|
Consider an effective differential ring
|
There exist essentially two approaches in computer algebra for the computation with formal solutions of systems of ordinary or partial differential equations.
The first approach is based on the theory of differential algebra [Rit50, Sei56, Ros59, Kap57,
Kol73, Bui92]. We start with an effective
differential field with derivations
and enrich
with the formal
solution to a system
of differential
polynomials. From an algebraic point of view, this formal solution lies
in the differential quotient ring
of
by the perfect ideal
generated by
. The theory of differential
algebra is well-established and admits an effective counterpart: see [Bou94] for an overview. In particular, there exists a
zero-test in
.
In practice however, a simple function such as
is not just the solution to the differential equation
, but it also satisfies the initial condition
. When taking into account
initial conditions, non-zero elements in
may
vanish when we consider them as functions. For instance, assume that we
redefine
to be the unique solution to
with
and
. Then
is non-zero as an
element of
, even though
vanishes as a function.
In certain cases, it may therefore be necessary to consider a second approach, in which the new functions are defined to be solutions to systems of differential equations with initial conditions. Since most functions which arise in this way are analytic, it is natural to systematically work with computable power series.
More precisely, given an effective field ,
a multivariate power series
is said to be
computable, if there exists an algorithm which takes
on input and which outputs the corresponding coefficient
of
in
. The set
of such
series forms a differential ring, even though
does not come with a zero-test.
Now assume that we are given an effective subring
of
and let
be the
quotient field of
. We wish
to extend
with power series solutions
to systems of partial differential equations
. In the ordinary differential case, the
coefficients of
can generally be computed as a
function of the equations
and a finite number of
initial conditions in
. In
the case of partial differential equations, the initial conditions are
taken in
with
.
When extending with the solution to a system of
differential equations which satisfies explicit initial conditions, an
important problem is to decide whether
for a
given differential polynomial
.
In the case of ordinary differential equations, there exists a variety
of algorithms to solve this problem [DL84, DL89,
Sha89, Sha93, PG95, PG97,
vdH01, vdHS06, vdH02]; see [vdH02] for a discussion. Progress in the case of partial
differential equations has been slower and the best currently known
result [PG02] reduces the zero-test problem to the Ritt
problem concerning the distribution of the singular zeros of a
differential system among its irreducible components. In fact, questions
related to the zero-test problem quickly tend to be undecidable:
has a power series solution in
.
In this paper, we propose a zero-test in the partial differential case,
which is based on a generalization of the technique from [vdH02,
vdHS06]. In a nutshell, assume that
is the unique solution to
with
and explicit initial conditions. In order to test whether
for
, we first
simplify the system of equations
using the
theory of differential algebra. At the end, we obtain a new system of
equations
. If we can show
that this system admits a solution
which
satisfies the same initial conditions as
,
then the fact that
implies
and
. If
, then it suffices to expand
sufficiently far in order to find a non-zero coefficient.
The main difficulty is to ensure the existence of a suitable solution
to the reduced system
, which may be highly singular. The main part of
this paper is devoted to further reduce the system
together with the initial conditions into an asymptotic normal form
(called an asymptotic basis). If this asymptotic normal form is
non-contradictory, then we will guarantee the existence of a solution in
a suitable field
of power series with
logarithmic coefficients. In order to make our zero-test work, we thus
have to assume that
is also the unique solution
in
to
with the
prescribed initial conditions. Fortunately, this is only a mild
additional hypothesis.
Let us outline the structure of the paper. In section 2, we start with some quick reminders on so called grid-based power series. For a more detailed treatment, we refer to [vdH06, Chapter 2].
In section 3, we review the elementary theory of
differential algebra. In order to lighten notations, we will only
consider equations in one differential indeterminate. In section 3.6, we also introduce Ritt co-reduction in the “dual
setting” where differential operators in
are replaced by operators in
.
This dual setting was considered before in the linear case [vdH07].
In section 4, we next switch to the case when we work over
a field of grid-based power series
with a suitable monomial ordering on
. This field comes with natural valuation
preserving derivations
. More
generally, for suitable
, we
will consider more general derivations
,
where
is an infinitesimal monomial. Since
is a field of series, it is naturally to study the
dominant parts of differential equations, which gives rise to a theory
of “asymptotic differential algebra”.
Although a general theory of asymptotic differential algebra is somewhat
difficult to design due to the possible presence of infinite powers of
initials and separants in the asymptotic analogue of Ritt reduction, a
systematic theory can be developed in the case of “quasi-linear
differential ideals”. We will define and outline the basic
properties of such ideals in section 5 and prove the
existence of zeros in in the case when the
derivations are
.
Starting with the equations ,
we next have to put the equations in quasi-linear form. This is
automatic for the derivations
if
is sufficiently small. Using a suitable process of
asymptotic dualizations and changes of derivations, we will show in
section 6 how to end up with a system of quasi-linear
equations w.r.t.
