 158–165, 1983.
T. Mora, G. Pfister, and C. Traverso. An introduction to the tangent cone algorithm. In Advances in Comp. Research, volume 6, pages 199–270. JAI Press, 1992.
N. Nilsson. Some growth and ramification properties of certain integrals on algebraic manifolds. Arkiv für Mathematik, 5:463–476, 1965.
J.F. Ritt. Differential algebra. Amer. Math. Soc., New York, 1950.
Lorenzo Robbiano. Term orderings on the polynominal ring. In European Conference on Computer Algebra (2), pages 513–517, 1985.
A. Rosenfeld. Specializations in differential algebra. Trans. Amer. Math. Soc., 90:394–407, 1959.
M. Saito, B. Sturmfels, and N. Takayama. Gröbner deformations of hypergeometric differential equations, volume 6 of Algorithms and Computation in Mathematics. Springer-Verlag, Berlin, Heidelberg, New York, 2000.
N. Takayama. Kan: a system for computation in algebraic analysis. http://www.math.kobe-u.ac.jp/KAN/, 1991.
J. van der Hoeven. Automatic asymptotics. PhD thesis, École polytechnique, France, 1997.
J. van der Hoeven. Complex transseries solutions to algebraic differential equations. Technical Report 2001-34, Univ. d'Orsay, 2001.
J. van der Hoeven. Generalized power series solutions to linear partial differential equations. Technical Report 2004-51, Université Paris-Sud, Orsay, France, 2004.
J. van der Hoeven. Transseries and real differential algebra, volume 1888 of Lecture Notes in Mathematics. Springer-Verlag, 2006.
M. van Hoeij. Formal solutions and factorization of differential operators with power series coefficients. JSC, 24:1–30, 1997.