Solutions of the equation
Abstract
We establish a novel upper bound for the real solutions of the equation specified in the title, employing a generalized derivation-division algorithm. As a consequence, we also derive a new set of Chebyshev functions adapted specifically for this problem.
Dedicated to Jorge Sotomayor
Introduction
The equation
(1) |
frequently emerges in problems concerning the qualitative theory of vector fields. In these problems, the variables , belonging to , are treated as free parameters. The goal is to estimate the number of real solutions to this equation, uniformly with respect to the parameters.
For instance, consider a smooth planar vector field whose phase portrait has a hyperbolic polycycle with saddle points and let be a smooth family of vector fields which unfolds . We fix a local transverse section (see picture below) such that and consider the first return map , which depends on the parameter . The limit cycles bifurcating from are related to the isolated fixed points of .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/e93f1825-f0d0-41b1-ac69-f37640a1f4be/x1.png)
According a result of Mourtada (see [9], Theorem 1), in the vicinity of the parameter value , there exists an appropriate domain on the transversal where the fixed-point equation has an expansion of the form (1), up to some error term which is innocuous under generic conditions. Here the parameters depend smoothly on and measure respectively the distance between the saddle separatrices and the ratio of eigenvalues at the saddle points. The additional parameter is related to the first multiplier of the Poincaré map along the unperturbed polycycle and is only relevant if the product equals one.
In such particular setting, one is interested in finding an upper bound for the cyclicity of the origin. Namely, the number of solutions (1) bifurcating from under small perturbation of the parameters.
Another situation where (1) appears is related to slow-fast systems. More specifically, it is shown in [4] that such equation appears in the study of limit cycles under bifurcation from multi-layer canard cycles. In contrast to the previous situation, here one is interested in finding a global uniform bound for the number of isolated solutions.
Independently, the structure of the group of real functions generated by translations and power maps is a subject of interest in group theory (see e.g. [2]). Here, one is interested in the abstract structure of such group and one natural question is whether is isomorphic to a free product. From a dynamical systems point of view, this problem is related to the following question, usually called center problem: Characterise those hyperbolic polycycles which lie on the boundary of a continuum of periodic orbits, i.e. those polycycles for which the Poincaré first return map is the identity. We refer the reader to [11] for some results on this direction.
In this paper, we will be interested in estimating the maximum number of real solutions of (1). To state the problem precisely, we need some definitions: Given , let denote the subset of all real numbers such that the functions defined by
(2) |
are all strictly positive. Since all these functions are strictly monotone (increasing or decreasing), it is easy to see that is an open (possibly empty) interval of . If the are all positive, is a nonempty neighborhood of infinity.
Remark 0.1.
The set is an open subset of which belongs to the o-minimal structure of sets definable in the expansion of the real field by the Pfaffian functions (in the sense of Khovanskii).
So, our goal is to estimate the number of connected components of the solution set
(3) |
Let denote the number of such connected components. It follows immediately from the analyticity of the functions in (2) that for each fixed parameter , either or is a finite set of points. As a consequence, , and we define
(4) |
The first obvious question is whether . This is a direct Corollary of Khovanskii’s Fewnomial theory [7]. Indeed, by introducing auxiliary variables , equation (1) can be equivalently written as a system,
In such form, we have a system of equations in -variables with distinct monomials, and the theory of Khovanskii provides the bound
(see [7]). In [1], Sotille and Bihan slightly improved the fewnomial’s bound. Their result implies that .
One of the difficulties of the problem is that, in general, the sequence (2) is not a Chebyshev system (see e.g. [3], Section 3.3, for the definition). To the author’s knowledge, this was first observed in Mourtada’s thesis [6]. In that work, Mourtada constructed a specific example of a generic hyperbolic polycycle with four singularities which bifurcates into five limit cycles. In our notation, this result translates to . We shall see in Section 2 that and, in Section 4, that in fact . The exact value of is unknown for .
In the next Sections, we shall describe a simple algorithm which gives a new upper bound for by a derivation-division algorithm. The algorithm is somewhat inspired in [8].
Such new upper bound coincides with the exact value of for and is significantly smaller the Khovanskii’s and Bihan-Sottile’s bound for . However, it is largely outperformed by such bounds for . We refer to Section 6 for a more detailed comparison.
