Complex transseries solutions to
algebraic differential equations

by Joris van der Hoeven

Dépt. de Mathématiques (bât. 425)
Université Paris-Sud
91405 Orsay CEDEX
France

November 4, 2023

Abstract

In our PhD. we have given an algorithm for the algebraic resolution of algebraic differential equations with real transseries coefficients. Unfortunately, not all equations do admit solutions in this strongly monotonic setting, even though we recently proved an intermediate value theorem.

In this paper we show that the algorithm from our PhD. generalizes to the setting of weakly oscillatory or complex transseries. Modulo a finite number of case separations, we show how to determine the solutions of an arbitrary algebraic differential equation over the complex transseries. We will show that such equations always admit complex transseries solutions. However, the field of complex transseries is not differentially algebraically closed.

1Introduction

In [vdH97], we have studied the asymptotic behaviour of solutions to algebraic differential equations in the setting of strongly monotonic or real transseries. We have given a theoretical algorithm to find all such solutions, which is actually effective for suitable subclasses of transseries. More recently, we have proved the following “differential intermediate value theorem”.

Theorem 1. [vdH00a] Let be the real field of grid-based transseries in and let be a differential polynomial with coefficients in . Then, given transseries with and , there exists a with and .

This theorem implies in particular that any algebraic differential equation of odd degree, such as

has at least one real transseries solution. This theorem is striking in the sense that it suggests the existence of theories of ordered and/or valuated differential algebra.

However, a main drawback of the setting of real transseries, is that not every algebraic differential equation can be solved; actually, even an equation like has no solutions. In order to get a better understanding of the asymptotic behaviour of solutions to algebraic differential equations, it is therefore necessary to search for a complex analogue of the theory of real transseries. This paper is a first contribution in this direction.

The first problem is to actually define complex transseries. The difficulty is that it is not clear a priori whether an expression like should be seen as an infinitely large or an infinitely small transmonomial. Several approaches can be followed. A first approach, based on pointwise algebras, was already described in chapter 6 of [vdH97]. However, this approach has the drawback that it is not easy to compute with complex transseries.

A second more computational approach is described in section 3. Roughly speaking, it is based on the observation that all computations with complex transseries can be done in a similar way as in the real setting, except for testing whether a monomial like is infinitely large or small. Now whenever we have to make such a choice, we will actually consider both cases, by applying the automatic case separation strategy (see [vdH97]). We implicitly reject the case when is bounded, which is “degenerate”, but which deserves to be studied later.

The last approach, which is described in section 2, is more structural and really allows us to define a complex transseries in a not too difficult way. The underlying idea is analogue to the concept of a maximal ideal. Intuitively speaking, we assume the existence of some “god”, who has decided a priori for us which monomials like are infinitely large and which ones are infinitely small. It turns out that all possible choices lead to isomorphic fields of transseries. However, the geometric significance of these fields is hard to grasp.

In section 4, we introduce parameterized complex transseries, which are necessary to express generic solutions to differential equations. Indeed, such solutions may involve integration constants. As usual, our approach is based on the automatic case separation strategy.

The remaining sections deal with the resolution of asymptotic algebraic differential equations with complex transseries coefficients. Our approach is similar to the one followed in [vdH97], but we have made a few simplifications and we corrected an error (see section 9.4). Our main results are stated in sections 9.1 and 9.2. We show that there exists a theoretical algorithm to express the generic solution to an algebraic differential equation by means of parameterized complex transseries and we give a bound for the logarithmic depth of the generic solution. We also show that an algebraic differential equation of degree admits at least complex transseries solutions when counting with multiplicities. As a consequence, each linear differential equation admits a full system of solutions. However, our fields of complex transseries are not differentially algebraically closed and several interesting problems still need to be solved (see section 9.5).

The reader should be aware of a few changes in notations w.r.t. [vdH97], which are summarized in the following table:

2Complex transseries

2.1Real trigonometric fields

In all what follows, let be a real trigonometric field and its complexification. This means that has the structure of a totally ordered field and functions , which are compatible with this ordering.

More precisely, we assume that admits an inverse function with domain and that the function restricted to admits a totally defined inverse. Here , where and . Furthermore,

for all . Finally, for each resp. and , we require that

Proposition 2. The real numbers form a real trigonometric field.

Proof. The functional equations are classical. The inequality for was first proved in [?]. As to the inequality for , we have

if . Otherwise,

since .

Remark 3. Analogue inequalities can be proved for and for expansions at order instead of .

Remark 4. Most of the most classical computations on complex numbers can be carried out in the context of real trigonometric fields. For instance, we the functions and may naturally be extended to , numbers in may naturally be written in polar form, etc.

Remark 5. We also have a natural partial analogue of real trigonometric fields; in this case, the functions and are no longer required to be total and the functional equations resp. inequalities are only required to hold, whenever they make sense. For instance, if , then we require that and .

2.2Series with complex coefficients and monomials

Let be a totally ordered monomial group (or set) with -powers. Then we recall that the field of grid-based power series is naturally totally ordered by , for all . This ordering is compatible with the multiplication: . More generally, if is only partially ordered, then we define an ordering on to be compatible with the asymptotic ordering on , if

(1)

for all .

In what follows, we are rather interested in the complexification of . Obviously, this -algebra can not be given an ordering which is compatible with the multiplication. Nevertheless, it is interesting to consider orderings on which are only compatible with the -algebra structure of . Such an ordering is again said to be compatible with the asymptotic ordering on , if (1) holds for all .

Assuming that such orderings on and are total, the condition (1) implies that for all non zero . Consequently, the ordering on is totally determined by the sets

where runs over . Each is actually the set of strictly positive elements of a total ordering on , which is compatible with the -module structure of . Therefore, each is characterized by an angle and a direction , via

This situation is illustrated in figure 1.

Figure 1. Shape of the region for resp. .

In is also possible to consider complex powers of monomials: a complex monomial group is a monomial group with -powers, with an asymptotic ordering which is compatible with the expo-linear -vector space structure of . For instance the formal group is a monomial group with -powers for the ordering . This group is not totally ordered, since and are incomparable. We may make the ordering total by deciding that .

2.3Pre-fields of complex transseries

Now consider a totally ordered grid-based algebra of the form , where is a totally ordered complex monomial group, and where the ordering on is assumed to be compatible with the asymptotic ordering on . Assume that we also have a partial logarithmic function on , such that

L1

coincides with the usual logarithm on ;

L2

If , then and .

We say that is a pre-field of complex transseries if the following conditions are satisfied:

T1

;

T2

, for all ;

T3

For all , we have , where ,

as well as the following conditions for the logarithm:

L3

for all and ;

L4

for all ;

L5

for all .

In L5, we write , if and only if for all . In view of L3, this means that for all .

Remark 6. The conditions L1 until L5 replace the requirement that should be a logarithmic function in the definition of real fields of transseries. It can indeed be checked that our conditions are equivalent to the usual conditions on in this case.

Remark 7. The domain of the logarithm on may be further extended, by setting for all , where is defined using the function on . Of course, such an extension of the logarithm to involves a choice of a principal determination. Furthermore, such an extension cannot satisfy both the properties L1 and L2.

On the other hand, the partial inverse of may be extended canonically in such a way that the equation admits a solution for each . Indeed, it suffices to extend via for all and . In what follows, we will always assume that the partial inverse of has been extended in this way.

2.4Logarithmic complex transseries in

Consider the formal -vector space generated by the formal symbols . Given angles and directions , we define a total ordering on as explained in section 2.2. Then the formal exponential of is a complex monomial group for the asymptotic ordering defined by , where . In order to avoid confusion, we will sometimes write instead of .

Assume from now on that and were chosen such that . Given a non-zero grid-based series with , we define its logarithm by

We may extend the total ordering on to in a similar way as in section 2.2, by extending the angle and direction families resp. into larger families resp. . It is easily verified that the field with this ordering is a pre-field of complex transseries.

Actually, the structure of does not really depend on the choices of , , and , modulo rotations and conjugations. Indeed, assume that and are a second family of angles and directions with indices in . Then we define an increasing isomorphism between and by

where

for all . We infer that is an isomorphism of complex monomial groups. Now if and are families of angles resp. directions with indices in , and which extend and , then we define an increasing isomorphism between and by

(2)

We notice that extends if and only if , which is again equivalent to the condition that for each we have

In this case, we say that and are strongly compatible. We say that and are compatible if the relation holds for all with sufficiently large .

