Transserial Hardy fields

Joris van der Hoeven

Dépt. de Mathématiques (bât. 425)

Université Paris-Sud, CNRS

91405 Orsay Cedex

France

Email: vdhoeven@texmacs.org

November 3, 2023

. This work has partially been supported by the ANR Gecko project.

It is well known that Hardy fields can be extended with integrals, exponentials and solutions to Pfaffian first order differential equations . From the formal point of view, the theory of transseries allows for the resolution of more general algebraic differential equations. However, until now, this theory did not admit a satisfactory analytic counterpart. In this paper, we will introduce the notion of a transserial Hardy field. Such fields combine the advantages of Hardy fields and transseries. In particular, we will prove that the field of differentially algebraic transseries over carries a transserial Hardy field structure. Inversely, we will give a sufficient condition for the existence of a transserial Hardy field structure on a given Hardy field.

1.Introduction

A Hardy field is a field of infinitely differentiable germs of real functions near infinity. Since any non-zero element in a Hardy field is invertible, it admits no zeros in a suitable neighbourhood of infinity, whence its sign remains constant. It follows that Hardy fields both carry a total ordering and a valuation. The ordering and valuation can be shown to satisfy several natural compatibility axioms with the differentiation, so that Hardy fields are models of the so called theory of H-fields [AvdD02, AvdD01, AvdD04].

Other natural models of the theory of H-fields are fields of transseries [vdH97, Sch01, MMvdD97, MMvdD99, Kuh00, vdH06]. Contrary to Hardy fields, these models are purely formal, which makes them particularly useful for the automation of asymptotic calculus [vdH97]. Furthermore, the so called field of grid-based transseries (for instance) satisfies several remarkable closure properties. Namely, is differentially Henselian [vdH06, theorem 8.21] and it satisfies the differential intermediate value theorem [vdH06, theorem 9.33].

Now the purely formal nature of the theory of transseries is also a drawback, since it is not a priori clear how to associate a genuine real function to a transseries , even in the case when satisfies an algebraic differential equation over . One approach to this problem is to develop Écalle's accelero-summation theory [Éca85, Éca87, Éca92, Éca93, Bra91, Bra92], which constitutes a more or less canonical way to associate analytic functions to formal transseries with a “natural origin”. In this paper, we will introduce another approach, based on the concept of a transserial Hardy field.

Roughly speaking, a transserial Hardy field is a truncation-closed differential subfield of , which is also a Hardy field. The main objectives of this paper are to show the following two things:

  1. The differentially algebraic closure in of a transserial Hardy field can be given the structure of a transserial Hardy field.

  2. Any differentially algebraic Hardy field extension of a transserial Hardy field, which is both differentially Henselian and closed under exponentiation, admits a transserial Hardy field structure.

We have chosen to limit ourselves to the context of grid-based transseries. More generally, an interesting question is which H-fields can be embedded in fields of well-based transseries and which differential fields of well-based transseries admit Hardy field representations. We hope that work in progress [AvdDvdH05, AvdDvdH] on the model theory of H-fields and asymptotic fields will enable us to answer these questions in the future.

The theory of Hardy fields admits a long history. Hardy himself proved that the field of so called L-functions is a Hardy field [Har10, Har11]. The definition of a Hardy field and the possibility to add integrals, exponentials and algebraic functions is due to Bourbaki [Bou61]. More generally, Hardy fields can be extended by the solutions to Pfaffian first order differential equations [Sin75, Bos81] and solutions to certain second order differential equations [Bos87]. Further results on Hardy fields can be found in [Ros83a, Ros83b, Ros87, Bos82, Bos86]. The theory of transserial Hardy fields can be thought of as a systematic way to deal with differentially algebraic extensions of any order.

The main idea behind the addition of solutions to higher order differential equations to a given transserial Hardy field is to write such solutions in the form of “integral series” over (see also [vdH05]). For instance, consider a differential equations such as

for large . Such an equation may typically be written in integral form

The recursive replacement of the left-hand side by the right-hand side then yields a “convergent” expansion for using iterated integrals

where we understand that each of the integrals in this expansion are taken from :

In order to make this idea work, one has to make sure that the extension of with a solution of the above kind does not introduce any oscillatory behaviour. This is done using a combination of arguments from model theory and differential algebra.

More precisely, whenever a transseries solution to an algebraic differential equation over is not yet in , then we may assume the equation to be of minimal “complexity” (a notion which refines Ritt rank). In section 2, we will show how to put the equation in normal form

(1.1)

where is “small” and admits a factorization

over . In section 4, it will be show how to solve (1.1) using iterated integrals, using the fact that the equation admits as a solution. Special care will be taken to ensure that the constructed solution is again real and that the solution admits the same asymptotic expansion over as the formal solution.

Section 3 contains some general results about transserial Hardy fields. In particular, we prove the basic extension lemma: given a transseries and a real germ at infinity which behave similarly over (both from the asymptotic and differentially algebraic points of view), there exists a transserial Hardy field extension of in which and may be identified. The differential equivalence of and will be ensured by the fact that the equation (1.1) was chosen to be of minimal complexity. Using Zorn's lemma, it will finally be possible to close under the resolution of real differentially algebraic equations. This will be the object of the last section 5. Throughout the paper, we will freely use notations from [vdH06]. For the reader's convenience, some of the notations are recalled in section 2.1. We also included a glossary at the end.

It would be interesting to investigate whether the theory of transserial Hardy fields can be generalized so as to model some of the additional compositional structure on . A first step would be to replace all differential polynomials by restricted analytic functions [vdDMM94]. A second step would be to consider postcompositions with operators for sufficiently flat transseries for which Taylor's formula holds:

This requires the existence of suitable analytic continuations of in the complex domain. Typically, if with , then should be defined on some sector at infinity (notice that this can be forced for the constructions in this paper). Finally, more violent difference equations, such as

generally give rise to quasi-analytic solutions. From the model theoretic point view, they can probably always be seen as convergent sums.