One of the advantages of our method is that it provides as a by-product a system of linearly independent monomials which allows to expand the left-hand side of (1) uniformly with respect to the parameters . This expansion can be interpreted as a generalization of the compensator-type expansion introduced independently by Roussarie [12] and Ecalle [5]. We refer to Section 5 for the details.
1 Derivation/division for Laurent polynomials
We initially consider a more abstract setting of a certain ring equipped with a derivation.
Given , let denote the commutative ring of polynomials variables with integer coefficients. Our basic object will be the ring
i.e. the ring of Laurent polynomials in with coefficients in . We will frequently use the fraction notation and write, for instance, simply as . We convention that .
A monomial in is polynomial of the form , with a nonzero coefficient and where we note
The vector will be called the exponent and we say unitary if . The (poly)-degree of a such monomial is the integer vector . We note that all unitary monomials are invertible element of .
Each non-zero polynomial can be written as a finite sum of monomials
(5) |
where the set will be called the support of . We now define successively three subrings of .
(6) |
Firstly, we denote by the subring of Laurent polynomials which are homogeneous, i.e. such that all monomials in their support have a same degree. We will note by the degree of an element of . For each , we will denote by the degree of with respect to the variables (i.e. the -component of ).
We now define as the subring of homogeneous Laurent polynomials which are polynomials in the -variables. In other words, we consider the Laurent polynomials such that each monomial has an exponent satisfying . Thus, if we let and denote the variables simply as , we can expand such element in the form
(7) |
with coefficients in .
Finally, we define the smallest subring appearing in (6), whose elements will be called -regular polynomials. The definition is by induction on :
-
(i)
For , the 0-regular polynomials are simply .
-
(ii)
For , we say that a Laurent polynomial is -regular if the coefficient in the expansion (7) is -regular.
Note that, in particular, if then is a positive vector and if then is -regular. Inductively, we show that if and only if belongs to the base ring .
Let us give a different characterisation of . Consider a homogeneous Laurent polynomial of degree . Then is -regular if and only if there exists a sequence of nonzero Laurent polynomials for such that:
-
(1)
and
-
(2)
for each , and we can inductively write
where is divisible by (as a polynomial in ). In other words, contains only monomials of the form with positive exponents such that and .
This result leads to the following nested form for ,
(8) |
for some .
Example 1.1.
For , the Laurent polynomials
are regular, of respective degrees and
We now turn into a differential ring by introducing a derivation
which satisfies and acts on the variables according to the following formulas
(9) |
where we define . The following result is an obvious consequence of the above expressions for and the Leibniz rule.
Proposition 1.2.
(1) The derivation maps the ring into itself, for each index .
(2) Given , we either have or else
in other words, preserves the degree.
Proof.
By linearity, it suffices to consider the case of a monomial . Then, from the Leibniz rule, have
where and .
We now remark that each term is a Laurent homogeneous polynomial of degree . Therefore, by the additivity of the degree under multiplication, we easily obtain
Finally, if we further suppose that belongs to (i.e. ) then it follows from the above formula that is a sum of monomials lying in . The same argument works if lies in . ∎
We observe however that the ring is not preserved by , as the following simple example shows:
Example 1.3.
Consider the polynomial given in Example 1.1. Then,
which is not a Laurent regular polynomial according to our definition.
In order to deal with such phenomena, we observe that any element of can be transformed into a regular Laurent polynomial upon division by an appropriate monomial.
More precisely, the regularization monomial associated to a non-zero Laurent polynomial is a unitary monomial inductively defined as follows:
-
(i)
Case : We define .
-
(ii)
Case : If we note then we observe that we can write an expansion of in the variables as
for some (possibly negative) integers and coefficients such that are non-zero. We then define , where is the regularization monomial of .
We will denote such monomial . The following result is obvious:
Proposition 1.4.
The Laurent polynomial is regular.
Example 1.5.
Consider the polynomial given in Example 1.3. Then, the above definitions gives . The Laurent polynomial
is regular, and has degree .
Theorem 1.6.
Suppose that . Then either or the regular polynomial
is such that .
Remark 1.7.
Here we denote by the usual reverse lexicographical ordering in , namely
Proof.
If then necessarily and hence . So, let us assume that is such that and , for some . We claim that the polynomial defined in the enunciate satisfies
(10) |
Indeed, it follows from the definition of regularity and the assumption on that the expansion (8) can be rewritten as
(11) |
where and, for each index , the coefficient is an element containing only exponents in . We recall that the derivation applied to each monomial gives
where . In particular, for the innermost term in the expansion (11), we can write
where is a Laurent polynomial in containing only exponents in (and hence is divisible by ).
