Vanishing Abelian Integrals on Zero-Dimensional Cycles
Abstract.
In this paper we study conditions for the vanishing of Abelian integrals on families of zero-dimensional cycles. That is, for any rational function , characterize all rational functions and zero-sum integers such that the function vanishes identically. Here are continuously depending roots of . We introduce a notion of (un)balanced cycles. Our main result is an inductive solution of the problem of vanishing of Abelian integrals when are polynomials on a family of zero-dimensional cycles under the assumption that the family of cycles we consider is unbalanced as well as all the cycles encountered in the inductive process. We also solve the problem on some balanced cycles.
The main motivation for our study is the problem of vanishing of Abelian integrals on single families of one-dimensional cycles. We show that our problem and our main result are sufficiently rich to include some related problems, as hyper-elliptic integrals on one-cycles, some applications to slow-fast planar systems, and the polynomial (and trigonometric) moment problem for Abel equation. This last problem was recently solved by Pakovich and Muzychuk ([28] and [27]). Our approach is largely inspired by their work, thought we provide examples of vanishing Abelian integrals on zero-cycles which are not given as a sum of composition terms contrary to the situation in the solution of the polynomial moment problem.
Key words and phrases:
Abelian integral, persistent center, tangential center, moment problem, center-focus problem, canard center, slow-fast system2010 Mathematics Subject Classification:
34C07, 34C08, 34D15, 34M351. Introduction
Zero-dimensional Abelian integrals were introduced by Gavrilov and Movasati in [18] and the problem of the bound for the number of their zeros was studied. In this paper we study conditions for their identical vanishing on families of zero-dimensional cycles.
Definition.
Let be a rational function of degree . Let denote an -tuple of analytic preimages , where and is the set of critical values of . Define a zero-dimensional cycle (shorter cycle) of as the sum
(1.1) |
Zero-dimensional cycles form the reduced -th homology group denoted . We say that a cycle is simple if it is of the form , and trivial if for every .
Let be a rational function. As in [18], define zero-dimensional Abelian integrals of along the cycle by
(1.2) |
Note that Abelian integrals on zero-cycles are simply algebraic functions, contrary to the one-dimensional case.
We study two problems:
Problem 1.1.
Characterize rational functions , such that for any family of cycles of the Abelian integral vanishes identically.
Problem 1.2.
(Tangential center problem) Characterize rational functions , and cycles of such that the Abelian integral vanishes identically.
In Section 2 we state our results on those two problems. Concretely, we solve completely Problem 1.1 and give an inductive solution of Problem 1.2 under the assumptions that are polynomials, and that the family of cycles we consider is unbalanced as well as all the cycles encountered in the inductive process. The definition of balanced and unbalanced cycles is related to the behavior of the branches as tends to infinity, and it is made precise in Section 2. We also solve the problem on some balanced cycles. In particular we provide examples of vanishing of Abelian integrals on zero cycles which are not given as a sum of composition terms contrary to the situation in the polynomial moment problem as solved by Pakovich and Muzychuk. The proofs of these results are included in Sections 5, 6 and 7.
In Section 3 we show that problems above are the zero-dimensional version of the analogous problems on the vanishing of Abelian integrals on one-cycles which appear in the study of periodic solutions of deformations of Hamiltonian vector fields in the plane.
We also push further the analogy between deformations of planar integrable systems and deformations of polynomials in one variable. In particular we define the displacement function along a zero-cycle, zero-dimensional Abelian integrals being its principal part. We study identical vanishing of the displacement function.
Problem 1.1 for one-dimensional Abelian integrals was studied by Gavrilov [17] and completely solved by Bonnet-Dimca [4] in the polynomial case and by Muciño in [22] in the generic rational case. Problem 1.2 for one-dimensional generic Abelian integrals was solved by Ilyashenko [19]. It was solved in the hyper-elliptic case (which is non-generic) for vanishing cycles by Christopher and the third author [12]. The general one-dimensional case remains open for both problems.