Finally, one may wonder about the respective merits of the theory of accelero-summation and the theory of transserial Hardy fields. Without doubt, the first theory is more canonical and therefore has a better behaviour with respect to composition. In particular, we expect it to be easier to prove o-minimality results [vdD98]. On the other hand, many technical details still have to be worked out in full detail. This will require a certain effort, even though the resulting theory can be expected to have many other interesting applications. The advantage of the theory of transserial Hardy fields is that it is more direct (given the current state of art) and that it allows for the association of Hardy field elements to transseries which are not necessarily accelero-summable.

2.Preliminaries

2.1.Notations

Let be the totally ordered field of grid-based transseries [vdH06]. Any transseries is an infinite linear combination of transmonomials, with grid-based support . Transmonomials are systematically written using the fraktur font. Each transmonomial is an iterated logarithm of or the exponential of a transseries with for each . The asymptotic relations and on are defined by

Given , one also defines variants of , etc. modulo flatness:

It is convenient to use relations as superscripts in order to filter elements, as in

Similarly, we use subscripts for filtering on the support:

We denote the derivation on w.r.t. by and the corresponding distinguished integration (with constant part zero) by . The logarithmic derivative of is denoted by . The operations and of upward and downward shifting correspond to postcomposition with resp. . We finally write if the transseries is a truncation of , i.e. for all .

2.2.Differential fields of transseries and cuts

Given , we define the canonical span of by

(2.1)

By convention, if contains less than two elements. We also define the ultimate canonical span of by

(2.2)

We notice that if and only if admits no minimal element for .

Example 2.1. We have

Consider a differential subfield of and let . We say that has span , if for all and for at least one (notice that we do not require ). Since is stable under differentiation, we have as soon as . Notice also that we must have if has span .

A transseries is said to be a serial cut over , if for every and admits no minimal element for . In that case, let be maximal for such that . Then and are called the head and the tail of . We say that is a normal serial cut if , which implies in particular that .

Assuming that has span , any serial cut over is necessarily in . Conversely, any with is a serial cut over . We will denote by the set of all which are either in or serial cuts over with . Notice that is again a differential subfield of .

The above definitions naturally adapt to the complexifications and of and differential subfields of . If has span , then the set coincides with the set of all which are either in or serial cuts over with .

2.3.Complements on differential algebra

Let be a differential field. We denote by the ring of differential polynomials in over and by its quotient field. Given and , we recall that denotes the homogeneous part of degree of . We will denote by the linear operator in with . Assuming that , we also denote the order of by , the degree of in by and the total degree of by . Thus, the Ritt rank of is given by the pair . The triple will be called the complexity of ; likewise ranks, complexities are ordered lexicographically.

As usual, we will denote the initial and separator of by resp. and set . Given with , Ritt reduction of by provides us with a relation

(2.3)

where is a linear differential operator, and the remainder satisfies .

Let be a differential field extension of . An element is said to be differentially algebraic over if there exists an annihilator with . An annihilator of minimal complexity will then be called a minimal annihilator and is also called the complexity of over . The order of such a minimal annihilator is called the order of over . We say that is a differentially algebraic extension of if each is differentially algebraic over .

We say that is differentially closed in , if contains no elements which are differentially algebraic over .. Given (resp. ), we say that is -differentially closed (resp. -differentially closed) in if (resp. ) for all . We say that is weakly differentially closed if every admits a root in . We say that is weakly -differentially closed if every of order admits a root in .

Given a differential polynomial and , we define the additive and multiplicative conjugates of by :

We have and

We also notice that additive and multiplicative conjugation are compatible with Ritt reduction: given and assuming (2.3), we have

Remark 2.2. The compatibility of Ritt's reduction theory with additive and multiplicative conjugation holds more generally for rings of differential polynomials in a finite number of commutative partial derivations (or with a finite dimensional Lie algebra of non-commutative derivations). Similar compatibility results hold for upward shiftings or changes of derivations (in the partial case, this requires the rankings to be order-preserving).

In the case when is a differential subfield of , we recall that a differential polynomial may also be regarded as a series in . Similarly, elements of the fraction field of may be regarded as series with coefficients in . Indeed, writing and , where denotes the dominant term of , we may expand

In the case when for some transbasis , then and may also be expanded lexicographically with respect to .

2.4.Linear differential operators and factorization

Let be a differential field and consider a linear differential operator . We will denote the order of by . Given , we define the multiplicative conjugate and the twist by

We notice that is also obtained by substitution of for in . We say that splits over , if it admits a complete factorization

(2.4)

with . In that case, each of the twists of also splits:

We say that is -linearly closed if any linear differential operator of order splits over .

Proposition 2.3. If is weakly -differentially closed, then s -linearly closed.

Proof. The proof proceeds by induction over . For , we have nothing to prove, so assume that and let be of order . Then the differential Riccati polynomial has order , so it admits a root . Division of by in yields a factorization where has order . By the induction hypothesis, splits over , whence so does .

Proposition 2.4. Let be an operator which splits over and let be such that . Then and split over .

Proof. Recall that greatest common divisors and least common multiples exist in the ring . Given a splitting (2.4), consider the operators

We have and . Moreover, the orders of and (resp. and ) differ at most by one for each . It follows that and split over .

Assume now that is a totally ordered differential field. A monic operator is said to be an atomic real operator if has either one of the forms

A real splitting of an operator over is a factorization of the form

(2.5)

where each is an atomic real operator. A splitting (2.4) over is said to preserve realness, if it gives rise to a real splitting (2.5) for or and .

Proposition 2.5. Let be an operator which splits over . Then admits a real splitting over .