More generally, applying the derivation to the nested expression of given above, we obtain
where, for each index , the coefficient is a Laurent polynomial in which contains only exponents in .
Considering separately the cases and , it now suffices to observe that the division of by its regularization monomial results into a regular polynomial such that (10) holds. ∎
Applying successively the above Theorem, we obtain the following:
Corollary 1.8 (Derivation-division Algorithm).
Let . Then the sequence of Laurent polynomials in , defined inductively as
eventually gives an element .
Proof.
It suffices to observe that if then and moreover if and only if . ∎
In the above setting, we define the derivation-division complexity of as the smallest integer such that .
Remark 1.9.
The basic idea behind the derivation-division algorithm is quite simple: The expressions given in (9) show that only depends on the variables and . Hence, we inductively eliminate all monomials in and then divide out by the maximal possible factors in .
Example 1.10.
Let . Writing the variable simply as , a polynomial of degree has the form
where the coefficients lie on the ring . Applying the derivation to each monomial gives
Therefore, we can write where
and . As a consequence, is divisible by and the polynomial has degree at most . We conclude by induction that .
Example 1.11.
Let . Denoting the variables and simply by and , we consider the special case where is such that
In other words, we assume that our initial Laurent polynomial is indeed an homogeneous polynomial. Writing , we claim that in this case .
If , we fall in the situation of the Example 1.10 and the estimate holds. So, let us suppose that and write the expansion
where and is a an element of . In fact, due to our hypothesis, are all homogeneous polynomials of degree in variables .
After at most steps of the derivation-division algorithm, we obtain a new Laurent polynomial of the form
of degree . Here and we have , where each coefficient is a Laurent polynomial with support contained in the set of monomials , with . In the subsequent derivation-division step, we will get , where is the monomial
and such that, has degree . One can verify that such new polynomial satisfies the same hypothesis of (i.e. lies in ). Therefore, we conclude by induction hypothesis that . Hence, since .
2 Studying the number of isolated roots
We are now ready to estimate the the real solutions of equation (1).
For each fixed parameter value , consider the sequence of functions defined inductively as ,
and, for each ,
As we have seen in the introduction, the functions are strictly positive and analytic on some open (possibly unbounded) interval . We therefore can consider the subring of given by
(12) |
i.e. the ring formed by finite linear combinations of monomials (with possible negative exponents ) and coefficients in the ring .
The following result relates this latter ring with the ring of Laurent polynomials defined in the previous Section.
Proposition 2.1.
The derivation maps the ring into itself. Moreover, if we consider the morphism of rings defined by and
Then, we have the relation .
Proof.
The proof is immediate. Indeed, we have
Similarly,
And, in general,
Therefore, comparing with the expressions given in (9), we immediately conclude that . ∎
Consider now the analytic function given by
(13) |
which we can also write as .
An easy computation shows that belongs to . More specifically, if we consider its image under the morphism given above and the monomial
Then, is assumes the very simple form
(14) |
where we define and recall that for .
Notation: We denote by the number of steps in the derivation-division algorithm for the above polynomial .
Using the fact that , we can easily obtain the following result:
Theorem 2.2.
The number of isolated roots of on (counted with multiplicities) is bounded by .
Proof.
By Rolle’s Theorem, the number of roots of on is bounded by the number of roots of plus two. Let us assume that is not identically zero. Then, the derivation-division algorithm described in Corollary 1.8, when to the polynomial applied given by (14), defines a sequence of analytic functions on which eventually (after at most steps) ends up into a constant function. Therefore, applying again Rolle’s theorem, we conclude that the number of roots of on is bounded by . ∎
Since the number does not depend on any specific choice of parameters , we conclude that:
Corollary 2.3.
The number (see (4)) is bounded by .
3 At most three solutions for
As a simple illustration of the algorithm, let us prove that the equation
(15) |
has a maximum of three isolated solutions. If we denote by the right-hand side of the above equation, the function has the form
Therefore, upon division by we obtain
which, under the morphism , gives the regular Laurent polynomial
where we set and . Now, applying one step of the derivation-division, we obtain
Therefore, and we conclude that .
The following explicit example shows that indeed . Take , , and . Numerical computations with SageMath shows that the equation
has the three isolated solutions , and .