In Sections 9, 10 we give applications of our results to study the one-dimensional tangential center problem and related problems. Concretely, we study the vanishing of hyper-elliptic integrals on one-cycles and some related problems for slow-fast planar systems.
In Section 8 we prove that the polynomial and the trigonometric moment problem can be stated as the vanishing of zero-dimensional Abelian integrals on special cycles. In particular, in the polynomial case the cycle is unbalanced and so are all the cycles in the inductive process so our results can be applied. The moment problem was started by Briskin, Françoise and Yomdim ([5, 6]) and it has received a lot of attention (see e.g. [7, 11, 23, 26]), but was only recently solved by Pakovich and Muzychuk ([28] and [27]). Our approach is largely inspired by their work.
In [5] Briskin, Françoise, Yomdin formulated the composition conjecture for the moment problem (see Section 8). Analogous conjecture in our context would be that for a cycle of the Abelian integral vanishes if and only if there exist decompositions and such that the projected cycle is trivial. A simple counter-example was given by Pakovich [27].
This allows the formulation of a weaker composition conjecture: The integral vanishes if and only if there exist decompositions , such that the projected cycles , are trivial. In [28] the authors solved the moment problem in the polynomial case. Their result can be interpreted as saying that the weak composition conjecture is valid for special cycles appearing in the polynomial moment problem. See Section 8 for more details. We show in Example 7.5 that even the weak composition conjecture is not true for general cycles in the polynomial case.
2. Main Results
We give a complete solution of Problem 1.1 in Theorem 2.1 and a partial solution of Problem 1.2 in a generic polynomial case in Theorem 2.2 and in some exceptional cases in Section 7. Our first result is the following:
Theorem 2.1.
Given , the following conditions are equivalent:
-
(1)
for every (simple) cycle .
-
(2)
There exists , such that .
In order to formulate the main theorem, Theorem 2.2, we must define the notions of unbalanced, totally unbalanced and projected cycles. Let be a polynomial of degree and its set of critical values. Then defines a fibration, with fiber consisting of points. Take a base point . Then this fibration defines a mapping from the first homotopy group to the group of automorphisms . Its image is called the monodromy group . It acts transitively on the fiber (see, e.g., [33]). Changing the base point conjugates all elements of the group, so the abstract group does not depend on the choice of the base point . Consider the loop in winding anti-clockwise around all critical values in . It corresponds to a permutation cycle of of order . We label the roots , of so that this permutation shifts the indices of the roots by one. The choice of the first root is arbitrary.
Let denote the conjugacy class of , that is, the set of all , for any . Note that is the set of permutations induced by all paths winding counter-clockwise once around infinity.
Definition.
We say that a cycle of is balanced if
where is any primitive -th root of unity. If is not balanced, we say that is unbalanced.
For each such that and each cycle of we define a cycle of called the projected cycle by
(2.3) |
Here are all the different roots of .
We say that is totally unbalanced when for every such that , the projected cycle is unbalanced or trivial.
The notion of balanced cycle is related to the behaviour of the branches when tends to infinity. It is well defined, i.e., independent on the way how permutation cycles are written. Indeed, if we write , then
Thus, if and only if
Note that in particular totally unbalanced cycles are unbalanced as seen from the trivial decomposition .
Theorem 2.2.
Let .
-
(1)
If is a totally unbalanced cycle, then
(2.4) if and only if there exist such that , and the projected cycles are trivial for every .
-
(2)
If is an unbalanced cycle, then is a solution of (2.4) if and only if there exist such that , , and for every
and the projected cycle is trivial or balanced.
In Example 7.5, we show that claim (1) in the Theorem is not true for general cycles contrary to the situation in polynomial moment problem. In order to completely solve Problem 1.2 for Abelian integrals, it remains to solve it for balanced cycles. Balanced cycles are exceptional cycles of codimension at least one. Moreover, in Section 7 we show that if is prime, then there are no balanced cycles. We also completely solve the problem for .
