Operators on generalized power series

by Joris van der Hoeven

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

November 3, 2023

Abstract

Given a ring and a totally (resp. partially) ordered set of “monomials” , Hahn (resp. Higman) defined the set of power series with well-ordered (resp. Noetherian or well-quasi-ordered) support in . This set can usually be given a lot of additional structure: if is a field and a totally ordered group, then Hahn proved that is a field. More recently, we have constructed fields of “transseries” of the form on which we defined natural derivations and compositions.

In this paper we develop an operator theory for generalized power series of the above form. We first study linear and multilinear operators. We next isolate a big class of so-called Noetherian operators , which include (when defined) summation, multiplication, differentiation, composition, etc. Our main result is the proof of an implicit function theorem for Noetherian operators. This theorem may be used to explicitly solve very general types of functional equations in generalized power series.

1Introduction

In [Hah07], Hahn introduced an abstract framework for algebraic computations on power series with generalized exponents like

One of his main results states that, given a field and a totally ordered monomial group , the set of series with well-ordered support in carries a natural field structure. This result was generalized by Higman [Hig52] to the case of partially ordered monomial monoids .

More recently, Dahn and Göring [DG86] and Écalle [É92] constructed so-called fields of “transseries”, which are fields of generalized power series in the sense of Hahn, with additional structure, such as exponentiation, differentiation, integration, composition, etc. Examples of transseries are

In [vdH97], we have shown how to differentiate, integrate and compose such transseries, and how to solve algebraic differential equations (whenever possible).

In this paper, we will be concerned with the development of an abstract operator theory for generalized power series, in the setting of partially ordered monomial sets introduced by Higman. We start by recalling some basic results about Noetherian orderings (also called well-quasi-orderings) in section 2. In Higman's setting, generalized power series have Noetherian support. For this reason, we shall actually call them Noetherian series.

In section 3, we recall the definition of Noetherian series and develop the theory of strongly linear and strongly multilinear operators. More precisely, it is possible to define a notion of infinite summation on algebras of Noetherian power series. One may think of this as something analoguous to normal summable families in analysis. Strongly linear mappings will then be linear mappings which also preserve infinite summation.

The remainder of this article focuses on the resolution of certain functional equations. Translated into the terminology of operators, this comes down to the isolation of nice classes of operators on which some kind of implicit function theorem holds (actually, we will rather prove “parameterized fixed point theorems”). As a basic example, one would like to solve implicit equations like

(1)

in fields of transseries, where is a sufficiently small parameter (say ) and the unknown.

In section 4, we start by developing a theory of continuous and contracting functions for Noetherian series and we will prove the existence of a solution to equations like (1) using the technique of fixed points. Actually, we will prove an implicit function theorem which is very similar to fixed point theorems from [PC90] and [PCR93], although our proof is more constructive.

A more natural and even more explicit way of getting solutions to (1) would be to replace the left hand side by the right hand side in a recursive manner, while expanding all sums. This would lead to a formal solution of the form

The main difficulty then resides in proving that the obtained formal expansion is indeed summable in our generalized sense. In sections 5 and 6, we will prove that this is indeed the case for a suitable class of “Noetherian operators”.

2Noetherian orderings

Throughout this paper, orderings are understood to be partial, except when we explicitly state them to be total. Actually, almost all ordered sets considered in this paper are monomial sets, and we denote them by fraktur letters . We denote by (or by ) the orderings on such monomial sets. Usually, is even a monomial monoid or group, on which the multiplication is assumed to be compatible with the ordering, i.e.

for all .

Example 1.

  1. with is a totally ordered monomial group.

  2. If and are monomial sets, then their disjoint union is naturally ordered, by taking the orderings on and on each part of the disjoint union, and by taking and mutually incomparable in .

  3. If and are monomial sets, then the Cartesian product is naturally ordered by .

  4. Let be the set of non-commutative words over a monomial set (and where one may think of the elements of as infinitesimals). Such words are denoted by sequences , with . The empty word is denoted by . The set is “naturally” ordered by , if and only if there exists a strictly increasing mapping , such that for all .