2.5Complex transseries in

Assume now that we are given a complex field of transseries , which is not stable under exponentiation (modulo the extension of the exponentiation as described in remark 7). Let and be the associated families of angles and directions. Now consider the formal complex monomial group

whose asymptotic ordering is given by , for all . Given extensions and of and to families indexed by monomials in , we may totally order as explained in section 2.2. It is easily verified that is a pre-field of complex transseries, which we call the exponential extension of , relative to and . In cases of confusion, we will write instead of . Notice that the exponential of any series in is defined in .

Again, the structure of does not really depend on the choice of . Indeed, if and are two different such choices, then

(3)

is an increasing isomorphism between and .

Starting with from the previous section, we may now consider the iterated exponential extensions , , of . The union

of these fields is called a field of complex transseries in . Of course, the construction of depends on the successive choices of angles and directions for , with indices in . The angles and directions for coincide with these choices on each . We will write instead of whenever confusion may arise.

We claim that and are isomorphic as soon as the restrictions of and to are compatible. We have already shown (see formulas (2) and (3)) that there exist isomorphisms

for each . Now let be such that for each . Then we observe that for all and . By induction over , it then follows that for for all and . Given , this shows that the value of does not depend on the choice of , for sufficiently large . In other words, the can be glued together into an isomorphism between and .

Remark 8. It is possible to slightly generalize the construction of pre-fields of complex transseries in , when starting with instead of . Notice that is not necessarily a monomial when adopting this generalization.

2.6Fields of complex transseries

Actually, in our construction of pre-fields of complex transseries in , it is reasonable to require that require that for all sufficiently large , thereby eliminating all ambiguity (up to isomorphism) in the construction of . More generally, a pre-field of complex transseries is a field of complex transseries, if it satisfies the following axiom:

T4

For each , there exists an , such that for all we have

Then up to isomorphism, we have constructed the field of grid-based complex transseries in . Actually, the same procedure of exponential extensions and direct limits can be used to close any field of complex transseries under exponentiation. Again, this closure is unique up to isomorphism.

Remark 9. In this paper we restrict our attention to grid-based complex transseries. Nevertheless, the results of this sections can easily be generalized to the case of Noetherian complex transseries. In this case, we recommend to replace the axiom T4 by the following more complicated but better axiom:

T4

Let be a sequence of monomials in , such that . Then there exists an , such that for all we have

This axiom allows the resolution of certain functional equations like

which admits natural solutions of the form

which are called nested transseries.

2.7Extra structure on the field of transseries in

Consider a tuple of non zero complex transseries in with . We call a complex transbasis if the following conditions are satisfied:

TB1

for some , which is called the level of .

TB2

for each .

TB3

(i.e. ).

Such a transbasis generates a complex asymptotic scale . We say that can be expanded w.r.t. if . If , then we say that (and any ) is purely exponential. The following incomplete transbasis theorem is proved in a similar way as in the case of real transseries:

Theorem 10. Let be a transbasis and a complex transseries. Then can be expanded w.r.t. a super-transbasis of .

We define a strong derivation w.r.t. on in the usual way: we take

for all monomials . This yields a derivation on through extension by strong linearity. Given a derivation on , we define

for all monomials . This again yields a derivation on through extension by strong linearity. By induction over , we thus obtain a derivation on .

We recall that a derivation on is said to be strictly valuated resp. strictly positive if the following conditions are satisfied:

Contrary to the case of real transseries, our derivation on cannot be strictly positive. Indeed, either or , say . Then we have , so either or . On the other hand, the following may be proved in the usual way:

Theorem 11. The derivation on is strictly valuated.

Actually, the proof involves upward shiftings of transseries: given , its upward (resp. downward) shifting is defined by (resp. ). Contrary to the case of real transseries, this transseries does not necessarily live in the same field of transseries as : if , then we have , where and for all transmonomials . In the case of downward shiftings, one may have to consider the generalized fields of complex transseries in from remark 8.

It is more difficult to extend functional composition from the real to the complex setting due to possible incompatibility between the angles and directions. For instance, if , then the transseries can not be composed on the right with . In general, right composition with a given transseries is only defined on a certain subfield of . Contrary to the case of real transseries, certain functional equations like

with seem to fall outside the scope of the theory of complex transseries, unless someone comes up with some really new ideas to incorporate the solutions to such equations inside this theory.

2.8Further generalizations

One of the main ideas behind the construction of fields of complex transseries is that we do not longer require the ordering on the constant field to be compatible with the multiplication. Indeed, we just need the compatibility with the addition (or multiplication with reals), in order to obtain ordered monomial groups via exponentiation.

The above idea may be used to generalize the results from this section to other circumstances. Consider for instance the set of -adic complex numbers, where . Then it is classical that there exists a partial logarithm on , which is defined for all with . By Zorn's lemma, there exists a total ordering on the -vector space . The theory of this section may now be adapted in order to construct the field of complex -adic transseries.

A first change concerns the condition T1, which should now become

Furthermore, it is not as easy as before to characterize the total orderings on , which are compatible with the -vector space structure. Consequently, there is no natural analogue to the condition T4 and we have to satisfy ourselves with the construction of pre-fields of complex transseries. Also, the exponentiation on is not total.

Notice that it seems to be possible to take itself for the indeterminate in the construction of . This would yield a field of transseries which contains and such that the logarithm is defined for all non zero elements.

3Generic complex transseries

In practical computations with complex transseries the angles and directions are not known in advance and we have to choose them (or more precisely, to put constraints on them) as the computation progresses. This can be done by introducing a closed interval for each transmonomial , which corresponds to the constraint

(4)

on . Given such sets , we will work with generic complex transseries which are in the “intersection” of all such that and satisfy the above constraints. Actually, it is convenient to always work w.r.t. generic complex transbases, which we will introduce now.

3.1Generic complex transbases

Let be an -tuple of symbols. Assume that each comes with closed interval modulo , such that . Then we may order the monomial group by

for each non zero monomial with . We call a generic complex asymptotic basis of the scale . Such a basis is called a generic complex transbasis, if

TB1

for some , which is called the level of , and .

TB2

is a regular, infinitely large transseries in for each .

TB3

.

An important question is whether the asymptotic constraints on the determine a non empty region of the complex transplane (see chapter 6 of [vdH97]). This question will be addressed in a forthcoming paper.

Example 12. The triple is a transbasis, for the constraints . Computations with respect to this transbasis are valid in regions of , where . This is for instance the case for , such that in a region where for some small and .

A generic complex transseries is an element of for some complex transbasis . It can be shown that two transbases which have a non empty region of definition in common can be merged together. In the remainder of the paper we will follow an easier approach, which consists of working with respect to a current transbasis, which may be enlarged and on which we may impose additional asymptotic constraints during computations with complex transseries.

3.2Case separations and the field operations

By construction, all ring operations can already be carried out in an algebra of the form . In order to invert a complex transseries, we first have to be able to compute its dominant monomial. In principle, both or might be “the” dominant monomial of a transseries like . Nevertheless, given a transseries with dominant monomials , then we may always separate cases

in each of which has only one dominant monomial. This case separation technique is explained in detail in [vdH97]. In the present context, the imposition of a constraint

with reduces to the insertion of in the interval . If the length of the new interval exceeds , then (4) can not be satisfied, so that the corresponding case does not need to be considered.

Remark 13. In order to be really complete, we should also consider the cases when several dominant monomials are asymptotic. For instance, in the case of the series , we should consider the cases and , but also . However, in the present paper, we argue that the situation when is “degenerate” in the sense that it corresponds to a single “direction” among a continuous number if possibilities.

As a consequence, we notice that the process of “regularization” of a complex transseries is much easier than in the case of multivariate transseries studied in [vdH97]. Indeed, in the case when one has to consider the possibility that , one also has to consider the possibility of cancellation or . This would necessitate refinements of the coordinates and rewriting of the series in .

Example 14. Modulo cases separations, we may thus carry out all field operations. For instance, the inverse of is either given by

or

3.3Logarithms of complex transseries

Consider a non zero complex transseries . Modulo case separations, we may assume that is regular, so that we can write

with and . Consequently,

If , then this series is already in . Otherwise, it still is, modulo the insertion of a new element in front of the transbasis, subject to the constraint . Since is a new symbol, this constraint is non contradictory with the existing expo-linear constraints on the . The relation is automatically verified, since .