.
In section 7, we will give algorithmic counterparts of this
construction and show how this can be used to elaborate our zero-test.
When applying the algorithm from this paper in the ordinary differential case, we notice a small difference with [vdH02, vdHS06]. In the present paper, we expand sufficiently far so as to make the equations in the reduced system quasi-linear. This simplifies the existence proof of solutions, but may require expansions up to a higher order than what was necessary before. Indeed, the previous algorithms relied on a theoretical result [vdH02, Theorem 3] for the existence proof (see also [vdH06, Chapter 8]).
For convenience of the reader, we will briefly recall the definition of grid-based power series, as well as some notational conventions. For a more detailed exposition, we refer to [vdH06, Chapter 2].
Let be an arbitrary ring of coefficients (
is usually a field) and
a
totally ordered multiplicative group (one may also consider partially
ordered monoids). The ordering
on
is called an asymptotic dominance relation and we
call
a monomial group. A subset
is said to be grid-based if
for some infinitesimal monomials
and an
arbitrary monomial
. Here
for any
.
A grid-based series is a formal sum
with
and grid-based support
. For reasons which will become clear later in
the paper (when
will be allowed to be a
non-commutative ring of operators), we multiplied
on the right with its corresponding coefficient
. The set
of grid-based
series forms a ring (and a field if
is a field).
Any non-zero grid-based series admits a unique
dominant monomial
(the maximal element
in its support) and a corresponding dominant coefficient
. By convention,
. The dominance relation
may be extended to
by setting
if
or
and
and
. This
notation further extends to the case when
and
for different rings
and
. We also write
if
and
if
but
.
Sometimes, it is convenient to use Landau's notation and write
or
instead of
resp.
.
A family is said to be a grid-based or
summable, if
is grid-based and
is finite for each
(notice
the double index convention
).
In that case, its sum
with
is again in
.
We assume that the reader is familiar with basic results from
differential algebra. Let be a field with a
finite number of derivations
.
We assume that the space
is closed under the Lie
bracket; the
are not required to commute. We
will denote
Let be fixed
-linearly
independent numbers in
with
![]() |
(1) |
It is easy to construct such :
starting with any vector
of
-linearly independent real numbers, it suffices
to produce a sufficiently precise rational approximation
of
and take
. We define a partial ordering
and a total ordering
on
by
The total ordering is admissible in Ritt's
sense.
We will restrict our attention to the ring of
differential polynomials in one single indeterminate
. Given
,
we will write
for its leader,
for its initial and
for its separant. The
relation (1) ensures that
for all
. Let
a vector of differential polynomials and denote
Given , Ritt reduction of
w.r.t.
yields a relation
![]() |
(2) |
where ,
is reduced w.r.t.
and
is a vector of linear differential operators. We
understand that
. In
particular, if
is a Rosenfeld basis
(i.e. if
is a coherent auto-reduced
set), then Ritt reduction of the
-polynomial
of
and
(
) gives rise to a relation
![]() |
(3) |
We will call (3) a critical relation. It can be
rewritten in the form .
Given a and
,
there exist unique differential polynomials
, called additive and multiplicative
conjugates of
, with the
properties that
for all . These notions
naturally extend to vectors and operators
;
for instance,
. We have
Consequently, if Ritt reduces to
modulo
, as in
(2), then
and
Ritt reduce to
and
modulo
resp.
, with
In particular, if is a Rosenfeld basis with
critical relations
, then so
are
and
,
with critical relations
resp.
.
Another important construction is to change the derivations. Assume that
for certain . Then any
differential polynomial
can be rewritten into a
differential polynomial
with respect to the
derivations
and vice versa. Similarly,
any operator
can be rewritten as an operator in
. We again have
Consequently, if Ritt reduces to
modulo
, as in
(2), then
Ritt reduces to
modulo
, with
In particular, if is a Rosenfeld basis with
critical relations
, then so
are
, with critical relations
.
Let be the set of linear homogeneous
differential polynomials. A differential ideal
is said to be linear, if it is generated by
. Given
,
we will denote by
and
its linear and constant parts. We also denote
.
Given , its initial and
separant
are equal and invertible. Ritt
reduction of
with respect to a system
yields a relation
![]() |
(4) |
with and
.
Similarly, the
-polynomial of
two polynomials in
is again in
. In fact,
as a
-module is isomorphic to the
skew-ring
. This leads to the
alternative interpretation of (4) as the reduction of
w.r.t.
in the
theory of non-commutative Groebner bases. Similarly,
-polynomials correspond to
-polynomials and Rosenfeld bases to Groebner
bases.
Applying Buchberger's algorithm to a linear system , we thus obtain a Rosenfeld basis
with
. Any
linear differential ideal
may be represented as
for such a basis
.
From now on, we will freely use the concepts of Ritt reduction, critical
relations, etc. for operators in . Given a Rosenfeld basis
and
, let
be the critical relation for
and
, with
.