Let be a monomial set. A chain in is a subset of which is totally ordered for the induced ordering. An antichain is a subset of of pairwise incomparable elements. The ordering on is said to be well-founded, if there are no infinite sequences of elements in . A Noetherian ordering is a well-founded ordering without infinite antichains.

Remark 2. In the literature, an ordered set is usually said to be well-founded, if there are no infinite sequences of elements in . This definition is compatible with ours, if one interprets a monomial set to be ordered by the opposite ordering of (as we did).

Let be a monomial set. A final segment is a subset of , such that , for all . Given an arbitrary subset of , we denote by the final segment generated by . Dually, an initial segment is a subset of , such that , for all . The following characterizations of Noetherian orderings are classical [Mil85], [Pou85].

Proposition 3. Let be a monomial set. Then the following are equivalent:

  1. The ordering on is Noetherian.

  2. Any final segment of is finitely generated.

  3. The ascending chain condition w.r.t. inclusion holds for final segments of .

  4. Each sequence admits a subsequence .

  5. Any extension of the ordering on to a total ordering on yields a well-ordering.

The most elementary examples of Noetherian orderings are well-orderings, and orderings on finite sets. Proposition 3 allows us to construct more complicated Noetherian orderings from simpler ones:

Proposition 4. Assume that and are Noetherian monomial sets. Then

  1. Any subset of with the induced ordering is Noetherian.

  2. Let be an increasing mapping into a monomial set . Then is Noetherian.

  3. Any extension of the ordering on is Noetherian.

  4. is Noetherian.

  5. is Noetherian.

The following theorem is due to Higman [Hig52]. We will recall a proof due to Nash-Williams [NW63], because a similar proof technique will be used in section 6.1.

Theorem 5. Let be a Noetherian monomial set. Then is Noetherian.

Proof. We say that is a bad sequence in , if there do not exist with . An ordering is Noetherian if and only if there are no bad sequences. Now assume for contradiction that is a bad sequence in . Without loss of generality, we may assume that each is chosen in such that it has minimal length as a word. We say that is a minimal bad sequence.

Now for all , we must have , so we can factorize , where is the first letter of . By proposition 3(d), we can extract a sequence from . Now consider the sequence . By the minimality of , this sequence is good. Hence, there exist and with , or with . But then, resp. . This contradicts the badness of .

3Noetherian series

3.1Noetherian series and infinite summation

Let be a commutative additive group of coefficients and a set of monomials. The support of a mapping is defined by

If is Noetherian for the induced ordering, then we call a generalized power series or a Noetherian series. We denote the set of all Noetherian series with coefficients in and monomials in by . We also write for the coefficient of in such a series and for . Each with is called a term occurring in .

Given two Noetherian series , we define their sum by

This gives the structure of a commutative group. More generally, consider a family of series in . We say that is a Noetherian family, if is Noetherian and for each there exist only a finite number of such that . In that case, we define its sum by

(2)

This sum is again a Noetherian series. In particular, given a series , the family is Noetherian and we have in the sense of (2).

It is useful to see as a strong commutative group, i.e. a commutative group with an additional “infinite summation structure” on it. In our case, this structure is reflected through the infinite summation of Noetherian families; it satisfies the following fundamental properties:

Proposition 6.

  1. Any zero family is Noetherian, and .

  2. For any , the family is Noetherian, and .

  3. If and are Noetherian and , then is Noetherian and .

  4. If is a Noetherian family, then for any bijective mapping , the family is Noetherian, and .

  5. If is a Noetherian family and a decomposition of into pairwise disjoint subsets, then is a Noetherian family for each , is a Noetherian family, and .

Proof. All properties are straightforward to prove. For illustration, we will prove (e). Let be a Noetherian family and let a partition of . For each and , let and , so that

(3)

Now is a Noetherian family for all , since and is finite for all . Furthermore, and for all , the set is finite, because of (3). Hence, the family is Noetherian and for all , we have

This proves (e).