3.4Exponentials of complex transseries

Consider a complex transseries . Modulo case separations, we may assume that is regular. In order to compute the exponential of , we distinguish three cases:

Case 1:
is bounded. We may write , with and . Hence, , with and .

Case 2:
for some (where we understand that the left resp. right hand side relation is verified if resp. ). We decompose , where and . Inserting into by , we then have .

Case 3:
for some . We may write , with and . Then and we compute using the same algorithm. The computation of cannot give rise to infinite loops, since the transbasis would remain invariant in such a loop, while the index would strictly decrease.

3.5A worked example

Consider the complex “exp-log function”

and let us show how to expand it generically with respect to a generic complex transbasis. We start with and recursively expand all subexpressions of .

Expansion of .
In order to expand , we fall into the second case of the exponentiation algorithm, since and . Consequently, we insert into using , so that expands as .

Expansions of and .
Since is a ring, we immediately have . Since the expansions of sums and products do not present any problems, we will omit them in what follows.

Expansion of .
In order to expand , we first have to determine the dominant of . Two cases need to be distinguished for this, namely , which corresponds to , and , which corresponds to . In the first case, , so that needs to be inserted into . In the second case, , so we rewrite .

Expansion of .
In the case when , we have , so we rewrite . In the other case, when , the argument is bounded, so that .

Expansion of .
We first have to determine the dominant monomial of . If , then we separate the cases in which , and in which . In the first case, we obtain

In the second case, we get

If , then , so we separate the cases in which , and in which . If , then

Otherwise, we obtain

Figure 2. A plot of the function , which illustrates the four possible asymptotic behaviours of on non degenerate regions. The “rows” of singularities correspond to the borders between regions of different types.

4Parameterized complex transseries

In order to deal with integration constants when solving differential equations, we need to consider parameterized transseries. As in the case of generic transseries, if will often be necessary to distinguish several cases as a function of the values of the parameters. Again, this can be done by putting constraints on the parameters.

4.1Definition of parameterized complex transseries

Let be a -tuple of complex parameters. We call a subset of a region, if is the set of solutions of a system of polynomial equations or inequations

where , and “rational function inequalities on the real parts”

where and does not vanish on . Notice that may be seen as a special kind of semi-algebraic set, under the isomorphism . The polynomial algebra will also be called the coefficient or parameter algebra.

Given a non empty region , let be an -tuple of symbols. Assume that each comes with a finite set of directions, such that does not vanish on for all , , and such that

(5)

for all and with . In the case when , the directions correspond to the extremal angles in the intervals from the previous section.

For each , there exists a natural partial ordering on the -vector space , which is generated by the relations for all . Indeed, the constraints (5) in an arbitrary point guarantee the absence of relations

with . Consequently, we may define a natural neglection relation on the asymptotic scale by

for each non zero monomial with . We say that is a parameterized transbasis, if

TB1

for some , which is called the level of , and .

TB2

is a regular, infinitely large transseries in for each .

TB3

.

A parameterized transseries is an element of for some transbasis .

4.2Uniform regularization

A regular parameterized transseries is said to be uniformly regular, if either , or for all . In this section we prove that any parameterized transseries can be uniformly regularized modulo case separations. We notice that a uniformly regular parameterized transseries on a region remains uniformly regular on any subregion of .

Lemma 15. Let be a monomial. Then, modulo case separations, we may assume that either , or .

Proof. Write , with and separate the following cases :

In the cases A, we have . In the cases B we have . In case C, we have .

Notice that the imposition of the constraints of the form or may involve a reduction of the region and/or the insertion of new directions into . Indeed, is an additional algebraic constraint on . In order to impose , we first impose the constraints and on , for all . Next we insert into .

Lemma 16. Let be infinitesimal monomials in an arbitrary monomial group with -powers, such that , for certain . Then there exist infinitesimal monomials , such that for all .

Proof. Since , we may assume without loss of generality that , modulo a permutation of indices. We will prove the lemma by induction over . For the lemma is trivial. So assume that and let . Then we have either , or .

If , then there exist , such that , by the induction hypothesis. Consequently, . If , then , whence a fortiori . If , then there exist , such that , by the induction hypothesis. Hence . The lemma follows by induction.

Theorem 17. Any can be uniformly regularized modulo case separations.

Proof. Let be such that . By lemma 15, we may assume without loss of generality that either , or for each , modulo some case separations. Without loss of generality, we may therefore assume that admits a Cartesian representation in , i.e. for certain . Choosing minimal, we will prove the theorem by induction over . If , then , and we have nothing to prove. So assume that .

We will first show how to regularize modulo case separations. So let be the set of dominant monomials of . By repeated application of lemma 15, and modulo reordering, we may assume that . If all these inequalities are strict, then we are done, since will be the only dominant monomial. Otherwise, we have for certain , which yields a non trivial relation for certain . Then lemma 16 implies that we may find a Cartesian representation for in variables only, and we are done again, by the induction hypothesis.

In order to make uniformly regular modulo case separations, we use the following algorithm:

We have to show that this algorithm terminates. Assume the contrary and let be the successive dominant monomials of in step 1 on smaller and smaller subregions of . Ultimately, for each , there exists a with in step 2, and the next region is given by in step 3. Now the numerators of all coefficients belong to the Noetherian polynomial ring . Consequently, the increasing chain of ideals is stationary and so is the decreasing chain of subregions of : contradiction.

4.3Computations with parameterized complex transseries

Using the tool of uniform regularization, we may compute with parameterized complex transseries in a similar way as explained in sections 3.2, 3.3 and 3.4. Of course, it may happen that we need to exponentiate or to take logarithms of parameterized constants in . Nevertheless, this can only happen a finite number of times, so that we may see these exponentials resp. logarithms as new parameters. Furthermore, we will show that it is never necessary to exponentiate or take logarithms of parameterized constants during the resolution of algebraic differential equations.

Example 18. Consider the expansion of the function

Case .
We insert into and get

We thus have to determine whether and , which leads to the following cases and expansions for :

Case .
We insert into and next need to determine whether or . This leads to the following cases and expansions for :

In the last exceptional case when , we get

so that we need to determine whether or . This leads to the following final cases and expansions for :

5The differential Newton polygon method

In the remainder of this paper, we will be concerned with the resolution of asymptotic algebraic differential equations like

(6)

where is a differential polynomial with transseries coefficients and a transmonomial.

In this section, we describe the differential Newton polygon method, which enables us to compute the successive terms of solutions one by one. In the next sections, we will be concerned with the transformation of this transfinite process into a finite algorithm. In sections 5, 6 and 7 the transseries in are assumed to be as in section 2. In section 8, we will consider parameterized transseries solutions.

5.1Notations

5.1.1Asymptotic relations

Except for the usual asymptotic relations and , we will also need the flattened relations and there variants , where is an infinitely large or small transseries. These relations are defined by

Notice that , and .

5.1.2Natural decomposition of

The differential polynomial is most naturally decomposed as

(7)

Here we use vector notation for tuples and of integers:

The -th homogeneous part of is defined by

so that

5.1.3Decomposition of along orders

Another very useful decomposition of is its decomposition along orders:

(8)

In this notation, runs through tuples of integers in of length , and for all permutations of integers. We again use vector notation for such tuples

We call || the weight of and

the weight of .

5.1.4Logarithmic decomposition of

It is convenient to denote the successive logarithmic derivatives of by

Then each can be rewritten as a polynomial in :

We define the logarithmic decomposition of by

(9)

where

Now consider the lexicographical ordering on , defined by

This ordering is total, so there exists a maximal for with , assuming that . For this , we have

(10)

for all , whose dominant monomial is sufficiently large.

5.1.5Additive and multiplicative conjugations and upward shifting.

Given a differential polynomial and a transseries it is useful to define the additive and multiplicative conjugates and of w.r.t. and the upward shifting of as being the unique differential polynomials, such that for all , we have

The coefficients of are explicitly given by

(11)

The coefficients of are more easily expressed using decompositions along orders:

(12)

The coefficients of the upward shifting (or compositional conjugation by ) are given by

(13)

where the are generalized Stirling numbers of the first kind:

5.2Differential Newton polynomials

Given a differential polynomial with transseries coefficients, its dominant monomial is defined by

(14)

and its dominant part (or coefficient) by

(15)

The following theorem shows how looks like after sufficiently many upward shiftings:

Proposition 19. Let be a differential polynomial with purely exponential coefficients. Then there exists a polynomial and an integer , such that for all , we have .