Proof. Assuming that , we claim that there exists an atomic real right factor of . Consider a splitting (2.4) over . If , then we may take . Otherwise, we write

and take to be the least common multiple of and in . Since , we indeed have . Since and , we also have . In particular, proposition 2.4 implies that splits over . Such a splitting is necessarily of the form

whence is atomic. Having proved our claim, the proposition follows by induction over . Indeed, let be such that . By proposition 2.4, splits over . By the induction hypothesis, therefore admits a real splitting over . But then is a real splitting of .

Corollary 2.6. An operator is atomic if and only if is irreducible over and splits over .

2.5.Factorization at cuts

Let be a differential subfield of of span . Given and , we say that splits over at , if and have the same order and splits over .

Lemma 2.7. Let be a differential subfield of of span . Let be a minimal annihilator of a differentially algebraic cut over , which splits over at . Then any minimal annihilator of over splits over at .

Proof. Since , Ritt division of by yields

(2.6)

for some and . Additive conjugation of (2.6) yields

(2.7)

By the minimality hypothesis for , we have and , so that and . Similarly, we have . Consequently, when considering the linear part of the equation (2.7), we obtain

whence divides in . Now splits over , whence so does . By proposition 2.4, we infer that splits over . Since , we also have and we conclude that splits over at .

Corollary 2.8. Let be a differential subfield of of span . Let be a minimal annihilator of a differentially algebraic cut over , which splits over at . Then any minimal annihilator of over splits over at .

Proof. Applying the lemma to , we see that splits over . Now , whence and also split over .

Lemma 2.9. Let be a differential subfield of of span , such that is -linearly closed. Let be a minimal annihilator of a differentially algebraic cut over , such that has order . Assume that and let be a minimal annihilator of over . Then splits over at .

Proof. Let be as in the above corollary, so that splits over at . Since has minimal complexity and , Ritt division of by yields

for some and . Additive conjugation and extraction of the linear part yields

so divides in . Since the separants of and don't vanish at , we have

and

Consequently, the quotient of and has order at most , whence it splits over . It follows that splits over and splits over at .

2.6.Normalization of linear operators

Let be a differential subfield of of span . Recall from [vdH06, Section 7.7] that with admits a canonical fundamental system of oscillatory transseries solutions with . We will denote by the set of dominant monomials of . The neglection relation on is extended to by if and only if with and .

We say that is normal, if we have or for each . In that case, any quasi-linear equation of the form

with admits as its only solution in . If is a first order operator of the form , then is normal if and only if for some or . In particular, we must have and .

Proposition 2.10. Let .

  1. There exists a such that is normal.

  2. If is normal and , then is normal.

Proof. Let . For each , the operator admits as solutions, which implies in particular that . Now for all . Choosing sufficiently large, it follows that for all with , so that is normal. Similarly, if for some with , then for all .

Proposition 2.11. Consider a normal operator , which admits a splitting

with . Then each is a normal operator.

Proof. We will call normal, if is normal. Let us first prove the following auxiliary result: given and such that and are normal and , then is also normal. If , then , whence . In the other case, we have . Now if , then , since . If , then implies , whence . It again follows that .

Let us now prove the proposition by induction over . For , we have nothing to do, so assume that . Since is normal, the induction hypothesis implies that is normal for all . Now let be the unique element in . Since is normal, is also normal for , by the auxiliary result. We conclude that is normal, since .

Let and be as above. The smallest real number with for all will be called the growth rate of , and we denote . For all , we notice that .

Proposition 2.12. Let be operators of the same order with

Then .

Proof. Given , we have

since . In particular, , whence and .

Proposition 2.13. Given a splitting

with , we have for all .

Proof. Assume for contradiction that for some and choose maximal with this property. Setting

the transseries

satisfies , as well as . But such an cannot be a linear combination of the with .

Remark 2.14. It can be shown (although this will not be needed in what follows) that an operator splits over if and only if there exists an approximation with which splits over for every . In particular, is -linearly closed if and only if is -linearly closed over .

2.7.Normalization of quasi-linear equations

Assume now that is a differential subfield of of span . We say that is normal if is normal of order and . In that case, the equation

(2.8)

is quasi-linear and it admits a unique solution in . Indeed, let be the distinguished solution to (2.8). By proposition 2.12, the operator is normal. If were another solution to (2.8), then would be in , whence , which is impossible.

Proposition 2.15. Let be a differential subfield of of span . Let be a minimal annihilator of a differentially algebraic cut over . Then there exists a truncation and such that is normal.

Proof. Let and . Modulo a multiplicative conjugation by for some , we may assume without loss of generality that . Modulo an additive conjugation by , we may also assume that . For any and , we have

whence

(2.9)

Since , we have . By proposition 2.10, there exists a for which is normal. Now take . Denoting , proposition 2.12 and (2.9) imply that is normal with and .

We say that is split-normal, if is normal and can be decomposed such that splits over and . In that case, we may also decompose for with . If is monic, then we say that is monic split-normal. Any split-normal equation (2.8) is clearly equivalent to a monic split-normal equation of the same form.

Proposition 2.16. Let be a differential subfield of of span such that is -linearly closed. Let be a minimal annihilator of a differentially algebraic cut of order over . Let be a minimal annihilator of and assume that . Then there exists a truncation and such that is split-normal.

Proof. By proposition 2.15 and modulo a replacement of by , we may assume without loss of generality that is normal. By lemma 2.9, splits over at . Let be such that

Setting , we notice that . Now take

Then and proposition 2.12 implies that is normal, with . Denoting , we finally have .

3.Transserial Hardy fields

3.1.Transserial Hardy fields

Let be the field of grid-based transseries [vdH06] and the set of infinitely differentiable germs at infinity. A transserial Hardy field is a differential subfield of , together with a monomorphism of ordered differential -algebras, such that

TH1

For every , we have .

TH2

For every , we have .

TH3

There exists an , such that for all .

TH4

The set is stable under taking real powers.