Remark 7. Given two monomial sets and , it is often convenient to identify with via the natural isomorphism

In particular multivariate operators may actually be regarded as a univariate operators . Similarly, given a monomial set , the Noetherian families may be identified with series in , where is strictly ordered by . We may thus view an operator as an operator “in infinitely many variables”, which assigns to each Noetherian family a series in .

3.2Algebras of Noetherian series

Assume now that is a (not necessarily commutative) ring, and a (not necessarily commutative) monomial monoid. Then we may naturally see and as subsets of via resp. . Given and in , we define their product by

The right hand side is well defined by propositions 4(e) and 4(b). Higman [Hig52] first observed that is a ring for this product. Actually, it is even a strong ring, because the product is compatible with the infinite summation structure on in the following way:

Proposition 8. For all Noetherian families and , the family is also Noetherian, and

Proof. First of all,

is Noetherian. Given , the set of couples with forms a finite anti-chain; let denote those couples. Then

is finite, whence is a Noetherian family. Given , we also have

with as above.

Remark 9. Also, if is a Noetherian family, then so is , for each family of scalars.

3.3Extension by strong linearity

Let be a ring and let , be monomial sets. In all what follows, we understand that operates on the left on -modules and -algebras. A linear mapping is said to be strongly additive, if for all Noetherian families , the family is also Noetherian and

Notice that this condition implies that is strongly linear, i.e. , for every Noetherian family and every family of scalars. Notice also that the composition of two strongly linear mappings is again strongly linear.

A mapping is said to be Noetherian, if is a Noetherian family for every Noetherian subset of .

Proposition 10. Let and be -modules of Noetherian series. Then any Noetherian mapping extends to a unique strongly linear mapping .

Proof. Let . By definition, is a Noetherian family, and so is . We will prove that

is the unique strongly linear mapping which coincides with on .

Given and we clearly have . Now let be a Noetherian family and denote . We claim that is a Noetherian family. First of all,

is Noetherian. Secondly, given , the set is finite, since is a Noetherian family. Finally, for each with , the set is also finite, since is a Noetherian family. Hence, the set is finite, which proves our claim. Now our claim, together with proposition 6(d) proves that is a Noetherian family and

This establishes the strong linearity of .

In order to see that is unique with the desired properties, it suffices to observe that for each , we must have by linearity and by strong linearity.

Actually, the above proposition generalizes to the “strongly multilinear” case. If and are monomial sets, then we call a multilinear mapping

strongly multilinear (or strongly multi-additive), if for all Noetherian families , the family is also Noetherian and

In particular, if is a monomial monoid, then the multiplication on is strongly bilinear, by proposition 8. Also, compositions

of strongly multilinear mappings and for are strongly multilinear.

Recall that a mapping is Noetherian, if is a Noetherian family for every Noetherian subset of . The following proposition is proved in a similar way as proposition 10:

Proposition 11. Let and be -modules of Noetherian series. Then any Noetherian mapping extends to a unique strongly multilinear mapping .

Remark 12. In a similar way as we identified with in remark 7, we may see as the strong tensor product of and . We have a natural strongly bilinear mapping . Furthermore, for any strongly bilinear mapping , there exists a unique strongly linear mapping , such that .

3.4Applications of strong linearity

Corollary 13. Let and be monomial monoids and let be a Noetherian mapping which preserves multiplication. Then preserves multiplication.

Proof. The mappings and are both strongly bilinear mappings from into , which coincide on . The result now follows from the uniqueness of strongly bilinear extensions in proposition 11.

Corollary 14. Let be a monomial monoid and a Noetherian mapping, such that for all . Then is a (strong) derivation on .

Proof. The mappings and are both strongly bilinear mappings from into , which coincide on . The result again follows from the uniqueness of strongly bilinear extensions in proposition 11.

Corollary 15. Let and be two Noetherian mappings. Then

Proof. This still follows from the uniqueness of extensions by strong linearity, since and coincide on .