Proof. Let be minimal, such that there exists an with and . Then we have and

(16)

by formula (13). Since , we must have . Consequently, . Hence, for some , we have . But then (16) applied on instead of yields . This shows that is independent of , for .

In order to prove the proposition, it now suffices to show that implies for some polynomial . For all differential polynomials of homogeneous weight , let

(17)

Since , it suffices to show that if and only if . Now implies that . Furthermore, (13) yields

(18)

Consequently, we also have . By induction, it follows that for any iterated exponential of . We conclude that , by (10).

Given an arbitrary differential polynomial , the above proposition implies that there exists a polynomial and an integer , such that for all sufficiently large . We call

the differential Newton polynomial of . More generally, given a monomial , we call the differential Newton polynomial of associated to .

5.3Potential dominant monomials and terms

Returning to the asymptotic differential equation (6), we call a potential dominant monomial, if admits a non trivial root , where stands for the algebraic closure of . If , then the corresponding term is called a potential dominant term. The multiplicity of (and of ) is the differential valuation of , i.e. the least such that . The Newton degree of (6) is the largest possible degree of for monomials .

Proposition 20. Assume that is a regular, non-zero transseries solution to (6). Then is a potential dominant term.

A potential dominant monomial is said to be algebraic if is non homogeneous, and differential if . A potential dominant monomial, which is both algebraic and differential, is said to be mixed. Notice that (12) implies

if the coefficients of and are purely exponential.

5.3.1Algebraic potential dominant monomials

The algebraic potential dominant monomials correspond to the slopes of the Newton polygon in a non differential setting. However, they can not be determined directly as a function of the dominant monomials of the , because there may be some cancellation of terms in the different homogeneous parts during multiplicative conjugations. Instead, the algebraic potential dominant monomials are determined by successive approximation:

Proposition 21. Let be such that and .

  1. If is purely exponential, then there exists a unique purely exponential monomial , such that .

  2. Denoting by the monomial in (a), there exists an integer , such that for all we have .

  3. There exists a unique monomial , such that is non homogeneous.

Proof. In (a), let be a purely exponential transbasis for the coefficients of . We prove the existence of by induction over the least possible , such that we may write . If , then we have . Otherwise, let with . Then

so that for some and . By the induction hypothesis, there exists a purely exponential monomial , such that . Hence we may take . As to the uniqueness of , assume that with . Then

This proves (a).

With the notations from proposition 19, we have already shown that and that equality occurs if and only if . Because of (12), we also notice that for all . It follows that

and similarly for instead of , since we necessarily have for some . We finally notice that and imply that , since whenever and . Consequently, and stabilize for with . For this , we have (b).

With the notations from (b), is actually the unique monomial such that

is non homogeneous for all sufficiently large . Now for sufficiently large . This proves (c) for purely exponential differential polynomials , and also for general differential polynomials, after sufficiently many upward shiftings.

The unique monomial from part (c) of the above proposition is called an equalizer or the -equalizer for . An algebraic potential dominant monomial is necessarily an equalizer. Consequently, there are only a finite number of algebraic potential dominant monomials and they can be found as described in the proof of proposition 21. Notice that, given a transbasis for the coefficients of , all equalizers for belong to .

5.3.2Differential potential dominant monomials

In order to find the differential potential dominant monomials, it suffices to consider the homogeneous parts of , since , if and . Now we may rewrite as times a differential polynomial of order in . We call the -th Ricatti equation associated to . Since solving is equivalent to solving , we are entitled to expect that finding the potential dominant monomials of w.r.t. is equivalent to solving “up to a certain extent”.

Proposition 22. The monomial is a potential dominant monomial of w.r.t.

(19)

if and only if the equation

(20)

has strictly positive Newton degree.

Proof. We first notice that for all and . We claim that the equivalence of the proposition holds for and if and only if it holds for and . Indeed, is potential dominant monomial w.r.t. (19), if and only if is a potential dominant monomial w.r.t.

(21)

and (20) has strictly positive Newton degree if and only if

(22)

has strictly positive Newton degree. Now the latter is the case if and only if

has strictly positive Newton degree. But

This proves our claim.

Now assume that is a potential dominant monomial w.r.t. (19). In view of our claim, we may assume without loss of generality that and are purely exponential and that . Since is homogeneous, we have for some and

Since is purely exponential, it follows that has degree , so that the Newton degree of (20) is at least . Similarly, if is not a potential dominant monomial w.r.t. (19), then and

for some . Consequently, for any infinitesimal monomial , and the Newton degree of (20) vanishes.

5.4Refinements

Now we know how to determine potential dominant terms of solutions to (6), let us show how to obtain more terms. A refinement is a change of variables together with an asymptotic constraint

(23)

where . Such a refinement transforms (6) into

(24)

We call the refinement admissible, if (24) has strictly positive Newton degree.

Proposition 23. Let be the dominant term of and assume that . Then the Newton degree of (24) is equal to the multiplicity of as a root of .

Proof. Let us first show that for any monomial . Modulo replacing by we may assume without loss of generality that . Modulo a sufficient number of upward shiftings, we may also assume that , that , and that , and are purely exponential. The differential valuation of being , we have in particular . Hence,

for all . We infer that .

At a second stage, we have to show that . Without loss of generality, we may again assume that , that , and that and are purely exponential. The differential valuation of being , we have for all . Taking , we thus get

for all . We conclude that .

5.5A worked example

Consider the algebraic differential equation

(25)

Let us start by computing the potential dominant monomials of . We first have to find the -equalizer relative to (25). Since , we cannot have , so we have to compute

In order to “equalize” and , we have to conjugate multiplicatively with :

At this point, we observe that , so we have found the -equalizer, which is . Since , the corresponding algebraic potential dominant term of is . As to the differential potential dominant monomials, we have

Clearly, has no roots and has all constants as its solutions modulo . Consequently, is a potential dominant monomial of for all , such that . The corresponding differential potential dominant terms are of the form , with and .

In order to find more terms of the solution to (25), we have to refine the equation. First of all, consider the refinement

which transforms (25) into

(26)

Since , we first observe that is actually a solution to (25). On the other hand, since is a potential dominant term of multiplicity 1 of , the Newton degree of (26) is one. The only potential dominant monomials of therefore necessarily correspond to solutions modulo of the Ricatti equation

These solutions are of the form and , which leads to the potential dominant monomials and , from which we remove , since . Expanding one term further, we see that the generic solution to (26) is

with and where the case recovers the previous solution. In other words,

is the first type of generic solution to (25).

As to the second case, we consider the refinement

which transforms (25) into

(27)

Again, this equation has Newton degree one. On the one hand, we observe that the linear part of this equation only admits solutions with dominant monomial or . Consequently, (27) admits at most one solution. On the other hand, we will show in the next section that quasi-linear equations (i.e. of Newton degree one) always admit at least one solution. In our case, this leads to the following second type of generic solution to (25):

For the present example, we actually even found exact solutions. Of course, the expansions are infinite in general.

6Distinguished solutions

6.1Distinguished left inverses of linear differential operators

Let be a purely exponential transbasis. A linear operator on

is said to be grid-based if its operator support

is grid-based. For all transseries we have

In particular, the differentiation on is grid-based with

Consequently, any linear differential operator with coefficients in is also grid-based, since

We will now show that also admits a so called distinguished left inverse , which is linear and grid-based. Here a distinguished solution to the equation

is a solution , such that for all other solutions , we have . Distinguished solutions are clearly unique. We say that is a distinguished left inverse of , if is a distinguished solution to for each .

In what follows, we will often consider linear differential operators as linear differential polynomials. In this case, you should keep in mind that denotes the coefficient of in and not the -th homogeneous part. We will also denote for any linear differential operator as above.

Theorem 24. Let be a linear differential operator with coefficients in and . Then admits a distinguished linear left inverse on . This left inverse is grid-based and

where

Proof. Let and . There exists a unique strongly linear operator , such that

for all . The operator admits a natural left inverse , which is constructed as follows. Let , where is purely exponential. By proposition 21(a), there exists a purely exponential monomial with . Let and respectively denote the dominant monomial and dominant part of . Let

where is minimal with . Then we observe that

so that . Consequently, we may take and extend to the whole of by strong linearity.