4 At most five solutions for
Let us prove that the equation
(16) |
has a maximum of five solutions, for all parameter values .
A similar computation to the previous example leads to the polynomial
where , and .
The following lines show the derivation-division algorithm ends in at most three steps. We omit the dependence on the coefficients to simplify the notation:
Therefore, we conclude that .
The following explicit example shows that indeed . We take
Numerical computations with 20 digits of precision in SageMath give the following five solutions
5 Disconjugacy and compensator-type expansions
We now show that, as a byproduct of the derivation-division algorithm, the function appearing in (1) can be expanded in an uniform basis of functions satisfying the Chebyshev property.
Firstly, we recall some basic definitions about disconjugate differential operators. For a thoughtful treatment, we refer the reader to the excellent Coppel’s book [3]. A -order linear differential operator defined on an interval is called disconjugate if it can be written in the form
(17) |
where, for each , is a strictly positive analytic function on . A well-known result of Polya (see e.g. [3], Chapter 3) states that is disconjugate if and only if no solution of has more than zeros in , counted with multiplicities.
Assuming that is written as in (17), we can easily construct a basis of solutions for as follows: fix an arbitrary base point and set
and, in general, define by the -iterated integral
(18) |
Notice that this collection of functions has the following properties:
-
(i)
They form a Chebyshev system, i.e. any non-trivial -linear combination has at most zeros on .
-
(ii)
For each , the subset is a basis of solutions of the truncated operator
(with the convention that )
-
(iii)
has a zero of multiplicity precisely at and .
The last two properties imply that any solution of can be expanded as
(19) |
where the coefficients are determined inductively as follows
Let us now apply these results to our original problem.
For each and each parameter value , we consider the function given by (13). Since the parameter will be innocuous in the following discussion, we suppose from now on that and will omit it to simplify the notation.
Proposition 5.1.
The function is a solution of a disconjugate linear differential operator of order .
Proof.
We recall from the proof of Theorem 2.2 and Corollary 1.8 that if we set and consider the Laurent polynomial then there exists a sequence of monomials for , such that
(20) |
where and is the derivation on defined in (9). Using the morphism defined in Proposition 2.1, this relation gives , where is the disconjugate operator of order given by
In this expression, for each , we choose to be a function in ring (given in (12)) which is mapped to under . ∎
In particular, we observe that this result allows to expand in terms of a globally defined collection of analytic Chebyshev functions. To enunciate this result precisely, we consider the open subset in the parameter space given by
Note that, as we have observed in the introduction, contains in particular the set of all parameters such that . Now, we define an open set in the total space by
Proposition 5.2 (Compensator type expansion for ).
There exists a collection of analytic functions defined on such that
-
(i)
For each fixed parameter value in , the functions forms a Chebyshev system on .
-
(ii)
We can write the expansion
where the coefficients are analytic functions on .
Proof.
The -order differential operator defined in Proposition 17 obviously depends analytically on the parameters . In other words, the functions which appear when we decompose as (17), are analytic on .
Let be the basis of solutions of defined according to (18). Note that the base point used in computation of the iterated integrals must chosen as a point on the interval . Such base point, seen as a function of the parameters , can also be chosen to be analytic.
Finally, when we write in terms of such basis of solutions , the coefficients can be computed according to the recursive procedure described in (19). This shows that they are analytic functions of . ∎
We will say that ideal generated by the functions is the ideal of coefficients of .
Remark 5.3.
Although the functions obviously depend on the choice of the basis , the ideal itself is independent of this choice. In fact one can prove that is indeed a polynomial ideal, i.e. generated by polynomials in . Let us sketch the proof assuming that : We consider the Laurent polynomial defined in the proof of Proposition 5.1. By definition, we can write a finite expansion
where are Laurent monomials and are coefficients in . Under the morphism , each monomial corresponds to an analytic function defined on some open interval containing on the boundary. Therefore, we can consider the asymptotic expansion of at . In fact, a simple application of the binomial expansion to each monomial shows that the asymptotic expansion of has the form
where is the semigroup and each is a polynomial in . Now, it is not very hard to see that is generated by .
Example 5.4.
[Classical compensators] Let and consider the function
with and . This function lies in the kernel of the third order disconjugate linear differential operator
Hence, and . Let us compute the basis of solutions by integrating from the base point . We have , and
is the well-known Ecalle-Roussarie compensator. Explicitly, we obtain
The ideal of coefficients of is generated by the polynomials , and .
