A differential intermediate value theorem

by Joris van der Hoeven

Dépt. de Mathématiques (Bât. 425)

Université Paris-Sud

91405 Orsay Cedex

France

June 2000

Abstract

Let be the field of grid-based transseries or the field of transseries with finite logarithmic depths. In our PhD. we announced that given a differential polynomial with coefficients in and transseries with and , there exists an , such that . In this note, we will prove this theorem.

1Introduction

1.1Statement of the results

Let be a totally ordered exp-log field. In chapter 2 of [vdH97], we introduced the field of transseries in of finite logarithmic and exponential depths. In chapter 5, we then gave an (at least theoretical) algorithm to solve algebraic differential equations with coefficients in . By that time, the following theorem was already known to us (and stated in the conclusion), but due to lack of time, we had not been able to include the proof.

Theorem 1. Let be a differential polynomial with coefficients in . Given in , such that , there exists an with .

In the theorem, stands for the open interval between and . The proof that we will present in this note will be based on the differential Newton polygon method as described in chapter 5 of [vdH97]. We will freely use any results from there. We recall (and renew) some notations in section 2.

In chapter 1 of [vdH97], we also introduced the field of grid-based transseries in . In chapter 12, we have shown that our algorithm for solving algebraic differential equations preserves the grid-based property. Therefore, it is easily checked that theorem 1 also holds for . Similarly, it may be checked that the theorem holds if we take for the field of transseries of finite logarithmic depths (and possibly countable exponential depths).

1.2Proof strategy

Assume that is a differential polynomial with coefficients in , which admits a sign change on a non empty interval of transseries. The idea behind the proof of theorem 1 is very simple: using the differential Newton polygon method, we shrink the interval further and further while preserving the sign change property. Ultimately, we end up with an interval which is reduced to a point, which will then be seen to be a zero of .

However, in order to apply the above idea, we will need to allow non standard intervals in the proof. More precisely, and may generally be taken in the compactification of , as constructed in section 2.6 of [vdH97]. In this paper we will consider non standard (resp. ) of the following forms:

Here and respectively designate the infinitely small and large constants and in the compactification of . Similarly, and designate the infinitely small and large constants and in the compactification of . We may then interpret as a cut of the transline into two pieces . Notice that

Remark 2. Actually, the notations , , and so on are redundant. Indeed, does not depend on , we have whenever , etc.

Now consider a generalized interval , where and may be as above. We have to give a precise meaning to the statement that admits a sign change on . This will be the main object of sections 3 and 4. We will show there that, given a cut of the above type, the function may be prolongated by continuity into from at least one direction:

(In the cases , and so on, one has to interchange left and right continuity in the above list.) Now we understand that admits a sign change on a generalized interval if .

2List of notations

Asymptotic relations.

Logarithmic derivatives.

Natural decomposition of .

(1)

Here we use vector notation for tuples and of integers:

Decomposition of along orders.

(2)

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

We call || the weight of and

the weight of .

Additive, multiplicative and compositional conjugations or upward shifting.

Additive conjugation:

(3)

Multiplicative conjugation:

(4)

Upward shifting (compositional conjugation):

(5)

where the are generalized Stirling numbers of the first kind:

3Behaviour of near zero and infinity

3.1Behaviour of near infinity

Lemma 3. Let be a differential polynomial with coefficients in . Then has constant sign for all sufficiently large .

Proof. If , then the lemma is clear, so assume that . Using the rules

we may rewrite as an expression of the form

(6)

where and for each . Now consider the lexicographical ordering on , defined by

This ordering is total, so there exists a maximal for , such that . Now let be sufficiently large such that for all . Then

(7)

for all postive, infinitely large , since for all such .

3.2Behaviour of near zero

Lemma 4. Let be a differential polynomial with coefficients in . Then has constant sign for all sufficiently small .

Proof. If , then the lemma is clear. Assume that and rewrite as in (6). Now consider the twisted lexicographical ordering on , defined by

This ordering is total, so there exists a maximal for , such that . If is sufficiently large such that for all , then

(8)

for all postive infinitesimal .

3.3Canonical form of differential Newton polynomials

Assume that has purely exponential coefficients. In what follows, we will denote by the purely exponential differential Newton polynomial associated to a monomial , i.e.

(9)

where

(10)

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

Theorem 5. Let be a differential polynomial with purely exponential coefficients. Then there exists a polynomial and an integer , such that for all , we have .

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

(11)

by formula (5). Since , we must have . Consequently, . Hence, for some , we have . But then (11) applied on instead of yields . This shows that is independent of , for .

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

(12)

Since , it suffices to show that whenever . Now implies that . Furthermore, (5) yields

(13)

Consequently, we also have . By induction, it follows that for any iterated exponential of . We conclude that , by the lemma 3.

Remark 6. Given any differential polynomial with coefficients in , this polynomial becomes purely exponential after sufficiently many upward shiftings. After at most more upward shiftings, the purely exponential Newton polynomial stabilizes. The resulting purely exponential differential Newton polynomial, which is in , is called the differential Newton polynomial of .

4Behaviour of near constants

In the previous section, we have seen how to compute and for all . In this section, we show how to compute and for all and all transmonomials . Modulo an additive and a multiplicative conjugation with resp. , we may assume without loss of generality that and . Hence it will suffice to study the behaviour of for and positive infinitesimal (but sufficiently large) , as well as the behaviour of for positive infinitely large (but sufficiently small) .