TH5

We have for all with .

In what follows, we will always identify with its image under , which is necessarily a Hardy field in the classical sense. The integer in TH3 is called the depth of ; if for all , then the depth is defined to be . We always have , since is stable under differentiation. If , then is exponential for all and contains . If and , then contains for all sufficiently large .

Example 3.1. The field is clearly a transserial Hardy field. As will follow from theorem 3.12 below, other examples are

Remark 3.2. Although the axioms TH4 and TH5 are not really necessary, TH4 allows for the simplification of several proofs, whereas it is natural to enforce TH5. Notice that TH5 automatically holds for with since

whence for some . Since both and are infinitesimal in , we have . Consequently, it suffices to check TH5 for monomials with .

Proposition 3.3. Let be a transserial Hardy field with . Then the upward shift of carries a natural transserial Hardy field structure with .

Proof. The field is stable under differentiation, since for all .

Corollary 3.4. If has depth , then is a transserial Hardy field of depth .

We recall that a transbasis is a finite set of transmonomials with

TB1

and .

TB2

for some .

TB3

for all .

If , then is called a plane transbasis and is stable under differentiation. The incomplete transbasis theorem for also holds for transserial Hardy fields:

Proposition 3.5. Let be a transbasis and . Then there exists an supertransbasis of with . Moreover, if is plane and is exponential, then may be taken to be plane.

Proof. The same proof as for [vdH06, Theorem 4.15] may be used, since all field operations, logarithms and truncations used in the proof can be carried out in .

Given a set of exponential transseries in , the transrank of is the minimal size of a plane transbasis with . This notion may be extended to allow for differential polynomials in (modulo the replacement of by its set of coefficients).

Remark 3.6. The span and ultimate span of are not necessarily in . Nevertheless, if and is a transbasis for , then we do have for some (and similarly for the ultimate span of ).

3.2.Cuts in transserial Hardy fields

Let be a transserial Hardy field. Given and , we write if there exists a with

We say that and are asymptotically equivalent over if for each (or, equivalently, for each ), we have

We say that and are differentially equivalent over if

for all .

Lemma 3.7. Let be a transserial Hardy field and let be differentially algebraic over . Let be maximal for , such that . Then is differentially algebraic over and .

Proof. Let be a minimal annihilator of . Modulo upward shifting, we may assume without loss of generality that and are exponential. Since , all monomials in are in , whence there exists a plane transbasis for and . Modulo subtraction of from and , we may assume without loss of generality that . Let be such that and let be the dominant monomial of . Modulo division of and by , we may also assume that is a normal serial cut. But then the equation gives rise to the equation for . The complexity of is clearly bounded by .

Lemma 3.8. Let be a transserial Hardy field and . Let and be such that and are both asymptotically and differentially equivalent over . Then and are both asymptotically and differentially equivalent over .

Proof. Given , we either have and

or , in which case

This proves that and are asymptotically equivalent over .

As to their differential equivalence, let us first assume that is differentially transcendent over . Given , let us denote

We have , and

(3.1)
(3.2)

whence and .

Assume now that is differentially algebraic over and let be a minimal annihilator. Given , Ritt reduction of w.r.t. gives

where and is such that . Since and , we both have and , whence

If , this clearly implies . Otherwise, vanishes neither at nor at and the relations (3.1) and (3.2) again yield and .

Lemma 3.9. Let be a transserial Hardy field and let be a differentially algebraic cut over with minimal annihilator . Let be a root of such that and are asymptotically equivalent over . Then and are differentially equivalent over .

Proof. Let be such that . Modulo some upward shiftings, we may assume without loss of generality that and are exponential. Modulo an additive conjugation by and a multiplicative conjugation by , we may also assume that is a normal cut. Modulo a division of by and replacing by , we may finally assume that .

Now consider with . Since , there exists a with and . But then

For general , we use Ritt reduction of w.r.t. and conclude in a similar way as in the proof of lemma 3.8.

3.3.Elementary extensions

Lemma 3.10. Let and be such that

  1. is a serial cut over .

  2. and are asymptotically equivalent over .

  3. and are differentially equivalent over .

Then carries the structure of a transserial Hardy field for the unique differential morphism over with .

Proof. Modulo upward shifting, an additive conjugation by and a multiplicative conjugation by , we may assume without loss of generality that is an exponential normal serial cut. Let be such that . We have to show that is closed under truncation and that for all with (this implies in particular that is increasing). Notice that implies .

Truncation closedness. Given , let us prove by induction over the transrank of that . So let be a plane transbasis for and . Assume first that . Writing

the sum

is finite, whence

By the induction hypothesis, we also have and . If , then

for a sufficiently large truncation , whence .

Preservation of dominant terms. Given with , let us prove by induction over the transrank of that . Let be a plane transbasis for and and assume first that . Since , there exists a maximal with , when considering as a series in . But then

by the induction hypothesis. If , then there exists an such that, for all sufficiently large truncations , the Taylor series expansion of yields

Taking such that , we obtain

This completes the proof.

Theorem 3.11. Let be a transserial Hardy field. Then its real closure admits a unique transserial Hardy field structure which extends the one of .

Proof. Assume that and choose of minimal complexity. By lemma 3.7, we may assume without loss of generality that is a serial cut. Consider the monic minimal polynomial of . Since , we have

for a sufficiently large truncation of (we refer to [vdH06, Section 8.3] for a definition of the Newton degrees ). But then

(3.3)

admits unique solutions and in resp. , by the implicit function theorem. It follows in particular that . Let and consider with . Then

Since , we obtain , whence and are asymptotically equivalent over . By lemmas 3.9 and 3.10, it follows that carries a transserial Hardy field structure which extends the one on . Since (3.3) has a unique solution in , this structure is unique. We conclude by Zorn's lemma.