Assume that is a monomial monoid. We call a series infinitesimal, if for all . Then extension by strong linearity may in particular be used to define the composition of a multivariate power series with infinitesimal series . Indeed, if is the multiplicative mapping which sends each to , then we define . Then corollaries 13 and 15 yield the following result:

Corollary 16. Let be infinitesimal Noetherian series in . Then

  1. , for .

  2. , for and infinitesimal .

4The topological implicit function theorem

4.1Truncation of Noetherian series

Let be a monomial set and . Given a subset , we define the restriction of to by

Given two series , we say that is a truncation of (and we write ), if there exists an initial segment of , such that . Thus is an ordering on .

Let be a non-empty family of series. A common truncation of the is a series , such that for all . A greatest common truncation of the is a common truncation, which is greatest for . Such a greatest truncation actually always exists and we denote it by :

Proposition 17. Any non-empty family admits a greatest common truncation.

Proof. Fix some and consider the set of initial segments of , such that for all . We observe that arbitrary unions of initial segments of a given ordering are again initial segments. Hence is an initial segment of each . Furthermore, for each and , there exists an with . Hence for all . This proves that is a common truncation of the . It is also greatest for , since any common truncation is of the form for some initial segment of with .

Let again be a family of series. A common extension of the is a series , such that for all . A least common extension of the is a common extension, which is least for . If such a least common extension exists, then we denote it by .

Now consider a directed index set . In other words, we have an ordering on , such that for any , there exist a with and . Let be a -increasing family of series in , i.e. whenever . If is Noetherian or totally ordered, then there exists a least common extension of the :

Proposition 18. Assume that is Noetherian or totally ordered. Then any directed -increasing family of series in admits a unique least common extension and .

Proof. Let . We claim that is Noetherian. This is clear if is Noetherian. Assume that is totally ordered and that is an infinite sequence of monomials in . Since is directed and whenever , there exist with for each . But we also have for each , so that . Since is Noetherian, the sequence therefore stabilizes.

Given , we claim that the coefficient is independent of the choice of , under the condition that . Indeed, let be such that and , then there exists a with and . Hence, and , so that . Now the series is the least common extension of the .

4.2Stationary limits

Let be a directed index set and a family of series. We call a pseudo-limit of the , if for each final segment of and for all , we have

Equivalently, we may require that for each inital segment of and for each , we have

Assume from now on that is either Noetherian or totally ordered. Below, we will show that the stationary limit of the , which is defined by

is in particular a pseudo-limit. We first prove some useful properties of and .

Proposition 19. Let be a family of series and let be an initial segment of .

  1. If , then

  2. If is directed and -increasing, then

Proof. We first observe that for all we have . In particular, this ensures that exists in (b).

Now assume that and let . Then , whence , for all . This shows that is a common truncation of the . Inversily, assume that is such that for all . Then also for all , so that . Hence . This shows that is the greatest common truncation of the .

Assume now that is directed and -increasing and let. Then , whence , for all . Consequently, is a common extension of the . Furthermore, its support is the same as the support of the least common extension of the . Hence .

Proposition 20. Let be a directed family and . Then

Proof. Since , we have . On the other hand, given , we have for some . Choosing with and , we then have and .

Proposition 21. For any directed family , its stationary limit is a pseudo-limit.

Proof. Let be an initial segment of and let be such that for all . Then proposition 19 implies that

(4)

Hence , by proposition 20.

Given and in , we will write , if for all , there exists an with . The following properties of will be used frequently in the next section:

Proposition 22. Let . Then

  1. if and only if .

  2. .

  3. .

  4. If now stands for a directed family, then

Proof. The first three properties are trivial. Consider the final segment

Then our hypothesis means that for all . Now , by proposition 21. But this means that .

4.3The implicit function theorem

A final segment of a monomial set is said to be attractive, if for each there exists an with . If is totally ordered, then all non-empty final segments are attractive. The intersection of two attractive final segments is again an attractive final segment and arbitrary non-empty unions of attractive final segments are again attractive final segments. In other words, the attractive final subsets of together with the empty set are the open sets of a topology on .