Let
be the matrix whose rows are the vectors
. It is well-known that these
rows generate the module of
with
. In other words, we have a natural exact
sequence
![]() |
(5) |
where
Assuming that is a field of constants for
, we can also consider the natural
power series analogue
of
, and apply the theory of standard bases. It will be
convenient to regard this as a dual theory.
More precisely, given , we
will call the smallest variable
occurring in
its co-leader
. The corresponding coefficient is called the
co-initial
or co-separant
. Ritt co-reduction
of
w.r.t.
yields a relation
![]() |
(6) |
with and
such that
for all
.
If
is a Rosenfeld co-basis, then Ritt
co-reduction of the co-
-polynomial
of
and
with
yields a co-critical relation
with . Denoting by
the matrix of co-critical relations, we again have a
natural exact sequence
The process of Ritt co-reduction can be generalized slightly beyond the
linear case. More precisely, let be the set of
series
in the infinite number of variables
, such that
is polynomial in each particular variable in
. Now let
be a linear
system. We say that
is co-reduced
w.r.t.
if its co-leader
does not satisfy
for any
. In the contrary case, we may
consider
as a polynomial
in
and write
![]() |
(7) |
for the unique with
, with
.
After a finite number of such steps, we obtain a relation
![]() |
(8) |
where does no longer depend on
and
. Continuing the same
process on
as long as
is
not reduced w.r.t.
results in an
infinite sequence of eliminations which converges to a relation
![]() |
(9) |
where and
is co-reduced
w.r.t.
.
Given formal indeterminates ,
let
denote the valuation-preserving partial
derivatives
Let and let
be a vector
of
-linearly independent
positive real numbers. We give
the structure of
a monomial group by
Notice that is isomorphic to an additive
subgroup of
. From now on, we
will assume that
is a field of grid-based series
of the form
where is a differential field for the
derivations
.
Remark by any
subgroup of
which contains
and for any dominance relation
on
with
. However,
proofs by induction over grid-based supports in
generally have to be replaced by transfinite induction proofs.
Given in
and
in
, we will
have to consider generalized Newton slopes between terms
and
. Both
terms can be “equalized” modulo the change of derivations
, where
If , then the monomials
are all infinitesimal, and we have
![]() |
(10) |
for all . In other words,
although the
do not commute, they do commute
from the asymptotic point of view. Similarly, if
, then each derivation
asymptotically commutes with elements
in the
operator ring
:
![]() |
(11) |
Let be as before. Any differential polynomial
can also be considered as a series
in the ring
Similarly, an operator can be considered as a
series
in the ring
Recall our convention to multiply monomials in the series on the right with the corresponding coefficients.
Because of (10), the set is not
necessarily stable under differentiation. It will be convenient to
introduce the commutative differential ring
obtained through formal substitution of
by
pairwise commuting variants
.
The corresponding bijection
between
and
, with
inverse
, satisfies
for all and
.
If
, then (11)
implies that
is a field of constants for the
derivations
. Given
, we will denote
. This notation is extended to differential
ideals
of
using
The set is clearly an ideal of
. Assume that
.
Given
, there exists a
with
.
Replacing
by
,
we may assume without loss of generality that
, whence
with
. Now we have
and
for all
,
whence
and
.
This proves that
is a differential ideal, called
the dominant differential ideal of
.
Remark is a radical differential ideal, then
is not necessarily radical. For instance, if we take
,
and
, then
.
A differential ideal of
is said to be strong, if it is closed under grid-based
summation. Since
is archimedian, the strong
differential ideal
generated by a differential
ideal
is just the completion of
:
![]() |
(12) |
It follows that . Any strong
differential ideal
of
is
clearly a
-module. The
contrary is not always true: taking
,
the
-module generated by
,
,
,
does not contain
.
be a
-module such that
is
finitely generated. Then
is a strong
differential ideal.
Proof. Let with
be such that
is generated by
. Given
, we claim that there exists an operator
with
such that . We compute the
coefficients
of
by
induction over
. Assume that
with
has been computed.
Then
and
.
Since
is generated by
, there exists an operator
with
. Taking
, it follows that
,
so the induction hypothesis is satisfied for the next monomial
in
. By
induction, we thus compute an operator
which
satisfies
for all
.
It follows that
.
Having shown our claim, let be a grid-based
family. Then there exists a family of operators
with
and
for all
. For each
, and using the facts that
is finite and
is grid-based, there are only a
finite number of indices
with
. Consequently,
is a
grid-based family and
.
Remark -module is necessarily
finitely generated as well.
Let us consider a vector with
for each
. A series
is said to be asymptotically reduced
w.r.t.
if
is Ritt reduced w.r.t.
.
Due to the presence of in (2), a
similar division technique as for standard bases will not always work in
the case of asymptotic reduction: infinite sequences of partial
reductions may give rise to infinite powers of initials and separants.