3.4.Exponential and logarithmic extensions

Theorem 3.12. Let be a transserial Hardy field and let be such that . Then the set carries the structure of a transserial Hardy field for the unique differential morphism over with for all .

Proof. Each element in is of the form for and -linearly independent . Given , let be a transbasis for . We may write

with (or the obvious adaptations if or ). Modulo the substitution of by , we may assume without loss of generality that .

If , then we may regard as a convergent grid-based series in with coefficients in . In particular,

Furthermore, if admits as its dominant exponent in , then holds both in and in .

If , then we may consider as a series

in . Since is closed under truncation, both and lie in , whence

by what precedes. Similarly, if is the dominant term of as a series in and is the dominant term of as a series in (with ), then holds both in and in .

This shows that is truncation closed and that the extension of to is increasing. We also have . In other words, is a transserial Hardy field.

Theorem 3.13. Let be a transserial Hardy field of depth . Then carries the structure of a transserial Hardy field for the unique differential morphism over with for all .

Proof. The proof is similar to the proof of theorem 3.12, when replacing by .

3.5.Complex transserial Hardy fields

Let be a transserial Hardy field. Asymptotic and differential equivalence over are defined in a similar way as over .

Proposition 3.14. Let be a transserial Hardy field. Let be a serial cut over and . Then and are asymptotically equivalent over if and only if and as well as and are asymptotically equivalent over .

Proof. Assume that and are asymptotically equivalent over and let . Consider . We have , so that . Moreover, , whence and . The relation is proved similarly. Inversely, assume that and as well as and are asymptotically equivalent over . Given , we have , whence there exist with and . It follows that , whence .

Proposition 3.15. Let be a transserial Hardy field, and . Then and are differentially equivalent over if and only if they are differentially equivalent over .

Proof. Differential equivalence over clearly implies differential equivalence over . Assuming that and are differentially equivalent over , we also have

for every .

Remark 3.16. Given and , it can happen that and are differentially equivalent over , without and being differentially equivalent over . This is for instance the case for , and . Indeed, the differential ideals which annihilate resp. are both .

Most results from the previous sections generalize to the complex setting in a straightforward way. In particular, lemmas 3.7, 3.8 and 3.9 also hold over . However, the fundamental extension lemma 3.10 admits no direct analogue: when taking and such that the complexified conditions i, ii and iii hold, we cannot necessarily give the structure of a transserial Hardy field. This explains why some results such as lemmas 2.9 and 2.16 have to be proved over instead of . Of course, theorem 3.11 does imply the following:

Theorem 3.17. Let be a transserial Hardy field. Then there exists a unique algebraic transserial Hardy field extension of such that is algebraically closed.

4.Analytic resolution of differential equations

Recall that stands for the differential algebra of infinitely differentiable germs of real functions at . Given , we will denote by the differential subalgebra of infinitely differentiable functions on . We define a norm on by

Given , we also denote and define a norm on by

Notice that

An operator (resp. ) is said to be continuous if there exists an with (resp. ) for all . The smallest such is called the norm of and denoted by (resp. ). The above definitions generalize in an obvious way to the complexifications and .

4.1.Continuous right-inverses of first order operators

Let be a transserial Hardy field of span . Consider a normal operator with and let be sufficiently large such that does not change sign on . We define a primitive of by

Decomposing , we are either in one of the following two cases:

  1. The repulsive case when .

  2. The attractive case when both and .

Notice that the hypothesis implies .

Proposition 4.1. The operator , defined by

(4.1)

is a continuous right-inverse of on , with

(4.2)

Proof. In the repulsive case, the change of variables yields

It follows that

for all , whence (4.2). In the attractive case, the change of variables leads in a similar way to the bound

for all , using the monotonicity of . Again, we have (4.2).

Corollary 4.2. In the attractive case, the operator

is a continuous right-inverse of on , for any .

4.2.Continuous right-inverses of higher order operators

Let be a transserial Hardy field of span . A monic operator is said to be split-normal, if it is normal and if it admits a splitting

(4.3)

with . In that case, proposition 2.11 implies that each is a normal first order operator. For a sufficiently large , it follows that admits a continuous “factorwise” right-inverse on , where . We have

Proposition 4.3. is a continuous operator for every .

Proof. Given , the the first derivatives of satisfy

with

By proposition 2.13 and induction over , we have for all . Since , it follows that

(4.4)

for all and , where

We conclude that

Proposition 4.4. If and the splitting (4.3) preserves realness, then preserves realness in the sense that it maps into itself.

Proof. It clearly suffices to prove the proposition for an atomic real operator . If has order , then the result is clear. Otherwise, we have

for certain . In particular, we are in the same case (attractive or repulsive) for both factors of . Setting , let be as in the previous section. Consider and . In the repulsive case, we have

In particular, we have , whence , since satisfies the differential equation of order with real coefficients. In the attractive case, we have

so that . Since , the difference satisfies . Now is the only solution with to the equation . This proves that .

4.3.The fixed point theorem

Let be a transserial Hardy field of span and consider a monic split-normal quasi-linear equation

(4.5)

where has order and has degree . Of course, we understand that is a monic split-normal operator with . We will denote by the valuation of in (i.e. for and ). We will show how to construct a solution to (4.5) using the fixed-point technique.

Proposition 4.5. Given with , let be a continuous factorwise right-inverse of beyond and consider the operator

(4.6)

on . Then there exists a constant with

(4.7)

for all .

Proof. Consider the Taylor series expansion

Since for all and , we may define by

(4.8)

and obtain

On the other hand, for each with , we have

where

(4.9)

Consequently, the proposition holds for .

Theorem 4.6. Let (4.5) be a monic split-normal equation and let be such that . Then for any sufficiently large , there exists a continuous factorwise right-inverse of , such that the operator (4.6) satisfies

(4.10)

for all

Moreover, taking such that , the sequence tends to a unique fixed point for the operator .