Now let be a commutative additive group. The attractive open subsets of are the subsets of the form , where and where is an attractive final segment of . These sets form a basis for the open subsets of the natural or attractive topology on . We notice that the attractive topology makes an additive topological group. Given another monomial set , we also notice that the attractive topology on (remember remark 7) coincides with the usual product topology on (if and are given the attractive topologies).

Consider a mapping , where . We call contracting, if for all , we have . A contracting mapping is in particular continuous at each point , since for any attractive open neighbourhood of , the set is an open neighbourhood of with .

Theorem 23. Assume that is Noetherian or totally ordered and let be a continuous mapping, such that the mapping is contracting for each . Then there exists a unique mapping with for each , and is continuous.

Proof. Given , consider the transfinite sequence defined as follows:

We will show that converges to a solution of the equation .

The sequence decreases for . Let us prove by (weak) transfinite induction over that for all ordinals . This is clear for . Assume that is a successor ordinal. Since is contracting, the induction hypothesis then implies that for all .

If is a limit ordinal and , then let us prove by a second (weak) transfinite induction over that for all . This is indeed true for , by the first induction hypothesis. Assuming that , we also have

again by the first induction hypothesis and proposition 22(c). If is a limit ordinal, then the second induction hypothesis implies that for all . Hence,

by proposition 22(d).

At this point, we have proved that for all . Now proposition 22(d) implies that

In a similar way, one proves that . Since is contracting, also implies that . Consequently, , by proposition 22(c).

Existence and uniqueness. Having shown that the sequence is decreasing for , we now claim that we must have for some sufficiently large . Otherwise, each of the sets of -maximal monomials of would be non empty, so that for some . Indeed, this will happen as soon as the monomials in get exhausted, i.e. for some such that the cardinality of is the one larger than the cardinality of . Now let . Since , there exists an with . But this contradicts the -maximality of in . This shows our claim and we conclude that the with satisfies .

Assume now that two Noetherian series and both satisfy and . Then , since is contracting. But we can only have if . This establishes the existence and the uniqueness of the mapping .

Continuity. In order to prove that is continuous in any given , let be an attractive open neighbourhood of . Then there exists an attractive open subset of of the form , such that . We claim that . Indeed, let . Taking in our sequence above, it suffices to prove that for all . We prove this by transfinite induction.

For and , we are already done. If , then implies that , whence . If is a limit ordinal, then we have seen above that for all . Taking any such , we also have by the induction hypothesis, whence again and . This completes the induction and the proof of the theorem.

Remark 24. The theorem still holds for monomial sets without “infinite combs” [PCR93]. Our proof also generalizes to this setting, because it can be shown in this case that the stationary limit of a sequence exists, whenever is strictly decreasing for .

Remark 25. Although the above topological implicit function theorem may be very useful to solve certain parameterized functional equations over Noetherian series, one of its major drawbacks is that we needed the very strong Noetherianity assumption on in the partial context. Even the slightly weaker condition about the absence of infinite combs is usually not satisfied. The functional equation

with is an example which shows that there is not much hope for a stronger implicit function theorem in the same spirit. Indeed, the natural “solution” to this equation, which is obtained by recursively replacing the left hand side by the right hand side in the equation, does not have a Noetherian support.

Remark 26. Another drawback of theorem 23, is that it does not provide us with any additional information about the solutions. The solutions may even be quite pathological: consider the monomial group with . Given , we denote . We define a linear (but not strongly linear) operator by

Then it is easily verified that is contracting (whence continuous) on . The equation

will therefore admit a unique solution, which happens to be . However, we do not have .

5Noetherian operators and combinatorial representations

5.1Noetherian operators

Let and be sets of monomials. A Noetherian operator is a mapping , such that there exists a family of strongly multilinear mappings with

(5)

for all Noetherian families . In particular, this assumes that the family of summands is Noetherian. We will call a multilinear decomposition of . The number is the arity of .

By regrouping the of the same arity, it actually suffices to consider the case when and there is exactly one for each arity . In this case, we may write , with for all and . In section 5.4, we will see that this representation is unique, under the assumption that and that the are symmetric (we may always take the to be symmetric if ). However, for the purpose of combinatorial representations in the next section, it is natural to consider more general multilinear decompositions. Notice also that the space of Noetherian operators from has a natural strong group structure.