Modulo suffiently upward shiftings (we have and ), we may assume that has purely exponential coefficients. By theorem 5 and modulo at most more upward shiftings, we may also assume that

(14)

for some polynomial and . We will denote by the multiplicity of as a root of . Finally, modulo division of by its dominant monomial (this does not alter ), we may assume without loss of generality that .

4.1Behaviour of in between constants

Lemma 7. For all with , the signs of and are independent of and given by

(15)

Proof. Since is purely exponential and , there exists an such that

(16)

for all . Let be such that , where . Then , whence

(17)

Furthermore, , whence

(18)

Put together, (17) and (18) imply that . Hence , by (16). Now

(19)

since for all positive infinitesimal .

Corollary 8. If is homogeneous of degree , then

(20)

for all with .

Corollary 9. Let be constants such that . Then there exists a constant with .

Proof. In the case when is odd, then holds for any with , by (15). Assume therefore that is even and let denote the multiplicities of as roots of . From (15) we deduce that

(21)

In other words, the signs of for and are different. Hence, there exists a root of between and which has odd multiplicity . For this root , (15) again implies that .

4.2Behaviour of before and after the constants

Lemma 10. For all with , the signs of and are independent of and given by

(22)

Proof. Since is purely exponential and , there exists an such that

(23)

since . Furthermore and , whence . In particular, , so that , by (23). Now

(24)

since for positive infinitely large .

Corollary 11. If is homogeneous of degree , then

(25)

for all with .

Corollary 12. Let be a constant such that . Then there exists a constant with .

Proof. In the case when is odd, then holds for any with , by (15). Assume therefore that is even and let be the multiplicity of as a root of . From (15) and (22) we deduce that

(26)

In other words, the signs of for and are different. Hence, there exists a root of which has odd multiplicity . For this root , (15) implies that .

5Proof of the intermediate value theorem

It is convenient to prove the following generalizations of theorem 1.

Theorem 13. Let and be a transseries resp. a transmonomial in . Assume that changes sign on an open interval of one of the following forms:

  1. , for some with .

  2. .

  3. .

  4. .

Then changes sign at some .

Theorem 14. Let and be a transseries resp. a transmonomial in . Assume that changes sign on an open interval of one of the following forms:

  1. , for some with .

  2. .

  3. .

  4. .

Then changes sign on for some with .

Proof. Let us first show that cases a, b and d may all be reduced to case c. We will show this in the case of theorem 13; the proof is similar in the case of theorem 14. Let us first show that case a may be reduced to cases b, c and d. Indeed, if changes sign on , then changes sign on , or . In the second case, modulo a multiplicative conjugation and upward shifting, corollary 9 implies that there exists a such that admits a sign change on . Similarly, case d may be reduced to cases b and c by splitting the interval in two parts. Finally, cases b and c are symmetric when replacing by .

Without loss of generality we may assume that , modulo an additive conjugation of by . We prove the theorem by a triple induction over the order of , the Newton degree of the asymptotic algebraic differential equation

(27)

and the maximal length of a sequence of privileged refinements of Newton degree (we have , by proposition 5.12 in [vdH97]).

Let us show that, modulo upward shiftings, we may assume without loss of generality that and are purely exponential and that . In the case of theorem 13, we indeed have and . In the case of theorem 14, we also have . Furthermore, if is such that changes sign on , then is such that changes sign on .

Case 1: (27) is quasi-linear. Let be the potential dominant monomial relative to (27). We may assume without loss of generality that , modulo a multiplicative conjugation with . Since By , we have or for certain constants .

In the case when , there exists a solution to (27) with . Now and . We claim that and must be equal. Otherwise would admit a solution between and , by the induction hypothesis. But then the potential dominant monomial relative to (27) should have been , if is the largest such solution. Our claim implies that , so that . Finally, lemma 4 implies that admits a sign-change at . Lemma 7 also shows that .

In the case when , then any constant is a root of . Hence, for each , there exists a solution to (27) with . Again by lemmas 4 and 7, it follows that admits a sign change at and on .

Case 2: . Let be the largest classical potential dominant monomial relative to (27). Since (resp. ), one of the following always holds:

Case 2a

We have (resp. ).

Case 2b

We have .

Case 2c

We have .

For the proof of theorem 14, we also assume that in the above three cases and distinguish a last case 2d in which .

Case 2a. We are directly done by the induction hypothesis, since the equation

(28)

has a strictly smaller Newton degree than (27).

Case 2b. Modulo multiplicative conjugation with , we may assume without loss of generality that . By corollary 12, there exists a such that . Actually, for any transseries we then have . Take such that

(29)

is a privileged refinement of (27). Then either the Newton degree of (29) is strictly less than , or the longest chain of refinements of (29) of Newton degree is strictly less than . We conclude by the induction hypothesis.

Case 2c. Since is the largest classical dominant monomial relative to (27), the degree of the Newton polynomial associated to any monomial between and must be . Consequently,

(30)

By the induction hypothesis, there exists a monomial with and

(31)

In other words, is a dominant monomial, such that and

(32)

We conclude by the same argument as in case 2b, where we let play the role of .

Case 2d. Since is the largest classical dominant monomial relative to (27), the degree of the Newton polynomial associated to any monomial between and must be . Consequently,

(33)

By the induction hypothesis, there exists a monomial with and

(34)

In other words, is a dominant monomial, such that and

(35)

We again conclude by the same argument as in case 2b.

Corollary 15. Any differential polynomial of odd degree and with coefficients in admits a root in .

Proof. Let be a polynomial of odd degree with coefficients in . Then formula (7) shows that for sufficiently large we have , since is odd in this formula. We now apply the intermediate value theorem between and .

Bibliography

[vdH97]

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