Proof. Since for all , the number from (4.8) tends to for . When constructing using proposition 4.1, the number from (4.9) decreases as a function of . Taking sufficiently large so that , we obtain (4.10). By induction over , it follows that

Now let be the space of times continuously differentiable functions on , such that are bounded. This space is complete, whence converges to a limit . Since this limit satisfies the equation (4.5), the function is actually infinitely differentiable, i.e. .

4.4.Asymptotic analysis

With the notations from the previous section, assume now that is -differentially closed in , i.e. any solution to an equation with is already in . Each is the right-inverse of an operator with . Now also admits a formal distinguished right-inverse . Consequently, the operator also admits a formal counterpart

For each , we have

so the sequence also admits a formal limit in . In order to show that the fixed point from proposition 4.6 and are asymptotically equivalent over , we need some further notations. Given and , let us write if , i.e. for all . We also write if .

Proposition 4.7. For , and , we have

Proof. Trivial.

Proposition 4.8. For , and with , we have

Proof. Let us first show that

(4.11)

Given with , we have , whence . Moreover,

(4.12)

whence for some fixed . This proves (4.11). More generally, additional applications of (4.12) yield

Now assume that and write

By what precedes, we have . On the other hand,

for some . Since is normal, we either have (in which case for all ) or . In both cases, we get , so that .

Theorem 4.9. Let be a transserial Hardy field of span such that is -differentially closed in . Consider a monic split-normal quasi-linear equation (4.5) without solutions in . Then there exist solutions and to (4.5), such that and are asymptotically equivalent over .

Proof. With the above notations, let and be the limits in resp. of the sequences resp. . Given , there exists an with

At that point, we have

In other words, and are asymptotically equivalent over .

Theorem 4.10. Let be a transserial Hardy field of span . Consider a monic split-normal quasi-linear equation (4.5) without solutions in such that and have coefficients in . Assume that one of the following conditions holds:

  1. is -differentially closed in and .

  2. is -differentially closed in .

Then there exist solutions and to (4.5), such that and are asymptotically equivalent over .

Proof. In view of propositions 2.5 and 4.4, we may assume that and preserve realness in all results from sections 4.3 and 4.4. In particular, the solutions and in the conclusion of theorem 4.9 are both real.

5.Differentially algebraic Hardy fields

5.1.First order extensions

Lemma 5.1. Let be a transserial Hardy field of span . Let be a normal operator. Let and be such that is transcendental over and . Then there exists an with , such that and are both differentially and asymptotically equivalent over .

Proof. With the notations of section 4.1, let . Given a truncation , we claim that

Indeed, consider

In the attractive case, implies . In the repulsive case, we have and again . By proposition 4.8, we also have

Since , it follows that , whence and are asymptotically equivalent over . Furthermore, is a minimal annihilator of over , since is transcendental over . Lemma 3.9 therefore implies that and are differentially equivalent over .

Theorem 5.2. Let be a transserial Hardy field. Let be the smallest differential subfield of , such that for any with and we have . Then the transserial Hardy field structure of can be extended to .

Proof. By theorems 3.11, 3.12 and 3.13, we may assume that is closed under the resolution of real algebraic equations, exponentiation and logarithm. Assume that and let be of minimal complexity , such that for some . Without loss of generality, we may make the following assumptions:

By Zorn's lemma, it suffices to show that carries the structure of a transserial Hardy field, which extends the structure of .

If , then lemma 5.1 implies the existence of an such that and are both asymptotically and differentially equivalent over . Hence, the result follows from lemmas 3.8 and 3.10.

If , then and are -differentially closed in resp. . Now , since is exponential. Therefore, theorem 4.10 provides us with an with , such that and are asymptotically equivalent over . We conclude by lemmas 3.9, 3.8 and 3.10.

5.2.Higher order extensions

Lemma 5.3. Let be a transserial Hardy field of span . Let be a normal operator. Let and be such that has order over and . Then there exists an with , such that and are both differentially and asymptotically equivalent over .

Proof. The fact that and are asymptotically equivalent over is proved in a similar way as for lemma 5.1. It follows in particular that and are asymptotically equivalent. Since annihilates , , and , it also annihilates both and . The fact that has complexity over now guarantees that is a minimal annihilator of . We conclude by lemma 3.9.

Theorem 5.4. Let be a transserial Hardy field. Let be the smallest differential subfield of , such that for any and we have . Then the transserial Hardy field structure of can be extended to .

Proof. By theorems 3.12, 3.13 and 5.2, we may assume that is closed under exponentiation, logarithm and the resolution of first order differential equations. Assume that and let be of minimal complexity , such that for some with . Let be a minimal annihilator of and notice that , since . Without loss of generality, we may make the following assumptions:

By Zorn's lemma, it now suffices to show that carries the structure of a transserial Hardy field, which extends the structure of .

If , then lemma 5.3 and the fact that is -differentially closed imply the existence of an such that and are both asymptotically and differentially equivalent over . The result follows by lemmas 3.8 and 3.10.

If , then and are -differentially closed in resp. . Now , since is exponential. Therefore, theorem 4.10 provides us with a with , such that and are asymptotically equivalent over . We conclude by lemmas 3.9, 3.8 and 3.10.

Corollary 5.5. There exists a transserial Hardy field , such that for any and with and , there exists a with and .

Proof. Take and endow it with a transserial Hardy field structure. Let and with be such that . By [vdH06, Theorem 9.33], there exists a with and . But implies .

Corollary 5.6. There exists a transserial Hardy field , such that is weakly differentially closed.

Proof. Take . By a straightforward adaptation of [vdH06, Chapter 8] (see also [vdH01, theorem 9.3]), it can be shown that any differential equation of degree with admits distinguished solutions in when counting with multiplicities. Let be such a solution. Since , both and are differentially algebraic over , whence .