Let . By construction, the operator is strictly extensive, and the operator coincides with on . Now consider the functional

By the implicit function theorem from [vdH00b], there exists a linear operator

such that for all . Consequently,

is a strongly linear left inverse for .

In order to prove that is actually a grid-based operator, we first notice that, by construction,

Since

it therefore suffices to prove that and are grid-based. But this follows from theorem 17 when considering as a generic transseries in , for . Indeed, there exist a finite number of regions, each on which is uniformly regular. Consequently, can only take a finite number of values and

is contained in the union of the supports of the generic transseries on each of the regions.

Remark 25. Each induces a canonical solution to . This canonical solution satisfies and for all . Actually, the canonical solutions are polynomials of degree in with coefficients in . In order to see this, let

Then we observe that maps into and that maps into .

6.2Distinguished solutions of quasi-linear equations

Let be a subset of a monomial group. The notion of operator support can be extended to strongly -linear operators by

More generally, if is a Noetherian operator, then we define its operator support by

where stands for the -th homogeneous part of . We have

for all . We say that is grid-based, if is grid-based.

Let be a purely exponential transbasis and a differential polynomial with coefficients in . Notice that we may naturally consider as a grid-based operator on . The equation (6) is said to be quasi-linear if its Newton degree is one. A solution to such an equation is again said to be distinguished if we have for all other solutions to (6).

Theorem 26. Assume that the equation (6) is quasi-linear. Then it admits a distinguished transseries solution.

Proof. Without loss of generality, we may assume that and . We prove the proposition by induction over . If , then we must have , so that is the distinguished solution to (6). So assume that and let

be the dominant part of w.r.t. . By the induction hypothesis, there exists a distinguished solution to the quasi-linear equation

(28)

We first proceed with the refinement

so that , and a sufficient number of upward shiftings, so that is purely exponential. We next decompose as

where

Let . Since is stable under and for each , the operator is strictly extensive on . Consequently, the implicit function theorem from [vdH00b] implies that the operator can be inverted, like in the proof of theorem 24:

In particular,

is a solution to . Furthermore, we have

so that

We claim that is the distinguished solution. Indeed, let be another solution and let . If , then

is a solution to (28), so that . If , then let

Since , we have , so that is the dominant monomial of a solution to the homogeneous equation . Consequently, , since .

Remark 27. By induction over , it also follows that we need at most upward shiftings in order to express the distinguished solutions. In other words, if has coefficients in , where is purely exponential, then . Actually, if is the order of , then the number of upward shiftings we need is also bounded by .

Indeed, denoting and using a similar argument as in remark 25, we first observe that is bijective on if admits no solutions in . Moreover, if admits such a solution, then has a root with the same dominant part w.r.t. . The same observations recursively hold for all involved in the resolution of (28). Now if is the distinguished solution of (6), then the linear equation admits at most solutions. Hence, there are at most transbasis elements for which we need to make an upward shifting.

6.3A worked example

Consider the linear differential equation

(29)

under the assumptions and . Then has coefficients in and

for each , so that

Hence, theorem 24 implies that (29) has a distinguished solution in with . Actually, it is easily seen that

and the first terms of are given by

In order to find all solutions to (29), we have to solve the Ricatti equation associated to the linear part of (29):

(30)

This equation has two potential dominant terms and of multiplicities one. Consequently, we get quasi-linear equations when refining or . When setting resp. , these equations are conveniently rewritten as

(31)

and

(32)

By theorem 26, these equations admit distinguished solutions

More precisely, in the proof of theorem 26, and for (31), we would have , and . It follows that , so that . Similarly, . Returning to (30), we obtain the following solutions:

which yield a basis

of solutions to .

It is interesting to study the solutions to (29) from an analytical point of view. Indeed, the asymptotic conditions or and or divide complex space into four non degenerate regions. However, each of these regions has infinitely many “bounded connected components”. When moving from one connected component to another one, a “generalized Stokes phenomenon” occurs. Consequently, a specific formal solution to (29) only makes sense on a bounded connected component from the analytical point of view.

Nevertheless, it is possible to give an asymptotic meaning to the generic formal solution to (29) on each region, by associating a “generalized Stokes matrix” to each connected component of the region. This issue will be detailed in a forthcoming paper. An interesting remaining question is the asymptotic behaviour of the Stokes matrices. Actually, the generalized Stokes phenomenon might be qualified as multi-Stokes phenomenon, since the Stokes phenomena occur with respect to several generalized sectors of different types. Equation (29) is one of the simplest examples which exhibits this multi-Stokes phenomenon.

7Unravellings

7.1Total unravellings

Theorem 26 together with propositions 21, 22 and 23 suggest that the solutions to an arbitrary asymptotic algebraic equation (6) can be expressed using the field operations, exponentiation, logarithm and distinguished solutions of quasi-linear equations. This is indeed so, if the Newton degree decreases at each refinement in proposition 23.

The remaining case, when the Newton degree repeatedly does not decrease in proposition 23, occurs when there are “almost multiple solutions”. In order to “unravel” these solutions, we have to find their greatest common part. More precisely, consider an asymptotic algebraic differential equation (6) of Newton degree . Then an unravelling (or total unravelling) is a refinement

such that

U1

The Newton degree of equals .

U2

For any , the Newton degree of is .

Clearly, the series , which is also called an unraveller, may be replaced by any other series of the form with .

From a theoretical point of view it is possible to prove a certain number of facts about unravellings. First of all, any unraveller admits a truncation , which is a canonical unraveller, in the sense that

is an unravelling for all with , and that a similar property does not hold for any proper truncation of .

Secondly, it is possible to construct the so called canonical algebraic unraveller by transfinite induction: having constructed the first terms of , say of sum , one looks at the equation of Newton degree . If this equation has an algebraic potential dominant term of multiplicity , then this term is unique, and we take it to be the next term of .

However, in what follows, we are interested in more constructive ways to obtain unravellings. For this purpose, we recall that in the more classical context of algebraic equations, multiple roots are usually found by solving the derivative (or a higher derivative) of the equation with respect to the indeterminate. In the next sections, we will describe a similar strategy in order to find the almost multiple solutions to asymptotic algebraic differential equations. The price to be paid is that we will need a sequence of so called partial unravellings (and adjusted partial unravellings) in order to construct a total unravelling.

7.2Partial unravellings

Consider an asymptotic algebraic differential equation (6) of Newton degree . Given a monomial such that admits a root of multiplicity , we define by

Now let , and be such that

PU1

.

PU2

The Newton degree of is .

PU3

For any with , the Newton degree of is .

Then the refinement

(33)

is said to be a partial unravelling with as its associated monomial. Notice that the equations and are quasi-linear. Partial unravellings are constructed as follows.

Proposition 28. Let and be as above. Then there exists a which satisfies the conditions PU1, PU2 and PU3.

Proof. We construct sequences and of approximations of and , such that all and satisfy the conditions PU1 and PU2. We let be the distinguished solution to the equation and . As long as and do not satisfy the condition PU3, there exists a with , such that the Newton degree of is . Hence we may take and .

We claim that the sequences and are of length at most , so that we may take their last elements for and . Indeed, for each , the series with are solutions to the quasi-linear equation . Consequently, the dominant monomials of these series, which are pairwise distinct, are all dominant monomials of solutions to the homogeneous linear differential equation . But there are at most linearly independent solutions to this equation.

Proposition 29. Consider a partial unravelling (33) as above, followed by a refinement

such that the Newton degree of

(34)

is equal to . Then, for , we have

Proof. Without loss of generality, we may assume that , , , and are purely exponential, that and that . From PU3 it follows that is neither a potential dominant monomial for , nor for . Proposition 22, applied to and the “non potential dominant monomial” , therefore yields

Consequently,

On the other hand, we have

so that

Now recall that is the coefficient of in for some . It follows that

Now assume that is a monomial with (so that ). Then we have

We conclude that the degree of can not exceed . If is chosen such that (34) has Newton degree , it thus follows that

which completes the proof.

7.3Adjusted partial unravellings

Proposition 29 shows that by taking sequences of partial unravellings, we rapidly approach a total unraveling. The only problem which still remains to be solved is the appearance of highly iterated logarithms. We will first solve this problem in the particular case when the Newton degree of coincides with its normal degree. In the next subsection, we will show that the general case can be reduced to this case.