Motivated by Roussarie’s result for the saddle loop case [12], we expect that the generalized compensators introduced in the last Proposition will ultimately prove useful in analyzing the first return map of non-generic hyperbolic polycycles.
6 The number of steps in the D-D algorithm
We now give an upper estimate on the number of steps of the derivation-division algorithm, when applied on a given Laurent polynomial . Such bound depends solely on the support , seen as a subset of .
More precisely, denoting by the degree of , let us suppose that
(21) |
i.e. that is a homogeneous polynomial in the variables (with no negative exponents). For each , we denote by
the collection of functions defined recursively as follows:
-
(1)
If then
-
(2)
If then
-
(2.a)
If then
-
(2.b)
If then , where .
-
(2.a)
We can now state the following result, whose proof we will omit:
Proposition 6.1.
If satisfies the condition (21) then
Suppose now that has the special form given by (14). In this situation, the support of is quite special and we conjecture, based on an algorithmic implementation, that the number is given by
Conjecture.
.
In the enunciate, denotes the Ackerman function, defined inductively as follows:
Notice that such function grows extremely fast with respect to . Here are the first 5 values:
Here, the rightmost expression is a tower of powers of height .
We conclude by providing a summary of the various upper bounds and the exact values known for :
Exact Value of | ||||
---|---|---|---|---|
1 | 2 | 8 | 3 | 2 |
2 | 3 | 5184 | 21 | 3 |
3 | 5 | 562 | 5 | |
4 | 13 | 42554 | ? | |
5 | 65533 | ? |
For , the bound becomes much larger than both Khovanskii’s bound and Bihan Sotile’s bound .
Up to the author’s knowledge, the exact value of is unknown for .
Declaration
The author states that there is no conflict of interest.
References
- [1] Bihan, F.; Sottile, F.New fewnomial upper bounds from Gale dual polynomial system.Moscow Mathematical Journal, 2007, Volume 7, Number 3, Pages 387–407.
- [2] Cohen, Stephen D.The group of translations and positive rational powers is free. Quart. J. Math. Oxford Ser. (2) 46 (1995), no. 181, 21–93.
- [3] Coppel, W. A. (1971). Disconjugacy. In Lecture Notes in Mathematics. Springer Berlin Heidelberg. https://doi.org/10.1007/bfb0058618
- [4] Dumortier F.; Roussarie R. Multi-layer canard cycles and translated power functions, Journal of Differential Equations, Volume 244, Issue 6, 2008, Pages 1329-1358.
- [5] Ecalle, J. Compensation of small denominators and ramified linearisation of local objects, Astérisque, tome 222 (1994), p. 135-199.
- [6] Jacquemard, A.; Khechichine-Mourtada F.; Mourtada, A. Algorithmes formels appliqués à l’étude de la cyclicité d’un polycycle algébrique générique à quatre sommets hyperboliques Nonlinearity, Volume 10, Number 1, pp. 19-53.
- [7] Khovanskii, A.G.Fewnomials, Trans. of Math. Monographs, 88, AMS, 1991.
- [8] Li, T.; Rojas, J.M.; Wang, X. Counting real connected components of trinomial curve intersections and -nomial hypersurfaces. Discrete Comput. Geom. 30 (2003), no. 3, 379–414.
- [9] Mourtada, A. Cyclicité finie des polycycles hyperboliques de champs de vecteurs du plan. Algorithme de finitude. Annales de l’Institut Fourier, Volume 41 (1991) no. 3, pp. 719-753.
- [10] Pauer F.; Unterkircher A. Gröbner Bases for Ideals in Laurent Polynomial Rings and their Application to Systems of Difference Equations. Applicable Algebra in Engineering, Communication and Computing 9 (1999), 271–291.
- [11] Panazzolo, D. , the exponential and some new free groups, The Quarterly Journal of Mathematics, Volume 69, Issue 1, March 2018, 75–117.
- [12] Roussarie, R. On the number of limit cycles which appear by perturbation of separatrix loop of planar vector fields. Bol. Soc. Bras. Mat 17, 67–101 (1986).
- [13] Zampieri, Sandro A solution of the Cauchy problem for multidimensional discrete linear shift-invariant systems. Linear Algebra Appl. 202 (1994), 143–162.
D. Panazzolo, IRIMAS, Université de Haute-Alsace, 68093 Mulhouse, France
E-mail address: daniel.panazzolo@uha.fr