Corollary 5.7. There exists a differentially Henselian transserial Hardy field , i.e., such that any quasi-linear differential equation over admits a solution in .

5.3.Differential Newton polynomials for Hardy fields

Let be a differentially algebraic Hardy field extension of a transserial Hardy field .

Proposition 5.8. Given , there exists an with .

Proof. The functional inverse of satisfies an algebraic differential equation over . Let be the leading term of for its logarithmic decomposition. As in [vdH06, Section 8.1.4]. there exists an with for all . It follows that and .

Given a differential polynomial , we define its dominant part to be the unique monic such that for some and . Here is said to be monic if its leading coefficient w.r.t. equals .

Theorem 5.9. Given , there exists a polynomial with

for all sufficiently large .

Proof. As in the proof of [vdH06, Theorem 8.6], we have

so we may assume without loss of generality that is constant for all . Now

whence

(5.1)
(5.2)
(5.3)

Indeed, we must have

because would imply . Applying [vdH06, Lemma 8.5] to (5.2), and similarly for , we get

for all .

By proposition 5.8 and (5.3), we have and for some . Modulo upward shiftings, we may thus assume without loss of generality that . More generally, assume that for some . By (5.3), this implies for all and

(5.4)

for all of weight . We claim that there exists an with

(5.5)

Assume the contrary and consider a coefficient of weight with

for all . Without loss of generality, we may assume that and are in . Then proposition 5.8 implies and even (by integrating from when possible). Again by proposition 5.8, it follows that and for some . But then (5.4) yields

which contradicts the fact that . The relations (5.5) and (5.4) imply the existence of an with . By induction over and modulo upward shiftings, we may thus ensure that for all .

The polynomial in theorem 5.9 is called the differential Newton polynomial of . The generalization of this concept to allows us to mimic a lot of the theory from [vdH06, chapter 8] in . In what follows, we will mainly need a few more definitions. The Newton degree of an equation

(5.6)

with and is defined by . Setting

we also define

We say that is a solution to (5.6) modulo if . We say that is differentially Henselian, if every quasi-linear equation over admits a solution. Given a solution to (5.6), we say that has algebraic type if is not homogeneous and differential type in the other case. The following proposition is proved along the same lines as [vdH06, proposition 8.16]:

Proposition 5.10. Let be a solution to (5.6) of differential type and let be the degree of . Then is a solution modulo of .

Remark 5.11. In this section, we assumed that is a differentially algebraic Hardy field extension of a transserial Hardy field . We expect that the theory can be adapted to even more general H-field. This is one of the objectives of a current collaboration with Lou van den Dries and Matthias Aschenbrenner [AvdDvdH].

5.4.Transserial models of differentially algebraic Hardy fields

Theorem 5.12. Let be a transserial Hardy field and a differentially algebraic Hardy field extension of , such that is differentially Henselian and stable under exponentiation. Then there exists a transserial Hardy field structure on which extends the structure on .

Proof. By theorems 3.11, 3.12 and 5.2, we may assume that is closed under the resolution of real algebraic equations, exponentiation and integration. Assume that and choose of minimal complexity , such that either

C1

for some .

C2

modulo for some , and admits no roots in modulo . Moreover, is -differentially closed in .

Modulo upward shifting, we may assume without loss of generality that is exponential. In view of Zorn's lemma, it suffices to show that there exists a transserial Hardy field structure on which extends the structure on .

Let be the set of such that . The set is totally ordered for , so there exists a minimal well-based transseries with for all . We call the initializer of over . Assume first that . Then we may assume without loss of generality that , modulo an additive conjugation by . Now is of differential type, since for no . Let be such that modulo . Since has lower complexity than , there exists a with modulo . Since is truncation closed we may take . But then . This contradiction proves that we cannot have .

Let us now consider the case when . Since , there exists a root of in the set of well-based transseries with complex coefficients. But admits only grid-based solutions, whence . By construction, and are asymptotically equivalent over . Let be such that . Modulo an additive and a multiplicative conjugation we may assume without loss of generality that is a normal cut. In case C2, we notice that , whence , since . Consequently, we always have .

We claim that the cuts and are differentially equivalent over . Assume the contrary and let be a minimal annihilator of . By lemma 2.15 and modulo an additive and multiplicative conjugation, we may assume without loss of generality that and that is normal. Since is differentially Henselian, it follows that admits a root in . Now in case C1 and in case C2, so this root is already in , by the induction hypothesis. But admits at most one solution in , whence . This contradiction completes the proof of our claim. By lemma 3.10, we conclude that carries the structure of a transserial Hardy field extension of .

Corollary 5.13. Let be a transserial Hardy field and a differentially algebraic Hardy field extension of , such that is differentially Henselian. Assume that admits no non-trivial algebraically differential Hardy field extensions. Then satisfies the differential intermediate value property.

Proof. The fact that admits no non-trivial algebraically differential Hardy field extensions implies that is stable under exponentiation. By theorem 5.12, we may give the structure of a transserial Hardy field. By theorem 5.4, we also have . We conclude in a similar way as in the proof of corollary 5.5.

It is quite possible that there exist maximal Hardy fields whose differentially algebraic parts are not differentially Henselian, although we have not searched hard for such examples yet. The differentially algebraic part of the intersection of all maximal Hardy fields is definitely not differentially Henselian (and therefore does not satisfy the differential intermediate value property), due to the following result [Bos87, Proposition 3.7]:

Theorem 5.14. Any solution of the equation

is contained in a Hardy field. However, none of these solutions is contained in the intersection of all maximal Hardy fields.