Remark 27. The formula (5) should hold in particular for families that consist of only one element. In other words, we should have

for all . However, the more complicated assumption (5) is essential, as you will notice in example 31 below.

Remark 28. In view of remark 7 the present definition of Noetherian operators also provides a definition of multivariate Noetherian operators.

Example 29.

Example 30. Let be Noetherian operators.

Example 31. Let and be two Noetherian operators. Then we claim that is also a Noetherian operator. Indeed, let resp. be multilinear decompositions of and . Then for each Noetherian family we have

This establishes our claim, since the operators are strongly multilinear. Notice that example 30 may be looked at as a combination of the present example and the last two cases in example 29.

One obtains interesting subclasses of Noetherian operators by restricting the strongly multilinear mappings involved in the multilinear decompositions to be of a certain type. More precisely, let be a monomial monoid and let be a set of strongly multilinear mappings . We say that is a multilinear type if

MT1

The constant mapping is in for each .

MT2

The -th projection mapping is in for .

MT3

The multiplication mapping from into is in .

MT4

If , then .

Given subsets of , we say that a strongly multilinear mapping

is of type , if for , there exists a mapping in , such that coincides with the restriction of the domain and image of to resp. . We say that a Noetherian operator

is of type , if it admits a multilinear decomposition consisting of strongly multilinear mappings of type only. In examples 30 and 31, we may then replace “Noetherian operator” by “Noetherian operator of type ”.

Example 32. For any set of strongly linear mappings , there exists a smallest multilinear type which contains . Taking to be the field of transseries whose logarithmic and exponential depths are bounded by , interesting special cases are obtained when taking or . Noetherian operators of type resp. may then simply be called differential resp. integral Noetherian operators. Given a finite subset of positive infinitely large transseries in , another interesting case is obtained by taking , where stands for right composition with .

5.2Combinatorial representations of Noetherian operators

Let be a Noetherian operator with a multilinear decomposition . Then is uniquely determined by the action of the on monomials in . For the deeper theory of Noetherian operators, it is convenient to represent this action in a combinatorial way.

Abstractly speaking, a set of -labeled structures is a set , together with a map that assigns to each a labeling , where stands for the size or arity of ; for simplicity, we denote such a set of labeled structures also by . For each subset of , we denote the subset of -labeled structures in by

We strictly order couples in by . A mapping is called a choice operator. We say that is Noetherian, if for any Noetherian subset of , the subset

of is Noetherian.

Example 33. Let be a strictly increasing -ary operation and let , with for all and . Then is a Noetherian choice operator.

Figure 1. Graphical representation of the action of on the structure with input , for the strongly bilinear operator . Notice that .

Returning to our Noetherian operator , we may see each tuple as an -labeled combinatorial structure with and for all . Let denote the set of such structures. We get a natural Noetherian choice operator by taking . Graphically speaking (see figure 1), we may represent the action of on by a box with (a tuple of) “inputs” in and (a set of) “outputs” in .

Inversely, given a Noetherian choice operator and an operator with for all , we define a Noetherian operator by

(6)

As to its multilinear decomposition, we associate an to each by

For Noetherian families , we indeed have

since for each , there are only finitely many tuples , such that .

5.3Composition of choice operators

In example 31, we have shown that the composition of two Noetherian operators and is again Noetherian. Let us now show how to interpret the composition in a combinatorial way. Denote the natural choice operators associated to and by resp. . We first define the composition of the choice operators and . Then , and will be given by (6) and similar formulas, for certain mappings , resp. . Here we may assume that and are given and we have to construct .

Let be given together with a tuple , such that for each . Then these data determine a unique -labeled structure , with and , for all and . We define to be the set of all such combinatorial structures (see figure 2). Then we claim that the choice operator is Noetherian.