In the sequel, we assume that (6) is an asymptotic differential equation of degree and Newton degree , such that the following additional conditions are satisfied for a certain purely exponential transbasis :

E1

has coefficients in .

E2

have coefficients in .

E3

admits only potential dominant monomials in .

The two first conditions can clearly be met after a sufficient number of upward shiftings. In section 9, we will show that this is also the case for the last condition.

Proposition 30. Let be a potential dominant term of multiplicity for (6). Then

  1. Modulo the insertion of new elements into , we have .

  2. There exists a unique , such that either

    1. is a total unravelling and is -minimal in .

    2. The Newton degree of is .

Proof. Let us first prove (a). If is differential, then E3 implies that is purely exponential, so after a suitable extension of . If is algebraic, then is the -equalizer for each , since has multiplicity . Proposition 21(a) implies that there exists a unique purely exponential monomial with , such that for all . More precisely, in the algorithm in proposition 21(a), is chosen such that , whence for some , and satisfies . In particular, for , this yields . We claim that .

If , then we have and for all . Since , this can only happen if . Hence is the -equalizer w.r.t. for all and . If , then E3 implies that is not a potential dominant monomial for . Consequently, the coefficients of and in both do not vanish. It follows that is the -equalizer w.r.t. and again .

We prove the existence of in (b) by induction over ; the uniqueness of follows from E3. If , then clearly satisfies assumption ii. If , then we refine

and remark that satisfies the hypothesis E1, E2 and E3, due to part (a). Now consider the equation

(35)

of Newton degree . If this equation admits a potential dominant term of multiplicity with , then the induction hypothesis implies that there exists a , which satisfies the assumption i or ii, and we may take . If there does not exist such a potential dominant term , then there either do not exist potential dominant terms of multiplicity at all for (35), so that i holds for , or such potential dominant terms do exist, and we have ii for .

Given a potential dominant term of multiplicity , let be as in proposition 30(b). In case i, we say by convention that is an adjusted partial unravelling. In case ii, let

(36)

be a partial unraveling relative to the equation

and with as its associated monomial. Then we say that

is an adjusted partial unravelling. Notice that a partial unravelling like (36) always exists, by propositions 28 and 30(a).

Notice also that we necessarily have . Indeed, consider the differential polynomial with in PU1. Since , this differential polynomial is actually linear. Furthermore, since , the coefficients of are in and the coefficients of in . We conclude that all solutions to , and in particular , are in . A consequence of our observation is that again satisfies the hypotheses E1, E2 and E3, so that we may consider sequences of adjusted partial unravellings.

Proposition 31. Any sequence of adjusted partial unravellings

is finite, say of length , and its composition

is a total unravelling.

Proof. Let denote the length of the sequence of adjusted partial unravellings. For each , let . For each , let denote the exponentiality of . Given , proposition 29 implies that

Since and , this yields

By induction, it follows that . We conclude that . The composition of the sequence of adjusted partial unravellings is clearly a total unraveling.

7.4Construction of total unravellings

Let us now return to the case of a general asymptotic differential equation (6) of Newton degree . Assume that (33) is a partial unravelling with and that is a potential dominant monomial of multiplicity for

(37)

Modulo a sufficient number of upward shiftings, we may assume that and the coefficients of can be expanded w.r.t. a purely exponential transbasis . Let be the transbasis element such that , and consider the dominant part of with respect to :

On the one hand, since , we have

for all , so that . Consequently, satisfies the conditions E1 and E2 from the previous section; we will see in section 9 that it also satisfies E3, modulo some additional upward shiftings. On the other hand, the following proposition reduces the problem of determining the unravellings for (6) to a similar problem for . In view of the previous section, this completes the effective construction of unravellings.

Proposition 32. With the above notations, a refinement

(38)

with is a total unravelling w.r.t. (37) if and only if

(39)

is a total unravelling w.r.t. the equation

(40)

Proof. Modulo a multiplicative conjugation, we may assume without loss of generality that . Now if (38) is an unravelling, then proposition 29 implies that

so that . Actually, in the proof of proposition 29 we showed that

so that . We infer that for all with . In other words, for (39) to be an unravelling, it is again necessary that .

The above argument shows that it suffices to prove the equivalence under the assumption that . Now we notice that for each transseries and each monomial , the dominant parts of and w.r.t. coincide. Consequently,

for such and . In particular, we have

for all sufficiently close to , so that the Newton degrees of

and

coincide. Hence U1 holds for (38) if and only if it holds for (39). Similarly, for all , such that , the Newton degrees of

(41)

and

(42)

coincide. Furthermore, for a similar reason as above, the Newton degrees of (41) and (42) are both bounded by if . In other words, U2 holds for (38) if and only if it holds for (39). We conclude that (38) is a total unravelling w.r.t. (37) if and only if (39) is a total unravelling w.r.t. (37).

7.5Worked examples

One of the easiest examples which illustrates the importance of unravellings is

(43)

where . This equation admits as it's unique potential dominant term of multiplicity . However, the refinement

leads to the equation

which again admits a unique potential dominant term of multiplicity . Continuing like this leads to an infinite sequence , of refinements, and we do not succeed in separating the two solutions.

Therefore, we rather construct a total unravelling. In order to do so, we first compute the distinguished solution to the quasi-linear differential equation

and then refine

This refinement (and partial unravelling with associated monomial ) is actually already the total unravelling we were looking for and the equation (43) transforms into

This time, the new equation admits two potential dominant terms and of multiplicities , which allows us to compute the solutions to (43) by recursion.

In general, total unravellings can only be achieved via successions of partial unravellings. An important example which illustrates this phenomenon is the following:

(44)

This equation becomes purely exponential after upward shiftings:

(45)

This new equation admits a unique potential dominant monomial of multiplicity . Indeed, this is easily seen when substituting for :

(46)

Now is a partial unravelling w.r.t. (46) and a partial unravelling w.r.t. (45). However, the partial unravelling transforms (46) into an equation of the same form as (45), but with decreased by . Consequently, we need a succession of unravellings

in order to attain the total unravelling

Notice that the ratios of the successive potential dominant terms indeed satisfy proposition 29.

An open question is whether there exist examples which essentially need the technique of adjusted partial unravellings in order to limit the appearance of iterated logarithms. If not, this would allow some major simplifications in the algorithm solve in the next section. Actually, the whole computation process of total unravellings needs to be better understood in order to generalize it to other fields of transseries (see section 9.5).

8Computation of parameterized solutions

8.1About the algorithms in this section

In this section, we give algorithms to solve an asymptotic differential equation (6), where the coefficients of and can be expanded with respect to a parameterized transbasis . Actually, we show how to solve such an equation modulo a transmonomial . Here a solution modulo to (6) is a transseries , such that the Newton degree of is strictly positive. In our algorithms, we will use the following conventions without further mention.

Automatic case separations.
Our algorithms are non deterministic in the sense that we allow automatic case separations, in a similar way as sections 3 and 4. The algorithms are constructed in such a way that each solution modulo to (6) occurs in precisely one case. In other words, for each specialization of the initial parameters and, for each specialization of the directions (which corresponds to a choice of a field of transseries as in section 2, which satisfies the constraints on the ) and for each solution in , there exists precisely one case and precisely one specialization of the newly introduced parameters, which yields as a solution.

The termination of our non deterministic algorithm follows from the termination of each of its branches by a similar Noetherianity argument as in the proof of theorem 17. For more details about the automatic case separation strategy, see chapter 8 in [vdH97].

Automatic upward shiftings.
We assume throughout our algorithms that all computations are done w.r.t. a purely exponential parameterized transbasis . In order to do this, we are allowed to insert new elements into and to “shift the whole computation upwards” whenever necessary. Upward shifting can for instance be implemented via an exception, which starts over the whole non deterministic computation, after shifting the input upwards. A better strategy, which consists of associating an “exponential level” to each transseries, is explained in more detail on page 276 of [vdH97].

Non standard monomials.
It is useful to allow a few non standard transmonomials in our algorithms, like the formal infinitely large and small monomials resp. , as well as monomials of the form .