In addition to the problem of persistence of centers, another motivation for our study is the relation of Problem 1.2 to the moment problem. The polynomial moment problem consists for a given polynomial in searching for all polynomials such that
In Section 8 we show that the moment problem is equivalent to solving Problem 1.2 for a concrete cycle . We show that this cycle is always totally unbalanced, so the first part of Theorem 2.2 extends Pakovich and Muzychuk’s theorem [28] solving the polynomial moment problem. In [24] Pakovich gives an explicit solution of the polynomial moment problem. In particular he shows that the function can be written as a sum of at most three reducible solutions (a polynomial is a reducible solution if there exists such that , and ). In Example 6.5 we show that the analogous statement for the tangential center problem is not even true for general totally unbalanced cycles.
3. Displacement function along a family of zero-cycles
Zero-dimensional Abelian integrals , are defined for a couple and of polynomials in one variable.
The definition is analogous to the definition of one-dimensional Abelian integrals , , defined for a polynomial and a polynomial differential -form in two variables. In this section we push further the analogy between systems depending on one variable and planar systems. Assume that some point is a strict local extremum of a real polynomial function . Then the system has a center at , parametrized by a family of one-cycles . Consider the deformation
(3.5) |
For most deformations , the center will be broken. A natural question is for what deformations , the center is preserved? This is the infinitesimal center problem.
The problem is tackled by considering the displacement function on a transversal to the family of cycles parametrized by . The displacement function measures the displacement along a trajectory of the deformed system (3.5) from the starting point on the transversal to the first return point to the transversal. A center is preserved if . A classical result is that Abelian integrals give the first order term of the displacement function:
(3.6) |
Of course the vanishing of the first order (Abelian integral) term is a necessary, but not sufficient condition for the persistence of a center. It is named the tangential (or first order) center problem.
The above problem can be complexified. One considers the foliation defined by (3.5). Taking a transversal and lifting each loop to nearby leaves the holonomy associated to is defined. By assumption, for all holonomies are identity (i.e., the associated displacement functions are zero-functions).
Consider now the problem in one-dimensional space. Let be a function and a family of cycles of . Consider a perturbation of by
and the deformed cycle , where , for , where is the set of critical values of . We define the displacement function of along the cycle by
As in the case of one-dimensional loops, zero-dimensional Abelian integrals correspond to the principal part of the displacement function of the perturbation along a family of zero-dimensional cycles. Indeed, from , it follows that , giving
(3.7) |
Problems 1.1 and 1.2 can be seen as first order (called also tangential) problems of analogous problems for the displacement function along families of zero-cycles:
Problem 3.1.
Characterize rational functions , such that for any family of cycles of the displacement function of , along any zero cycle vanishes identically.
Problem 3.2.
(Infinitesimal center problem) Characterize rational functions , and cycles of such that the displacement function of , along the family of zero-cycles vanishes identically.
Corollary 3.3.
Given , the following conditions are equivalent:
-
(1)
for every (simple) cycle .
-
(2)
There exists , such that .
Problem 3.2 would probably require studying vanishing of some iterated integrals on zero-cycles. We hope to address this problem in the future.
4. Imprimitivity systems of the monodromy group and equivalence classes of decompositions
The aim of this section is to recall some definitions and results about the monodromy group and to study the relationship between the decompositions of a rational function and its monodromy group.
Given a rational function of degree , let be the set of critical values of . The monodromy group is defined as in the polynomial case and acts transitively on generic fibers .
As shown in [15] the monodromy group is the Galois group of the Galois extension of by the preimages of by , that is,
Let and let be a transitive permutation group on . A subset is called a block ([32]) of if for each the image set and are either equal or disjoint. Given a block , the set forms a partition of into disjoint blocks of the same cardinality which is called an imprimitivity system (or a complete block system). Each permutation group has two trivial imprimitivity systems: and .
Given we say that two decompositions of , and , are equivalent if and define the same field over every preimage of by , that is, if for every .
Proposition 4.1.
Let . There exists a one-to-one correspondence between imprimitivity systems of and equivalence classes of decompositions of .