So let be a Noetherian subset of . We will prove that for any sequence of elements in the set

there exist with . Since is Noetherian, is a Noetherian subset of , and we observe that for each . Since is Noetherian, we may therefore assume that , modulo the extraction of a subsequence. If for some , then we have and we are done. Hence, we may assume that . We conclude by the observation that given there exist only a finite number of , such that . Indeed, for each , there are only a finite number of with , since is Noetherian.

Now consider the operator . Clearly, for all . We claim that

(7)

for all . Indeed,

This yields the desired combinatorial description of the composition .

Figure 2. Illustration of the action of on a structure in . For each that we attach to , we require the “output” of to coincide with the “input” of .

5.4Canonical multilinear decompositions

We already noticed that each Noetherian operator has a multilinear decomposition of the form , such that has arity for each . Setting for all and , we then have

(8)

Now assume that (so that is in particular torsion-free). Then, modulo replacing each by the operator with

we may assume without loss of generality that the are symmetric. Under this additional symmetry assumption, the decomposition (8) is actually unique, and we call the homogeneous part of of degree .

Proposition 34. Let be a Noetherian operator with a multilinear decomposition , such that is symmetric and of arity for each . If is torsion-free and , then for each .

Proof. We observe that it suffices to prove that for each , since the are symmetric and is torsion-free. Assume the contrary and let be such that for some . Choose is Noetherian. The Noetherianity of implies that there exist only a finite number of indices , such that . Let be those indices.

Let for all . For any , we have , by multilinearity. On the other hand, for each , so that

The matrix on the left hand side admits an inverse with rational coefficients (indeed, by the sign rule of Descartes, a real polynomial cannot have distinct positive zeros unless ). Consequently, an integer multiple of the vector on the right hand side vanishes. We infer that , since is torsion-free. This contradiction completes the proof.

6The algebraic implicit function theorem

Let and be monomial sets and let be a Noetherian operator. We call strictly extensive in if there exists a multilinear decomposition of , such that for all , , and , we have . In particular, such a is contracting in . The main objective of this section will be to prove the following theorem:

Theorem 35. Let be a Noetherian operator, which is strictly extensive in . Then for each the operator on has a unique fixed point , and the operator is Noetherian.

6.1Iteration of choice operators with parameters

Let be as in theorem 35 and let be the natural Noetherian choice operator associated to . The fact that is strictly extensive in implies that may be assumed to be strictly extensive on , i.e.

Also, let be the natural Noetherian choice operator associated to the identity mapping . Actually, we take , with and for all .

Now consider the sets of -labeled combinatorial structures, where the are defined by

For each , the minimal with is called the depth of . We have a natural choice operator , which is defined componentwise by

Here stands for the choice operator which coincides with on and with on . Similarly, the componentwise definition of means that we take . In figure 3 one finds an illustration of the action of on a structure in . We will also call the iteration of with parameters in .

Figure 3. Illustration of the action of the iterated choice operator on a structure in . The connected “inputs” and “outputs” should match in a similar way as in figure 2. The white and black dots correspond to monomials in resp. .

Theorem 36. Let be a set of -labeled structures and a Noetherian choice operator which is extensive on . Then is Noetherian.

Proof. Let be a Noetherian subset of . Assume that there exists a bad sequence

(9)

with and for each . We may assume that we have chosen this bad sequence minimally in the sense that the depth of each is minimal in the set of all bad sequences with fixed . Writing for each , we claim that the induced ordering on is Noetherian.

Indeed, suppose for contradiction that the claim is false, and let

be a bad sequence. Notice that for all , since is strictly extensive on . Hence, taking such that is minimal, the sequence

is also bad. This contradicts the minimality of (9).

At this point we have proved that is Noetherian. In particular, is Noetherian. Hence, there exist with , since . If for some , then and we are done. Otherwise, . Now for every , the with are finite in number, since they form an antichain. Consequently, can only take a finite number of values and there exist with . This contradicts the badness of (9).

6.2Proof of the implicit function theorem

With the notations from the previous section, let be a mapping, such that for all , and such that (6) holds for all . We now define componentwise as follows:

where . Theorem 36 implies that we may define a function by the formula

(10)

We can now prove the following more explicit version of the implicit function theorem.