Nevertheless, we will now show that the mechanism does work in the
special but important case when all initials and separants are
.
We say that is normal if
and
for all
. Given
,
we claim that there exists an expression
![]() |
(13) |
where satisfies
and
is reduced w.r.t.
. We will call
and
the quotient and remainder of the asymptotic
reduction of
w.r.t.
. The coefficients of
and
are computed by induction over
the grid-based set
So let and assume that we computed
and
for all
, in such a way that
![]() |
(14) |
with and
.
We may assume without loss of generality that
(if
, then we may simply take
). Ritt reduction of
w.r.t.
yields a
relation
with and
.
In view of (10) and (11), we get
![]() |
(15) |
with ,
and
with
.
Taking
, it follows that
If , then
is reduced w.r.t.
and the
construction stops. Otherwise, we have
and
and the induction hypothesis is satisfied for the
next monomial
.
We will say that is quasi-normal if
for all
.
In that case, let
for each
, where
is the leader
of
and
the coefficient
of
in
.
Then
is normal and we call it the
normalization of
.
Remark is not quasi-normal, then the same argument still leads to
a relation
with whenever the reduction process stops.
However, we no longer obtain a nice relation in the case when
reduces to
.
As we did in section 3.6 for Ritt reduction, it is natural
to introduce asymptotic co-reduction, the dual notion of asymptotic
reduction. Since cannot be applied to
, we will assume
. The set
plays the role of in the dual setting.
Similarly, operator algebra
is the natural counterpart of .
In a similar way as before, one introduces the commutative variant
of
.
Again, the dominant part
of a differential ideal
is a differential ideal.
Any series can also be regarded as a series in
the variables
and
,
with a support
which is a subset of
. A subset
is said to
be admissible if it occurs as the support of a series in
. A family
is said to
be summable if
is admissible and
is finite for each
.
In that case,
. A
differential ideal
of
is
said to be strong if it is closed under infinite summation.
Let be the set of series in
which are strong linear combinations of monomials
of weight
at least
.
The strong differential ideal
generated by a
differential ideal
coincides with a suitable
completion of
:
![]() |
(16) |
It follows that . Any strong
differential ideal of
is a
-module. Any
-module
, such that
is strong and finitely generated as a strong differential ideal, is a
strong differential ideal.
Consider a vector with
for all
. We will say that
is co-normal, if
,
and
is linear for all
. Given
, we say that
is
asymptotically co-reduced w.r.t.
if
or
for all
. In a similar way as above,
asymptotic co-reduction of
w.r.t.
yields a relation
where satisfies
and
is co-reduced w.r.t.
.
Let be normal in the sense of section 4.3
and assume that
is linear for each
. We say that
is an
asymptotic basis, if
is a Rosenfeld
basis and if each pair
with
satisfies an asymptotic critical relation
![]() |
(17) |
where and
are such that
is a critical relation for
and
. We call
the asymptotic
-polynomial
of
and
.
We will denote by
the matrix formed by the
row vectors
,
such that
is an asymptotic critical relation.
More generally, we say that is an asymptotic
basis, if
is quasi-normal and if its
normalization is an asymptotic basis in the above sense.
Proof. We will construct
by induction over the grid-based set
Let be such that
has
been computed. As the induction hypothesis, assume that
,
and
Since , we have
. Hence, the exactness of (5)
implies the existence of a vector
with entries
in
, such that
. Taking
,
we have
and . By induction over
, we obtain a vector
with
.
be a normal asymptotic basis. Then
the strong differential ideal generated by
is
given by
and coincides with the set of
which reduce asymptotically to
w.r.t.
.
Proof. Given with
, the lemma implies that there
exists an operator
with
and
. In particular,
, so that
, by proposition 4. Furthermore,
implies that
is not
asymptotically reduced w.r.t.
.
Now let
be the remainder of the asymptotic
reduction of
w.r.t.
. Since
,
the above argument shows that
.
A differential ideal of
is said to be quasi-linear if
is
linear.
be a strong
differential ideal of
.
Then the following conditions are equivalent:
is quasi-linear.
admits a normal asymptotic basis.
Proof. Assume that is
quasi-linear. By definition, there exists
with
,
,
for each
, and such that
is a
Rosenfeld basis. Given
, we
claim that asymptotic Ritt reduction of the asymptotic
-polynomial of
and
yields an asymptotic critical relation (17).
Indeed, the remainder
of the asymptotic
reduction must vanish, since
and
is reduced w.r.t.
. We thus get a relation (17).
Furthermore, Ritt reduction of
w.r.t.
yields
, which is thereby a critical relation for
and
.
Inversely, assume that admits a normal
asymptotic basis
. Given
, corollary 9
implies that
reduces to
modulo
. In particular,
and
is a linear differential
ideal.
It is rather straightforward to dualize the theory from the previous section. Let us briefly state the dual versions of the main results, while omitting proofs.