Moreover, if , is the imprimitivity system corresponding to and , then .
Remark 4.2.
Let us point out that the second statement in the proposition is in particular telling us that given a cycle of , the projected cycle of given by the expression (2.3) can be written as
where is the block containing of the imprimitivity system corresponding to the decomposition .
Proof.
Let be a rational function of degree , and let us denote by the preimages of as above.
Let be a decomposition and let us find the associated imprimitivity system . To do so, we only have to find a block , since by definition, is an imprimitivity system of . For any preimage of , we have that
and therefore
By the Fundamental Theorem of Galois theory (in Artin’s version, see [2, 31]) there exists a subgroup of the monodromy group containing , the stabilizer of , such that
As a consequence, for every . The orbit of by the action of this subgroup is the block we are looking for:
It is obvious that for every , and it is also immediate to see that when . Then, we also have proved that is constant on . Conversely, if , then : since the monodromy group is transitive, then there exists some such that , and therefore
which implies that is invariant by . Then must belong to and belongs to . For another different block , a similar argument proves the same statement. Then, the blocks of the imprimitivity system verify that if and only if belongs to the same block of , proving that the imprimitivity system we get does not depend on the choice of .
Now let be an imprimitivity system for . The block containing the element provides a subgroup of containing the stabilizer of . Precisely,
By the Fundamental Theorem, the following inclusions of groups
yield the following inclusions of fields
By Lüroth’s Theorem [31] applied to , for some rational function . Since , there exists another rational function such that . Moreover, for any .
If we choose an element of a different block , then we get another subgroup of , which contains the stabilizer of . Again by the Fundamental Theorem, it provides a field , which by Lüroth’s Theorem is generated by some rational function , that is, . Therefore, there exists another rational function such that . Let us see how different these functions and are. Since is transitive, there exists a permutation such that . Therefore, (or ), and it is immediate to check that
(4.8) |
Since by the Fundamental Theorem we know that and are the Galois groups of over and respectively, let us see what equality (4.8) gives: an element is an algebra automorphism of such that . Since for some , that means
which implies that . Also, as leaves invariant , it holds that
which implies that . To sum up, equality (4.8) implies that is not only the Galois group of over , but also over . Similarly, is not only the Galois group of over , but also over . By the Fundamental Theorem, we obtain that and . Therefore, given an imprimitivity system , we get not only one decomposition of , but a whole equivalence class of decompositions. ∎
5. Zero-dimensional Abelian integrals and displacement functions vanishing on any cycle
The aim of this section is to solve Problems 1.1 and 3.1 proving Theorem 2.1 and Corollary 3.3. First we give a simple sufficient condition for the vanishing of the displacement function (and hence also the corresponding Abelian integral) of the family , , along a family of cycles of .
Proposition 5.1.
Assume that , , where . Let be a cycle of and its continuation as a cycle of . Let
and
be the cycle of and obtained by projection of and by respectively. Then,
-
(1)
and .
-
(2)
In particular, if the projected cycle is trivial, then is trivial,
and
Remark 5.2.
In other words, the assumption (2) of the proposition can be reformulated by saying that the cycle projects by to a trivial cycle
of and, moreover, the function factors through the same , so a change of coordinates leading to integration on the trivial cycle is possible.
Proof.
(1) Let be a cycle and assume that , . Set . Then
The second claim of (1) follows the same way.
(2) If is a trivial cycle, then all its coefficients are equal to zero. The dependence with respect to being continuous, it follows that the cycle is trivial too. Now the claim follows from (1). ∎
Next result shows that in case of simple cycles condition (2) of Proposition 5.1 is necessary and sufficient for the vanishing of Abelian integrals.
Proposition 5.3.
Let , and consider a simple cycle . Then
if and only if there exist such that , and .
Remark 5.4.
Let us observe that since , by Proposition 4.1 the set is included in some block of the imprimitivity system corresponding to the decomposition .
Proof.
The sufficient condition is a consequence of Proposition 5.1.