Theorem 37. Let be a Noetherian operator, which is strictly extensive in . Then the Noetherian operator defined by (10) is unique with the property that for all .

Proof. Identifying and via the natural isomorphism, we have

for all . Similarly, for all , we have

Applying (7), we conclude that

for all . The uniqueness of follows in the same way as in the proof of theorem 23, since is contracting in .

Corollary 38.

Let be a multilinear type. If is of type in theorem 35, then so is .

6.3Applications

Example 39. Let us first show that the classical implicit function theorem for bivariate power series follows from theorem 23. So let be a bivariate power series with and . Then we have to prove that there exists a unique power series with

Modulo division of by and passing to the other side of the equation, the problem can be reduced to solving the equation

(11)

for with . Under these assumptions, the series corresponds to an operator . Theorem 23 then provides us with a unique mapping with . Taking , we thus find the unique solution to (11).

Moreover, theorem 37 actually tells us that the “natural solution” to (11), which is obtained by recursively plugging in the left hand side of the equation in the right hand side, is indeed a solution. We also notice that by applying theorem 37 to the operator

instead of the previous , we actually get a solution in terms of the coefficients of .

Example 40. The above example naturally generalizes to the multivariate case. What is more, we may consider non commutative power series in several variables. Given symbols , we order the free monomial monoid in by the ordering from example 1.4. Then the ring of non commutative power series in over is given by . Now consider the equation

(12)

for with . Then it may be proved in a similar way as in the previous example that this equation admits a unique infinitesimal solution. Again, this solution is equal to the natural expression which is obtained when repreatedly plugging in the left hand side of (12) into the right hand side. Again, the solution may be expressed naturally in terms of the coefficients of the equation.

Example 41. Let be the field of transseries in , whose logarithmic and exponential depths are bounded by some integer [vdH97]. The transseries is an example of an element in if . Now consider the integral equation

(13)

for and where . Taking we may consider the operator . Theorem 23 then implies that there exists a unique function , such that satisfies (13) for all . Theorem 37 and its corollary imply that is actually an integral Noetherian operator. Modulo regrouping terms, this means that the series

is indeed a solution to (13) for all .

Example 42. Let now be the field of transseries in , whose exponential and logarithmic depths are bounded by . Consider the functional equation

(14)

for and . Taking , theorem 37 yields a Noetherian operator , such that is a solution to (14). Moreover, is what one could call a “differential compositional Noetherian operator”.

Example 43. For independent infinitely large variables consider the monomial group

and its subset

Then the equation

(15)

admits a unique solution , which can be expressed as a “partial differential series”. Theorem 23 can not be directly applied in this case.

Bibliography

[DG86]

B. I. Dahn and P. Göring. Notes on exponential-logarithmic terms. Fundamenta Mathematicae, 127:45–50, 1986.

[É92]

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

[Hah07]

H. Hahn. Über die nichtarchimedischen Größensysteme. Sitz. Akad. Wiss. Wien, 116:601–655, 1907.

[Hig52]

G. Higman. Ordering by divisibility in abstract algebras. Proc. London Math. Soc., 2:326–336, 1952.

[Mil85]

E. C. Milner. Basic wqo- and bqo- theory. In Rival, editor, Graphs and orders, pages 487–502. D. Reidel Publ. Comp., 1985.

[NW63]

C. St. J. A. Nash-Williams. On well-quasi-ordering finite trees. Proc. Cambridge Philos. Soc., 59:833–835, 1963.

[PC90]

S. Prieß-Crampe. Der Banachsche Fixpunktsatz für Ultrametrische Raüme. Results in Mathematics, 18:178–186, 1990.

[PCR93]

S. Priess-Crampe and P. Ribenboim. Fixed points, combs and generalized power series. Abh. Math. Sem. Hamburg, 63:227–244, 1993.

[Pou85]

M. Pouzet. Applications of well quasi-ordering and better quasi-ordering. In Rival, editor, Graphs and orders, pages 503–519. D. Reidel Publ. Comp., 1985.

[vdH97]

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