Let be co-normal in the sense of section 4.4. In particular,
is linear for each
. We say that
is an asymptotic co-basis, if
is a Rosenfeld co-basis and if for each pair
, we have an asymptotic co-critical
relation
![]() |
(18) |
where and
are such that
is a co-critical relation for
and
. We will denote by
the matrix formed by the
row
vectors
, such that
is an asymptotic critical relation. A differential ideal
of
is said to be
quasi-linear if
is linear.
be a co-normal
asymptotic co-basis. Then
coincides with the
set of
which co-reduce asymptotically to
w.r.t.
.
be a strong
differential ideal of
.
Then the following conditions are equivalent:
is quasi-linear.
admits a co-normal asymptotic co-basis.
In this section, we consider an asymptotic basis
in the special case when
and
is a field of constants for
.
We will show the existence of solutions to the equation
in the ring
, where
.
Proof. Applying the tangent cone algorithm to
, we obtain a matrix
with coefficients in
such
that
is a Rosenfeld co-basis. We claim that
is compatible with
in the
sense of [vdH07, Section 3.2], i.e. that
for any
with
. Indeed, let
be the
matrix of critical relations for the Rosenfeld basis
. Since
is a flat ring
over
, the exact sequence
transforms into an exact sequence
In other words, given with
, there exists a
with
. It follows that
In [vdH07, Section 4.2], we have shown how to compute a
solution in to
Now is in the strong ideal generated by
, so there exists a matrix
with coefficients in
with
. Consequently, there exists
a matrix
with entries in
and
. We conclude that
and
.
Proof. We may assume that
is normal. We compute the coefficients
of a
solution
with
by
induction over
. We have
. Given
, assume that
has been
computed for all
, in such a
way that
for
.
We have
and each asymptotic critical relation
gives rise to a relation
Extracting dominant parts, it follows that is a
Rosenfeld basis. Moreover,
,
whence
. By lemma 14,
it follows that
admits a solution
. Taking
and
, we have
and
, so the induction
hypothesis is verified for the next monomial in
.
be a Rosenfeld basis w.r.t.
and let
be the leader of
for each
.
Assume that
and
for
all
. Consider the critical
relation
![]() |
(19) |
obtained by Ritt reduction of modulo
. Then
. Moreover, if
for each
, then
is an asymptotic basis.
Proof. Let .
The process of Ritt reduction yields a sequence of relations
![]() |
(20) |
obtained by pseudo-division of by
, considered as polynomials in the common
leader
of
and
. In particular,
is free from
. At the end of
the reduction process, we have
for some
.
Let us proof by induction over that
and thus
. From
, we deduce
. Let
be the
coefficient of
in
.
Then
is also the coefficient of
in
, since
is free from
. In particular,
. Telescoping the relations
(20) for
, we
get
.
The hypothesis implies in particular that
. If we also have
for each
, then
, whence (19) can be divided by
and rewritten into an asymptotic critical
relation.
be a Rosenfeld basis
w.r.t.
and assume that
satisfies
,
but
. Then there exist
and
,
such that
is an asymptotic basis
w.r.t.
.
Proof. Consider the Rosenfeld basis w.r.t.
,
which satisfies
for all
. Let
be the leader of
for each
.
Since
, the coefficient
of
in
does not vanish. Taking
sufficiently small, we
have
for all
.
From now on, let us consider the entries of
and
as differential polynomials in
. Modulo a multiplicative conjugation by a
sufficiently small
, we may
assume without loss of generality that
for all
, while preserving the fact
that
for all
.
Modulo a division of each
by
, we are thus in a position where
satisfies the conditions of lemma 16. We
conclude that
is an asymptotic basis
w.r.t.
, and so
is
for any
.
In this section, we assume that .
Given a strong differential ideal
,
the dualization of
is the smallest
strong differential ideal
which contains
.
,
its dualization is again quasi-linear and we have
.
Proof. Let be an
asymptotic basis for
and let
be the vector of critical relations for
,
giving rise to an exact sequence
![]() |
(21) |
Since is a flat ring over
, we also get an exact sequence
![]() |
(22) |
Now consider an element with
. We claim that we may rewrite
in such a way that
. Indeed,
using the exactness of (22) instead of the exactness of (5), this is proved in a similar way as lemma 7.
Our claim implies that
. This
shows that
.
Now consider a matrix with entries in
such that
is normal and
is a Rosenfeld co-basis. We claim that
is an asymptotic basis for
.
Indeed, as in the first part of the proof of proposition 10,
asymptotic co-reduction of the asymptotic co-
-polynomial of
and
with
yields an asymptotic
co-critical relation for
and
. Our claim and proposition 12
imply that
is quasi-linear. We conclude that
.
is a strong quasi-linear differential ideal,
then so is
.
Proof. Assume that is
generated by
. Then
is again generated by
.