For the necessary condition, let us define
Since
by Lüroth’s Theorem for some rational function . Since , then , for some . ∎
Now, we are ready to prove Theorem 2.1.
Proof of Theorem 2.1 and Corollary 3.3.
By Proposition 5.1, in Theorem 2.1 implies , similarly in Corollary 3.3 implies . Therefore, to conclude it is sufficient to prove that in Theorem 2.1 implies .
Let be such that for every (simple) cycle . First, let us observe that the condition for every (simple) cycle is equivalent to
Therefore, is invariant under the action of the whole monodromy group , that is, , which means that there exists such that . Then . ∎
Under a generic condition on we show that the existence of a persistent center or a tangential center for any non-trivial cycle is equivalent to the integrability of the system as characterized in Theorem 2.1. Recall that a point is of Morse type if , . Generically all critical points are of Morse type and all critical values are different.
Proposition 5.5.
Given , assume that every critical value corresponds to only one critical point, and moreover that they are all of Morse type. For any non-trivial cycle the following conditions are equivalent:
-
(1)
.
-
(2)
along .
-
(3)
There exists such that .
Proof.
We prove that implies for a given cycle , since, by Theorem 2.1, implies and , and trivially implies .
The permutation in the monodromy group associated with a critical value with a Morse point is a transposition. Let the cycle be given by (1.1).
First, we shall prove that there is a transposition corresponding to a critical value and such that . Note that otherwise, for every and every critical value . Since the monodromy group is transitive and it is generated by the permutations corresponding to critical values, we would have for every , and certain . Now, since is a cycle, and the cycle would be trivial.
Take and such that . By analytic continuation,
Therefore,
so the integral vanishes. By Proposition 5.3 and Remark 5.4, there exists such that , and .
If and is a non-trivial imprimitivity system of , then for some block . Since is generated by the transpositions corresponding to the critical values and it is transitive, then for each , there is a chain of transpositions , , , . Therefore, there is only one block , and the unique imprimitivity systems of are the trivial ones. Thus, must be the trivial imprimitivity system with a unique block and, in consequence, . ∎
If the polynomial is not generic, then there are some solutions which are not of the form . Moreover, there exists solutions that are not of the form with .
Example 5.6.
Let . Let , , be the real roots of . Consider the cycle
Here can be analytically extended to . Then for any odd function it follows , because is a biquadratic equation and therefore , . Take, for instance, .
Let us show that there not exist with such that , . Indeed, assume that there exist such decompositions and let be the associated imprimitivity system, and . Since the imprimitivity system is not trivial, then and . On the other hand, does not admit decompositions, so , but for every .
6. Tangential center problem for polynomials
In this section we focus on the case when are polynomials. We shall prove Theorem 2.2 and some other results under the assumption that the cycle given by (1.1) is totally unbalanced.
Consider a cycle given by (1.1). It is represented by the vector . Let be an element of the monodromy group of . It transforms the cycle into the cycle
(6.9) |
We consider the action of on defined by
Replacing by , we see that the orbit of the cycle by the action of the monodromy group is represented by the space defined by
Let us choose and number the branches in such a way that .
Let be an imprimitivity system of . By Proposition 4.1 it corresponds to a class of decompositions , where can be chosen polynomials (see Lemma 3.5 of [12]). Let . Then and . Since , then each , consists of the congruence class of modulo in . For each decomposition as above, the corresponding projected cycle is of the form
where are all the different values of .
Let denote the set of divisors of such that there is a decomposition with . Note that in we have the partial order induced by divisibility. We say that is a is a complete set of elements of covered by if and are maximal among divisors of in .
For every , let us denote by the set of -periodic vectors.
Lemma 6.1 ([28]).
Each -irreducible subspace of has the form
where and is a complete set of divisors of covered by . The subspaces are mutually orthogonal and every -invariant subspace of is a direct sum of some as above.
Lemma 6.2 ([28]).