Effective computations with transseries.
Of course, our algorithms are not really effective, as long as we do not provide algorithms to compute the expansions of parameterized transseries and to test them for being zero. In this paper, we assume that we have oracles for deciding such questions. Similarly, in our algorithms for computing distinguished solutions in the next section, we will merely give infinite formulas for the results. Nevertheless, we notice that for certain classes of transseries, the oracles may be replaced by real algorithms, in a similar way as described in chapter 12 of [vdH97].

8.2Distinguished solutions

In order to get a parameterized version of theorem 24, we have to make sure that the operator is well defined. This can be done as follows:

Algorithm linear

Input A linear differential operator and transseries

Output The distinguished solution to

Step 1 [Introduce generic monomial]

Let denote the current purely exponential transbasis

Let be temporary new parameters in and set

Step 2 [Compute the distinguished solution]

Uniformly regularize

Let , with the notations from the proof of theorem 24

Step 3 [Clean up]

Destroy the parameters by projection of the regions

Return

In the algorithm, the uniform regularization of a differential operator or polynomial means that we uniformly regularize all its coefficients and that we make the corresponding dominant monomials pairwise comparable for using theorem 17 and lemma 15.

In the parameterized context, the equation (6) has a Newton degree , if each specialization of the directions and parameters leads to an equation of Newton degree . We will show below how to compute the Newton degree modulo case separations. The distinguished solution to a quasi-linear equation (i.e. of Newton degree ) is computed inductively as in the proof of theorem 26. The dominant part of w.r.t. is computed by the usual formula after the uniform regularization of .

Algorithm quasi_linear

Input

An integer (with as default value), a differential polynomial with coefficients in and a monomial , such that (6) is quasi-linear

Output The distinguished solution to (6)

Step 1 [Normalize]

If , then return times quasi_linear

Uniformly regularize

If then return quasi_linear

Step 2 [Recurse]

Compute the dominant part of w.r.t.

Let

Step 3 [Return]

Return , with the notations from the proof of theorem 26

8.3Determining the Newton polygon

The -equalizer of a differential polynomial is computed similarly as in the proof of proposition 21, by uniformly regularizing and at each recursion.

Algorithm equalizer

Input A differential polynomial and integers

Output The -equalizer for

Step 1 [regularize]

Uniformly regularize and

Step 2 [equalize]

Let

If then return

Step 3 [shift upwards]

If and , then return

Shift upwards and return to step 2

The analogue of the Newton polygon associated to an algebraic differential equation in the differential case is the determination of -equalizers, which occur as potential dominant monomials, and which are extremal in the sense that is maximal. These equalizers correspond to the slopes of the Newton polygon and the and to the first coordinates.

Algorithm Newton_polygon

Input A differential polynomial

Output

Indices and potential dominant monomials for , such that is the -equalizer for for each

Step 1 [Initialize]

Step 2 [Insert vertex]

If then return and

Step 3 [Search edge]

Compute for all with

Let and let be minimal with .

Set and go to step 2

The Newton degree of (6) can easily be read from the Newton polygon:

Algorithm Newton_degree

Input A differential polynomial and a monomial

Output The Newton degree of

Compute and by Newton_polynomial

Let be maximal, such that

Return

8.4Computing potential dominant terms

Given the Newton polygon associated to , let us now show how to determine the potential dominant terms of solutions to (6) and their multiplicities. We separate a case for each edge and each vertex of the Newton polygon, and determine the corresponding algebraic or mixed resp. differential potential dominant monomials and terms. In order to determine the differential potential dominant monomials, we recursively have to solve Ricatti equations modulo . The algorithm solve which does this will be specified in the next section.

Algorithm pdt [non deterministic]

Input A differential polynomial and a monomial

Output A potential dominant term for (6)

Step 1 [Determine Newton degree]

Compute and by Newton_polygon

Choose a , such that or , and set

If then go to step 3

Separate two cases and go to step 2 resp. 3

Step 2 [Algebraic and mixed potential dominant terms]

Let

Let be a new parameter in

Impose the constraint (as an algebraic constraint, since )

Return

Step 3 [Differential but non mixed potential dominant terms]

Let

Let , where the integral is computed using linear

If then impose the constraint

Otherwise impose the constraint

If then impose the constraint

Let be a new parameter in and impose the constraint

Return

Algorithm multiplicity

Input A differential polynomial and a term

Output The multiplicity of as a root of

Repeat

Uniformly regularize

If then shift upwards

Until

Return the multiplicity of as a root of

8.5Solving the differential equation

We can now state the main resolution algorithm for solving asymptotic algebraic differential equations (6) modulo monomials .

Algorithm solve

Input A differential polynomial and monomials and

Output A solution to modulo

Step M1 [Initialize]

modenormal

Step M2 [Are we done?]

If Newton_degree, then separate two cases and respectively

1. Return

2. Proceed with step M3

Step M3 [Compute potential dominant term]

Let

Let

If then kill this process

Step M4 [Does have multiplicity ?]

If then

modenormal

Go to step M2

Step M5 [Dispatch on mode]

If modenormal then go to step H1

If modeunravel then go to step H3

If modeadjust then go to step U4

Step H1 [Prepare first step unravelling loop]

modeunravel

serialhead

Step H2 [Prepare partial unravelling loop]

While shift upwards

If with then

Otherwise

Step H3 [Partial unravelling]

If then

Go to step M2

Step H4 [Dispatch on serial]

If serialhead then go to step T1

If serialtail then go to step T2

Step T1 [Prepare other steps unravelling loop]

Let be such that

Uniformly regularize

Compute the dominant part of w.r.t. and set

serialtail

Step T2 [Prepare next step unravelling loop]

If there is no purely exponential monomial in with , then set

Let be a purely exponential monomial in , such that

modeadjust

Step T3 [Adjust partial unravelling]

If then

Go to step M2

If then

modeunravel

Go to step H3

The algorithm solve gradually constructs a solution modulo to (6) via a succession of refinements. Each time we get back to the main entry M2 of the loop, we actually have to solve the equation . Given a potential dominant term for this equation, the next refinement (and value of ) depends on the mode variable.

The core of the algorithm consists of steps M1-M5 in which case modenormal. As long as we do not hit a potential dominant term of maximal multiplicity , the algorithm only executes steps M1-M5. Given a potential dominant term of multiplicity , we can simply take for our refinement, which corresponds to the assignments and in step M4.

The steps H1-H4 correspond to the first partial unravelling when we hit a potential dominant term of maximal multiplicity . As long as we do not enter T1-T3, we will have , modeunravel and serialhead. We start by computing the differential polynomial from section 7.2 (modulo an additive conjugation by ). We then keep refining as far as possible in step H3, where is the distinguished solution to the quasi-linear equation .

If the steps H1-H4 do not lead to a complete unravelling, we apply the theory from sections 7.3 and 7.4 in steps T1-T3. We start by computing once and for all the differential polynomial in T1, with this change that we apply a multiplicative and additive conjugation to it in order to make it “compatible” with . The transseries , which is initialized with , may become an iterated exponential as a result of upward shiftings. As long as the current potential dominant term has not yet the required form in order to start a partial unravelling, we have modeadjust, and we keep on adjusting in step T3.

The termination of solve is guaranteed by propositions 23 and 31, modulo the hypothesis that the resolution process requires only a finite number of upward shiftings. An upper bound for will be given in the next section.

9Main theorems and final remarks

9.1Complex transseries solutions to algebraic differential equations

The following main theorem describes the general form of solutions to asymptotic algebraic differential equations (6) with parameterized complex transseries coefficients.

Theorem 33. Consider an asymptotic algebraic differential equation (6) with transseries coefficients. Then, modulo case separations, there exist a finite number of parameterized transseries solutions to (6), with the following properties:

  1. The logarithmic depths of do not exceed the logarithmic depths of the coefficients of by more than a fixed constant , which only depends on the Newton degree , the order and the weight of (6).

  2. For each specialization of the parameters occurring in the coefficients of , for each specialization of the directions, and for each solution to (6) after these specializations, there exists exactly one and exactly one specialization of the remaining parameters on which depends , for which specializes to .

Proof. In view of the algorithm from the previous section, we only have to prove (a). We prove the bounds for by a double induction over and . For , we necessarily have , and clearly . So assume that . If , then (6) has no solutions, so that . Assume therefore that .