We will now study what happens to quasi-linear differential ideals when changing the derivations. Given two systems of derivations
with we will write
if
. In that case, we have
where and
denote the
natural counterparts of
and
for
. More generally, we will
use primes when working with respect to
.
From now on, let be an asymptotic co-basis. For
each
, let
be the co-leader of
. The
largest
for which
is not
reduced to a single term for some
,
is called the next critical derivation. If
is also quasi-normal w.r.t.
,
then it follows in particular that
is the leader
of
for each
.
be an asymptotic co-basis w.r.t.
and let
be the next
critical derivation. Assume that
is linear for
each
. Then
is also an asymptotic basis w.r.t.
.
Proof. Without loss of generality, we may assume
that is co-normal w.r.t.
. We will denote by
the normalization of
w.r.t.
. Let us
prove that any element
asymptotically reduces to
w.r.t.
. In particular, this will imply that the asymptotic
-polynomials of the form
asymptotically reduce to
w.r.t.
, and
thereby give rise to the required asymptotic critical relations for
.
Assume for contradiction that there exists a , such that
is reduced
w.r.t.
. We may
even assume that
is totally reduced
w.r.t.
,
i.e. that each variable
with
for some
has been eliminated
from
. Since
is an asymptotic co-basis w.r.t.
, asymptotic co-reduction of
yields a relation
. Let us
prove by induction over
that
. This yields the desired contradiction, since
this would imply
, by strong
linearity.
Let . Then
is the quotient of the co-reduction of
w.r.t.
.
Consequently, starting with
,
there exists an infinite sequence of equations
each of which is induced by a partial co-reduction of the type (7), and such that
More precisely, let be the co-leader of
and write
.
Then there exists an index
with
,
and
for
. Assuming that
, we must have
since
is totally reduced w.r.t.
and
involves
. Combined with the fact that
, this yields
and
. In other words, our
hypothesis
implies
for
all
, whence
.
Remark w.r.t.
obtained by asymptotic co-reduction of the
asymptotic co-
-polynomials
continue to provide asymptotic critical relations for
w.r.t.
which are sufficiently close
to
. Using lemma 10,
it follows that
is an asymptotic basis
w.r.t.
. If
, then asymptotic reduction
of the asymptotic
-polynomials
w.r.t.
yield new asymptotic
critical and co-critical relations w.r.t.
, which allows us to continue the process with
instead of
.
It can be shown that
is reached after a finite
number of steps.
Given an effective ring , we
recall that a multivariate power series
is
computable, if there exists an algorithm which takes
on input and which outputs the corresponding coefficient
of
in
. The set
of such
series forms a differential ring. Although
is
usually a field, it is also allowed to take a non-commutative ring of
operators for
.
Given an effective monomial group ,
a grid-based series
is said to be
computable if there exist infinitesimal
and
, as well as a computable
series
, such that
. We denote by
the ring of such series.
The definition of computable power series applies in the obvious way to
differential operators . More
generally, a series
is said to be computable, if
there exists an algorithm which takes a monomial
on input and which outputs the corresponding coefficient
. We denote by
the ring
of such series.
Throughout this section, we will assume that and
as in sections 3.1 and 4.1
have been fixed and that
is an effective field
with an effective zero-test. For instance, we may take
and
.
Assume that is an effective field with an
effective zero-test. Let
,
and
.
Consider a Rosenfeld basis
,
such that the leaders of the
are known (remind
that we do not have a zero-test in
).
Then we may apply the theory from section 6 in order to
find an asymptotic basis for
w.r.t.
.
Algorithm asymptotic_basis
Input: a Rosenfeld basis with
known leaders.
Output: an asymptotic basis for .
Step 1. [Initial asymptotic basis]
Let be the leader of
for each
.
Take sufficiently large, such that
for all
and
.
By lemma 16, is an asymptotic
basis w.r.t.
.
Go to step 4.
Step 2. [Dualization]
Normalize .
Compute with respect to
.
Using the tangent cone algorithm, find a Rosenfeld co-basis for
.
For each , let
be such that
.
Replace by
.
By lemma 18, is an asymptotic
co-basis w.r.t.
.
Step 3. [Change derivations]
Compute the next critical derivation and
replace
by
.
By lemma 20, is an asymptotic
basis w.r.t.
.
Step 4. [Done?]
If , then return
.
Otherwise, go to step 2.
Remark . For
instance, in step
, it
suffices to compute the coefficient
up to the
order of
: if
, then
for all
. Similarly, for the computation of
the next critical derivation
in step 3, it
suffices to evaluate
up to the order of
, where
stands for the co-leader of
.
Proof. The correctness being ensured by lemmas
16, 18 and 20, it suffices to
prove the termination. Let and
be the successive values of
and
at the start of step 2. For each
and
, let
be
the leader of
, and consider
the anti-chain
. Then
forms a decreasing sequence of anti-chains, which
implies the finiteness of the sequence.