Let be a polynomial and let
denote the Puiseaux expansion at infinity of , where is one of the branches of the preimage of the polynomial . For any , , if we define
there exist such that
Moreover, there exist , with , such that
Now we are in conditions of proving Theorem 2.2.
Proof of Theorem 2.2.
First we prove .
Assume that there exist such that , , . Set and let denote the projected cycle
Assume that for every , either the projected cycle is trivial or balanced, and
Then by Proposition 5.1, is a solution of (2.4) for every . By linearity, is also a solution of (2.4).
Now, assume that is a solution of (2.4). By analytic continuation this means that
(6.10) |
Assume that is unbalanced. Thus, there exists a permutation cycle of order , , such that
There is no loss of generality in assuming that . Consider the orbit of the cycle
Let us define the vectors
for every natural , where . Note that these vectors for form an orthogonal basis of . Moreover, for every , the (complexified) subspace is generated by the vectors such that .
Let us prove that . Since are invariant by the action of and is irreducible, then either or . We show that and , thus . Indeed, for every , is generated by the vectors where . Since , then and these are orthogonal to . Therefore , but since is unbalanced, is not orthogonal to . Then and are not orthogonal, thus showing that .
Now, take the Puiseux expansions of at infinity. Applying Newton’s Polygon Method to (see, e.g., [10, p. 45]) can be written as an analytic function in , and composing with we get
Now, (6.10) yields
(6.11) |
Assume that for some . Then is orthogonal to , and therefore, to . Since is generated by the vectors , , maximal in , this implies that is a linear combination of these vectors. But are linearly independent, therefore coincides with one of them. Consequently, for some maximal element of .
The Puiseux expansion of at infinity is
Therefore, satisfies (6.11) and as a consequence
Now, since , projecting the cycle by ,
where , and is the number of different .
Since has for every , by induction on the number of maximal elements of such that for some , we obtain
where the Abelian integral of along the cycle
vanishes for every . It only remains to prove that we can assume that either is trivial or balanced. If it is none of them, then it is unbalanced, and applying again the arguments above, , where now the cycles associated with each are projections of the cycle . Repeating this argument recursively, one obtains that either is trivial or balanced.
In particular, if is totally unbalanced, then every must be trivial, which proves . ∎
In [24] Pakovich proves in particular that any solution of the polynomial moment problem can be expressed as a sum of at most three reducible solutions. We show that in our case of general cycles there exist solutions with arbitrarily many terms which cannot be reduced.
Let be a polynomial of degree such that has different branches for every . Let be a cycle given by (1.1), which can be seen as a vector of the -vector space . Note that given a vector of , we can always get a multiple with integer coordinates.
As before, let us choose and choose the numbering of the branches such that . Then, for every positive integer the candidates to imprimitivity systems of are
where
for . By Proposition 4.1, if is an imprimitivity system, we get a decomposition (up to an equivalence) . Then, by Remark 4.2 the projected cycle is trivial if and only if for . Similarly, we shall say that for is the projection of the cycle by even when is not an imprimitivity system (note that if , then is always an imprimitivity system). These equations form a homogeneous linear system in unknowns . As every column in the associated matrix has just one and the rest of the coefficients zero, while every row has at least one , we can assure that the rank of the system is . In particular, a non-trivial (integer) solution of the system is a cycle, since the sum of all equations is the condition for being a cycle.
Let us check the dimension of the -vector space of the cycles with trivial projection for every .
Proposition 6.3.
Under the above conditions, the dimension of the -vector space of the cycles with trivial projection for every is , where is Euler’s totient function, defined as the number of positive integers less than that are relatively prime with .
Remark 6.4.
Let us observe that these cycles with trivial projections are not necessarily totally unbalanced. To make sure that they are we must ask for another condition to hold: that they are not balanced. Then the set of totally unbalanced cycles with trivial projections is obtained by substracting some codimension-one -vector subspaces from the -vector space of cycles with trivial projections.
Proof.
First, let us point out that it is enough to check those cycles with trivial projection for every , where is a prime divisor of .