We first observe that the number of upward shiftings needed to compute a potential dominant term is bounded by

by propositions 21(b), 22 and the induction hypothesis. We have to estimate the maximal number of upward shiftings which may occur in the main loop of solve, before we reach a lower Newton degree. Now the first partial unravelling requires at most upward shiftings, in view of remark 27. By the induction hypothesis and proposition 30, the second loop of adjusted partial unravellings requires at most upward shiftings. Finally, the decisive refinement, which decreases the Newton degree, again needs at most upward shiftings. Altogether, we obtain

Consequently,

In particular, for , we obtain

For , we obtain

By induction, we finally notice that

which implies our bound.

Remark 34. Of course, it is possible to improve the bound for for particular values of and . First of all, in the case when , it is easily checked that upward shifting is sufficient in proposition 21(b), so that

This observation implies the sharper bounds

for . A careful analysis of the differential Newton polygon method will probably lead to even sharper bounds for small values of . Similarly, it is possible to improve the bounds for small values of , by using the fact that the weight of is bounded by for .

Although the above theorem describes the general form of solutions to (6), it does not claim the actual existence of such solutions. We say that is a solution of multiplicity of (6), if the differential valuation of equals . The following theorem stipulates the existence of solutions to (6) of a very special form.

Theorem 35. Consider an asymptotic algebraic differential equation (6) of Newton degree and whose coefficients can be expanded w.r.t. a transbasis . Then there exist at least solutions to (6) when counting with multiplicities. Moreover, these solutions can be expanded w.r.t. for some .

Proof. Without loss of generality, we may assume that . Let us prove the theorem by induction over . For we have nothing to prove. For , the equation is quasi-linear and the distinguished solution can be expanded w.r.t. . Assume therefore that .

If there exists only one algebraic potential dominant term with multiplicity , then consider the unravelling we obtain by executing solve, but where we always choose the unique algebraic potential dominant term in pdt. Since this branch only involves the computation of equalizers and solutions of quasi-linear equations, can be expanded w.r.t. a transbasis of the form . Modulo replacing by , we may thus assume without loss of generality that (6) admits no algebraic potential dominant terms of multiplicity .

If there exists a mixed potential dominant monomial , then is a potential dominant term of multiplicity for each , and the coefficients of can be expanded w.r.t. . By the induction hypothesis each equation admits at least one solution which can be expanded w.r.t. for some . Hence, there exists an infinity of solutions with the required properties. In what follows, we therefore assume that all potential dominant monomials are algebraic, but not mixed.

Now let and be such that is the -equalizer for each . For each the Newton polynomial is a polynomial with valuation and degree , which has roots (when counting with multiplicities). These root induce at least potential dominant terms, which can be expanded w.r.t. , and whose multiplicities are . By proposition 23 and the induction hypothesis, this leads to at least solutions of the required form, when counting with multiplicities. The theorem now follows from the fact that is a solution of multiplicity .

9.2Linear differential equations

We recall that a differential field is said to be differentially algebraically closed, if for any pair of differential polynomials over , such that the order of is strictly larger than the order of , there exists a root of in , which is not a root of .

Let be a field of complex transseries as in section 2. Unfortunately, theorem 35 is not sufficient for to be differentially algebraically closed. Indeed, the only transseries solutions for the elliptic equation

are , and . Consequently, there are no transseries solutions to this equation, which are not solutions of the equation of lower order

Nevertheless, theorem 35 is sufficient for the following application.

Theorem 36. Let be a linear differential operator of order with coefficients in . Then

  1. can be completely factored over .

  2. There exist linearly independent solutions to in .

Proof. By theorem 35 the Ricatti equation associated to has at least one solution . Consequently, we may factor

for some linear differential operator of lower order and coefficients in . Part (a) now follows by induction over .

Now consider a factorization

(47)

with and let

where stands for distinguished integration. Then . Moreover, by the distinguished properties of the left inverses , we have

for all . This guarantees the linear independence of . Indeed assume that there exists a relation

with . Then . This contradiction completes the proof of (b).

Remark 37. When choosing the factorization 47 in such a way that , we even obtain the canonical basis of solutions of in the proof of theorem 36.

Remark 38. In the case of real transseries, it can be shown that each linear differential operator may be factored as the product of a transseries and operators of the form

or

9.3Bounding the number of integration constants

Although the algorithm solve provides us with the generic solution to (6), it is not clear a priori that the number of new parameters on which the solution depends does not exceed . In this section we sketch a proof of the fact that the number of such integration constants is indeed bounded by .

We first notice that the only place where we introduce (continuous) integration constants is in step 3 of pdt. Each integration constant can therefore be “attached” to a solution of a Ricatti equation of the form . Given an arbitrary moment during the algorithm solve, we actually search solutions of the form

where are the “active integration constants”. The idea is now to set

and to consider as a differential polynomial of order in , with coefficients in . In other words, we consider as new monomials and we give the natural “pointwise” quasi-ordering (see chapter 6 of [vdH97]).

The only obstruction for the computation with coefficients in instead of coefficients in is when the uniform regularization of a transseries in is not possible. Now this obstruction corresponds to the imposition of an algebraic constraint on an active integration constant , when performing the same computation in . In order to solve this problem, an “error handler” is installed each time that we introduce a new continuous integration constant . Whenever we impose an algebraic constraint on , we go back to the error handler and reperform the same computations while assuming that either did or did not (non determinism) satisfy the algebraic constraint right from the start.

In all branches of the new resolution process, the order of the asymptotic differential equation, when rewritten as an equation in , does not exceed . Consequently, at the end of each branch of the process.

9.4Comparison with previous work and errata

The reader may have noticed a certain number of changes with respect to the treatment of algebraic different equations in [vdH97]. Although the results of this paper were stated in the context of grid-based transseries, they may easily be adapted to the well-ordered context from [vdH97], except for the results about parameterized transseries, which become more complicated. The algorithm solve may still be applied in the well-ordered context, except that the introduction of new parameters should then be interpreted as a new source of (continuous) case separations.

During a careful reexamination of our previous work, we noticed that proposition 5.7(c) in [vdH97] does not hold for all . Consequently, our previous treatment of almost double solutions in section 5.5.1 does not work. The present, more complicated, treatment using unravellings corrects this error. When calling a refinement occurring in our construction of a total unravelling a privileged refinement, the proof of theorem 5.2 in [vdH00a] remains correct (except for the bound for the maximal length of a chain of privileged refinements, which may have to be replaced by a larger bound).

Some other changes with respect our previous work are the following:

9.5Conclusion and perspectives

In this paper, we have generalized the transseries technique for solving algebraic differential equation as far as reasonably possible. Three main problems remain to be solved.

Analytic counterpart.
First of all, one has to show that the complex transseries solutions to algebraic equations have a genuine analytic meaning. This problem, which will be treated in a forthcoming paper, can actually be subdivided in two parts:

Differentially algebraic closure.
We have already remarked that our fields of complex transseries are not differentially algebraically closed. In other words, we still miss most of the solutions to algebraic differential equations in our formalism. In order to get a full understanding of the asymptotic behaviour of solutions to algebraic differential equations, two approaches may be followed:

In any case, a purely local treatment seems not to be possible in order to describe all solutions to an algebraic differential equation. A good combination of a more global theory with our local results might lead to the resolution of interesting questions, such as

Is it possible for an analytic solution to an algebraic differential equation with coefficients in to admit a natural boundary somewhere on its Riemann surface?

For Liouvillian functions and, in view of the theorem 36, for functions which are obtained via the repeated resolution of linear differential equations, the answer seems to be negative.

Unravellings.
In relation to a joint project with Aschenbrenner and Van den Dries, which aims to describe the model theory of “real differential fields” and “valuated differential fields”, it seems important to better understand our technique of unravellings in cases where the transseries in do not necessarily admit finite logarithmic depths. A typical example of an equation which is “hard to unravel” is

A good question is whether there are essentially different examples of equations which are hard to unravel. Another question is whether we may avoid the adjusted partial unravellings from section 7.3.

Bibliography

[É92]

J. Écalle. Introduction aux fonctions analysables et preuve constructive de la conjecture de Dulac. Hermann, collection: Actualités mathématiques, 1992.

[vdH97]

J. van der Hoeven. Automatic asymptotics. PhD thesis, École polytechnique, France, 1997.

[vdH00a]

Joris van der Hoeven. A differential intermediate value theorem. Technical Report 2000-50, Univ. d'Orsay, 2000.

[vdH00b]

Joris van der Hoeven. Operators on generalized power series. Journal of the Univ. of Illinois, 2000. Submitted.