Glossary

is dominated by 4

is negligible w.r.t. 4

is asymptotic to 4

is asymptotically similar to 4

is flatter than or as flat as 4

is flatter than 4

is as flat as 4

and are similar modulo flatness 4

modulo elements flatter than 4

modulo elements flatter than 4

modulo elements flatter than or as flat as 4

modulo elements flatter than or as flat as 4

shorthand for 4

shorthand for 4

shorthand for 4

infinite part of 4

part of which is flatter than 4

shorthand for 4

shorthand for 4

derivation with respect to 4

integration with respect to 4

logarithmic derivative of 4

upward shifting 4

downward shifting 4

is a truncation of 4

canonical span of 4

ultimate canonical span of 4

completion of with serial cuts 5

ring of differential polynomials in over 5

quotient field of 5

linear part of as an operator 5

order of 5

degree of in its leader 5

total degree of 5

complexity of 5

initial of 5

separant of 5

the product 5

complexity of over 5

order of over 5

additive conjugation of by 6

multiplicative conjugation of by 6

multiplicative conjugate of by 6

twist of by 6

set of dominant monomials of solutions to 9

ring of infinitely differentiable germs at infinity 12

is asymptotically similar to over 13

real closure of 15

Newton degree of modulo 15

norm of for 17

shorthand for 17

norm of and its first derivatives for 17

operator norm for 18

operator norm for 18

first order differential closure of in 23

differentially algebraic closure of in 24

Bibliography

[AvdD01]

M. Aschenbrenner and L. van den Dries. Liouville closed -fields. J. Pure Appl. Algebra, 2001. to appear.

[AvdD02]

M. Aschenbrenner and L. van den Dries. -fields and their Liouville extensions. Math. Z., 242:543–588, 2002.

[AvdD04]

M. Aschenbrenner and L. van den Dries. Asymptotic differential algebra. In Contemporary Mathematics. Amer. Math. Soc., Providence, RI, 2004. To appear.

[AvdDvdH]

M. Aschenbrenner, L. van den Dries, and J. van der Hoeven. Linear differential equations over H-fields. In preparation.

[AvdDvdH05]

Matthias Aschenbrenner, Lou van den Dries, and J. van der Hoeven. Differentially algebraic gaps. Selecta Mathematica, 11(2):247–280, 2005.

[Bos81]

M. Boshernitzan. An extension of Hardy's class l of 'orders of infinity'. J. Analyse Math., 39:235–255, 1981.

[Bos82]

M. Boshernitzan. New 'orders of infinity'. J. Analyse Math., 41:130–167, 1982.

[Bos86]

M. Boshernitzan. Hardy fields, existence of transexponential functions. Æquationes Math., 30:258–280, 1986.

[Bos87]

M. Boshernitzan. Second-order differential equations over Hardy fields. J. London Math. Soc., 35:109–120, 1987.

[Bou61]

N. Bourbaki. Fonctions d'une variable réelle. Éléments de Mathématiques (Chap. 5. Hermann, 2-nd edition, 1961.

[Bra91]

B.L.J. Braaksma. Multisummability and Stokes multipliers of linear meromorphic differential equations. J. Diff. Eq, 92:45–75, 1991.

[Bra92]

B.L.J. Braaksma. Multisummability of formal power series solutions to nonlinear meromorphic differential equations. Ann. Inst. Fourier de Grenoble, 42:517–540, 1992.

[Éca85]

J. Écalle. Les fonctions résurgentes I–III. Publ. Math. d'Orsay 1981 and 1985, 1985.

[Éca87]

J. Écalle. L'accélération des fonctions résurgentes (survol). Unpublished manuscript, 1987.

[Éca92]

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

[Éca93]

J. Écalle. Six lectures on transseries, analysable functions and the constructive proof of Dulac's conjecture. In D. Schlomiuk, editor, Bifurcations and periodic orbits of vector fields, pages 75–184. Kluwer, 1993.

[Har10]

G.H. Hardy. Orders of infinity. Cambridge Univ. Press, 1910.

[Har11]

G.H. Hardy. Properties of logarithmico-exponential functions. Proceedings of the London Mathematical Society, 10(2):54–90, 1911.

[Kuh00]

S. Kuhlmann. Ordered exponential fields. Fields Institute Monographs. Am. Math. Soc., 2000.

[MMvdD97]

A. Macintyre, D. Marker, and L. van den Dries. Logarithmic-exponential power series. Journal of the London Math. Soc., 56(2):417–434, 1997.

[MMvdD99]

A. Macintyre, D. Marker, and L. van den Dries. Logarithmic exponential series. Annals of Pure and Applied Logic, 1999. To appear.

[Ros83a]

M. Rosenlicht. Hardy fields. Journal of Math. Analysis and Appl., 93:297–311, 1983.

[Ros83b]

M. Rosenlicht. The rank of a Hardy field. Trans. Amer. Math. Soc., 280(2):659–671, 1983.

[Ros87]

M. Rosenlicht. Growth properties of functions in Hardy fields. Trans. Amer. Math. Soc., 299(1):261–272, 1987.

[Sch01]

M.C. Schmeling. Corps de transséries. PhD thesis, Université Paris-VII, 2001.

[Sin75]

M.F. Singer. Asymptotic behavior of solutions of differential equations and Hardy fields: Preliminary report. http://www4.ncsu.edu/~singer, 1975.

[vdD98]

L. van den Dries. Tame topology and o-minimal structures, volume 248 of London Math. Soc. Lect. Note. Cambridge university press, 1998.

[vdDMM94]

L. van den Dries, A. Macintyre, and D. Marker. The elementary theory of restricted analytic fields with exponentiation. Annals of Math., 140:183–205, 1994.

[vdH97]

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

[vdH01]

J. van der Hoeven. Complex transseries solutions to algebraic differential equations. Technical Report 2001-34, Univ. d'Orsay, 2001.

[vdH05]

J. van der Hoeven. Integral transseries. Technical Report 2005-15, Université Paris-Sud, Orsay, France, 2005.

[vdH06]

J. van der Hoeven. Transseries and real differential algebra, volume 1888 of Lecture Notes in Mathematics. Springer-Verlag, 2006.