Let be the decomposition of as product of primes . For every prime we get a homogeneous linear system of linearly independent equations in unknowns. Every one of those equations gives a -periodic vector of :
So for every prime we have linearly independent vectors of , that is, we get a basis for .
The subspace of is the solution of the homogeneous linear system of equations we get for the primes . So, the codimension of is the rank of the matrix associated with this system of equations.
If we consider the equations associated with primes and , we have a set of vectors which consists of a basis for and a basis for . Then, we get a system of generators for and, in consequence, the rank of the matrix associated with those equations is the dimension of . That is,
Now let us observe that , since because and are relatively prime. Therefore,
In the general case, when we consider all the equations we get:
Therefore, is a -vector subspace of of dimension . Consequently, for every we have (up to a multiple) linearly independent cycles with trivial projections. ∎
Example 6.5.
Given , for the polynomial we can find at least linearly independent totally unbalanced cycles such that is a solution of (2.4) but it cannot be reduced to less than a sum of four reduced solutions. Then, in this case the results of Pakovich [24] about the minimum number of reducible solutions of which the solution of the moment problem is a sum do not apply.
In general, given , for the polynomial we can find at least linearly independent totally unbalanced cycles such that is a solution of (2.4), for arbitrarily large.
7. Balanced Cycles
We recall that in order to obtain a complete solution of the tangential zero-dimensional center problem it only remains to solve the problem for balanced cycles. We shall give some examples showing that in this case there exist solutions that are not sums of reduced solutions.
First we recall some definitions and give a characterization of balanced cycles.
Given a field extension and an algebraic element over , the minimal polynomial of is the monic polynomial of the least degree such that . The minimal polynomial is irreducible over and any other non-zero polynomial with is a multiple of in .
The -th roots of unity, that is, the roots of the polynomial form an abelian cyclic group denoted by . If the characteristic of is zero, then has order and
where is called a primitive -th root of unity because it generates . When , then there are exactly primitive -th roots of unity.
The -th cyclotomic polynomial is defined to be the minimal polynomial that has the primitive -th roots of unity as simple roots:
Obviously, is the degree of . Cyclotomic polynomials are irreducible polynomials with integer coefficients and is the minimal polynomial of over .
Given a divisor of , the primitive -th roots of unity are the elements of of order . As the order of any element of is a divisor of , we can conclude that
Then
Given a cycle by (1.1) and a permutation cycle , we associate the principal ideal
The definition above does not depend on the way we represent . Indeed, if we write , then
and consequently
where is invertible in .
We will define to be any generator of , that is,
(7.12) |
Note that the cycle is balanced if and only if is a root of every polynomial of for every , which is equivalent to being a root of for every .
Proposition 7.1.
The cycle is balanced if and only if
where is the -th cyclotomic polynomial and .
Proof.
Let be a cycle, and be the polynomial associated with and by (7.12). Since is a cycle, it follows that , that is, is a root of , thus giving that is a divisor of .
Now the cycle is balanced if and only if for every , that is, if and only if is a root of the polynomial . As is the minimal polynomial of , it is equivalent to dividing . The claim follows as and are relatively prime. ∎
Corollary 7.2.
If is prime, then every cycle is totally unbalanced.
Proof.
First, if is prime, then . So it is not a factor of , and the cycle is unbalanced. Moreover, since is prime, then the unique imprimitivity systems are the trivial ones, and in consequence, is totally unbalanced. ∎
Now we give some examples of solutions when the cycle is balanced. We will extend them to examples where the cycle is unbalanced but not totally unbalanced with respect to .
The first example is , which can be completely solved. In this case does not depend on , so to simplify we write .
Proposition 7.3.
Proof.
The -th roots of are for , where is an -th root of and is a primitive -th root of unity.
Let us compute the value of the expression :
Hence, if and only if whenever , . For and , because is a cycle and because it is balanced respectively. Then, we can suppose . For , is a primitive -th root of unity. As we saw in Proposition 7.1, is a root of if and only if divides . Now is equivalent to the condition that the -th cyclotomic polynomial does not divide , where . ∎
Example 7.4.