The argument from the proof of proposition 17 can be
generalized so as to provide asymptotic bases w.r.t. . More precisely, let us consider a
Rosenfeld basis
with
and
whose leaders are known. If
is quasi-linear for
a sufficiently small
, then
we will show how to compute such an
and an
asymptotic basis
. If the
algorithm fails, then we will provide a proof that
.
Algorithm ql_asymptotic_basis
Input: a Rosenfeld basis with
and known leaders.
Output: | an asymptotic basis for ![]() ![]() |
fail and a certificate that ![]() |
Step 1. [Initial asymptotic basis]
Let be the leader of
for each
.
Take sufficiently large, such that
for all
and
.
Replace by
,
where
is sufficiently small such that
is affine.
Go to step 4.
Step 2. [Dualization]
Normalize .
Compute with respect to
.
Using the tangent cone algorithm, find a Rosenfeld co-basis for
.
For each , let
be such that
.
Replace by
.
Step 3. [Change derivations]
Compute the next critical derivation and
replace
by
.
Replace by
,
where
is sufficiently small such that
is affine.
Step 4. [Done?]
If , then return
fail.
If , then return
.
Otherwise, go to step 2.
Proof. The algorithm is a perturbation of the
algorithm , in which case the previous correction and
termination proof generalizes. Whenever
in step
4, this implies in particular that
for the input
basis
.
Until now, we have not really used the fact that the derivations are
allowed to act in a non-trivial way on .
In the next section, we will work over
instead
of
. This requires some
adaptations of proposition 15.
be a field of
constants for
and consider an asymptotic basis
. Then
admits a solution in
.
Proof. This is proved using a straightforward
adaptation of the proof of proposition 15.
be a field of
constants for
and consider a Rosenfeld basis
. Then
admits a solution in
.
Proof. Using a procedure similar to ql_asymptotic_basis,
we may compute a monomial in
such that
is quasi-linear w.r.t.
. Indeed, since
admits no non-linear terms, instead of choosing
such that
is affine in steps 1 and
3, we may now choose
sufficiently large such
that
is linear. By proposition 25,
admits a zero
in
. It follows that
is a solution to
.
Assume now that we wish to work over instead of
. This fits in our setting by
taking
for a constant field
and a copy
of
with
on
. The
variables
and
will
simply be ignored.
be a field of
constants for
. Denote
and
.
Consider an asymptotic basis
.
Then there exists an
with
and
as a series in
.
Proof. The proof is analogous to the proof of
proposition 15. We first notice that the absence of the
variables from
implies
their absence in the asymptotic critical relations and during all
computations in the proof. This time, the Rosenfeld basis
lies in
instead of
. By proposition 26, the equation
admits a solution in
. Since
does not occur in
(regarding
and
as separate variables), the constant term
of
w.r.t.
is again a zero of
. This
ensures that the analogue of the proof of proposition 15
goes through.
Let be an effective field with an effective
zero-test. Assume that we are given an effective differential subring
of
for
with an effective zero-test. Assume also that the coefficients of
as a series in
are still in
and that there exists an algorithm to compute
them. These coefficients lie in the effective differential ring
. We will denote by
the quotient field of
.
In view of the previous section, we may apply the theory from this paper
for series in the field
,
where
.
Let be a partial differential equation with a
solution
and coefficients
. Assume that there exists an order
such that the equation
![]() |
(23) |
admits only as a solution in
. This is typically the case when
is sufficiently non-singular, so that the coefficients of
in
are given by a
recurrence relation. The differential ring
is
clearly a subring of
. We
will now give an effective zero-test in this ring.
Algorithm zero_test
Input: a polynomial .
Output: true if
and false otherwise.
Step 1. [Initial order]
Set and
.
Step 2. [Rosenfeld basis]
If then return false.
Compute a Rosenfeld basis for
.
Let be the subset of
with
.
If then set
and repeat
step 2.
Step 3. [Certification]
If ql_asymptotic_basis()
does not return fail, then return
true.
Step 4. [Increase order]
Set and
.
Return to step 2.
Proof. Let us first prove the correctness. If
the algorithm returns false, then we clearly have . If the algorithm returns
true, then the ideal
is
quasi-linear in step 3. By proposition 27, this implies the
existence of a zero
with
. Since
,
any solution of
is also a solution of
and of
. It
follows that
for
.
But we assumed that
is the only solution to (23). It follows that
.
Let us now prove the termination. For a fixed order , the loop in step 2 terminates because the
successive rankings of
are lower and lower. If
, then we clearly have
for a sufficiently large
. If
,
then, for a sufficiently large
,
we have
for all
considered in step 2. Consequently, the successive values of
in step
all satisfy
. When entering step 3, we therefore have
, whence ql_asymptotic_basis
does not return fail.
Several remarks can be made about the generality of our zero-test. In order to avoid unnecessary complications of the notations, we have presented our algorithm in the case when we adjoin a single solution to a partial differential equation with initial conditions. Of course, the theory of differential algebra also works for differential polynomials in a finite number of indeterminates. It should be straightforward to generalize our zero-test to the case when we adjoin several solutions at the same time.
Similarly, we may replace the single equation by
a system of equations
. In
that case, the initial conditions generally lay in the ring
for some
. In
particular, this requires to take
,
where
is the quotient field of
.
We assumed that is the unique solution in
to (23). This condition is usually met
in practice and in particular when Cauchy-Kovalevskaya's theorem
applies. Nevertheless, the condition can probably be weakened to the
existence of a unique solution in
or some ring
in between. In order to prove such a result, one would have to carefully
study the supports of the series which arise during our computations. We
notice that the existence of a unique solution in
is a minimal requirement. However, in view of results such as theorem 1, it might be hard to escape from the intrusion of
logarithms. As a matter of fact, the proof of this theorem fails if we
replace
by
.
It is appropriate to notice one important advantage of our setting: we
only required to be an effective subring of
with an effective zero-test. Such rings can be
constructed via a tower of extensions of
of the
type considered in section 7.4. However,
does not necessarily have to be of this type. For instance, in the case
of ordinary differential equations, we know that the series
in
occurring in Stirling's series
for
is differentially transcendental over
. We may thus take
.
Even though the proof of the new zero-test might seem complex, we notice
that the actual algorithm is actually quite simple and can expected to
be reasonably efficient. Indeed, apart from the usual (inevitable?) step
of computing Rosenfeld bases over ,
we only have to apply the tangent cone algorithm a few times for series
over the simpler field
.
Several points of a more theoretical interest might deserve further
study. First of all, we made a few artificial hypothesis on and
, which one
may try to remove as much as possible. More importantly, it would be
nice to have a genuine theory of asymptotic differential algebra, which
goes beyond the quasi-linear case. In the continuation of [AvdD04],
such a theory might even work for more general “asymptotic
functions fields” than our fields of grid-based series
.
M. Aschenbrenner and L. van den Dries. Asymptotic differential algebra. In Contemporary Mathematics. Amer. Math. Soc., Providence, RI, 2004. To appear.
F. Boulier. Étude et implantation de quelques algorithmes en algèbre différentielle. PhD thesis, University of Lille I, 1994.
A. Buium. Differential algebraic groups of finite dimension, volume 1506 of Lecture Notes in Mathematics. Springer-Verlag, 1992.
J. Denef and L. Lipshitz. Power series solutions of algebraic differential equations. Math. Ann., 267:213–238, 1984.
J. Denef and L. Lipshitz. Decision problems for differential equations. The Journ. of Symb. Logic, 54(3):941–950, 1989.
I. Kaplansky. An introduction to differential algebra. Hermann, 1957.
E.R. Kolchin. Differential algebra and algebraic groups. Academic Press, New York, 1973.
A. Péladan-Germa. Testing identities of series defined by algebraic partial differential equations. In G. Cohen, M. Giusti, and T. Mora, editors, Proc. of AAECC-11, volume 948 of Lect. Notes in Comp. Science, pages 393–407, Paris, 1995. Springer.
A. Péladan-Germa. Tests effectifs de nullité dans des extensions d'anneaux différentiels. PhD thesis, Gage, École Polytechnique, Palaiseau, France, 1997.
A. Péladan-Germa. Testing equality in differential ring extensions defined by pde's and limit conditions. AAECC, 13(4):257–288, 2002.
J.F. Ritt. Differential algebra. Amer. Math. Soc., New York, 1950.
A. Rosenfeld. Specializations in differential algebra. Trans. Amer. Math. Soc., 90:394–407, 1959.
A. Seidenberg. An elimination theorem for differential algebra. Univ. California Publ. Math. (N.S.), pages 31–38, 1956.
J. Shackell. A differential-equations approach to functional equivalence. In Proc. ISSAC '89, pages 7–10, Portland, Oregon, A.C.M., New York, 1989. ACM Press.
J. Shackell. Zero equivalence in function fields defined by differential equations. Proc. of the A.M.S., 336(1):151–172, 1993.
J. van der Hoeven. D-algebraic power series. Technical Report 2001-61, Prépublications d'Orsay, 2001.
J. van der Hoeven. A new zero-test for formal power series. In Teo Mora, editor, Proc. ISSAC '02, pages 117–122, Lille, France, July 2002.
J. van der Hoeven. Transseries and real differential algebra, volume 1888 of Lecture Notes in Mathematics. Springer-Verlag, 2006.
Joris van der Hoeven. Generalized power series solutions to linear partial differential equations. JSC, 42(8):771–791, 2007.
J. van der Hoeven and J.R. Shackell. Complexity bounds for zero-test algorithms. JSC, 41:1004–1020, 2006.