Let us study some degree four examples. First note that since has degree three, by Proposition 7.1, , . Then, the only candidates for balanced cycles are
Take . Then, by Proposition 7.3, is a solution if and only if it is of the form
with polynomials. Indeed, the only condition corresponds to coefficients , with . Then , does not divide and hence there are no terms of powers in .
Example 7.5.
Now let be the fourth Chebyshev polynomial and again assume that is a balanced cycle, thus,
By direct computation,
Rewrite , for certain . Then the Abelian integral along the cycle reduces to
Therefore is a solution if and only if there exist polynomials such that
Observe that the weak composition conjecture does not hold in this case: take, for instance, the solution , which does not share any non-trivial compositional factor with .
Example 7.6.
Consider now and a balanced cycle . Note that this implies that the associated polynomial must have the factors and . By Proposition 7.3, the solutions depend on which cyclotomic polynomials divide . The unique candidates are and . Moreover, it is not possible that both , divide , since . Therefore there are three possibilities: neither nor divide (e.g., ), divide (e.g., ), or divide (e.g., ).
Now, Proposition 7.3 gives us all the solutions:
-
(1)
For the cycle
satisfies if and only if for every .
-
(2)
For the cycle
satisfies if and only if for every .
-
(3)
For the cycle
satisfies if and only if for every .
Example 7.7.
To end with this section, let us show an example of an unbalanced cycle which is not totally unbalanced.
Take , with . Let be the element of the monodromy group corresponding to a simple clockwise loop around all critical values. We shall number the branches in order to have .
Consider the cycle . It is easy to check that this cycle is unbalanced, and that the non-trivial decompositions of are , , and .
Let us denote
Then is trivial, is unbalanced, and is balanced.
8. Moment Problem
Let us consider the following Abel equation
(8.13) |
where , are analytic functions. We say that a solution of (8.13) is closed if . Let us denote
Assume that we have a center at the origin for (thus, every bounded solution of the system is closed), which happens iff , and we want to study when the center persists after perturbation (i.e., for every we have a center). This problem comes from the center-focus problem for certain planar systems, which after a change to polar coordinates become of the form of (8.13) with trigonometric polynomials. The case when has been extensively studied and solved. We show in this section how both cases can be seen as zero-dimensional tangential center problems.
To first order in , the center is persistent if and only if for every close to zero, where is the solution of (8.13) determined by .
Differentiating with respect to and evaluating at , , one has that for every close to zero if and only if
(8.14) |
for every close to zero (see [6]).
The moment problem [5] asks for a given , to find all such that (8.14) is satisfied. It has been extensively studied in recent years (see, e.g., [7], [16] and references therein).
The first case to be studied is the case when and are polynomials, recently solved by Pakovich and Muzychuk [28]. More precisely, they prove:
Theorem 8.1 ([28]).
Take a polynomial such that . A polynomial is a solution of the polynomial moment problem if and only if there exist polynomials such that , for every and
We shall show that the polynomial moment problem is equivalent to a zero-dimensional tangential center problem for and defined as in Theorem 8.1, and for the cycle
where are all the solutions of close to for close to , and analogously for and . We shall prove that is totally unbalanced, so Theorem 8.1 follows from Theorem 2.2. Note however that in the proof of Theorem 2.2 we use methods developed in Theorem 8.1.
Proposition 8.2.
Given polynomials such that , the polynomial is a solution of (8.14) if and only if
Moreover, is totally unbalanced.
Proof.
By Theorem 1.1 of [28], is a solution of the polynomial moment problem if and only if there exists polynomials such that , for every and
On the other hand, if is totally unbalanced, then Theorem 2.2 implies that is a solution of (8.14) for the cycle if and only if there exist polynomials such that for every ,
and is trivial.
First, let us prove that is equivalent to being trivial. To this end, let be the imprimitivity system corresponding to some such that .
If , then the blocks of contain elements of both and : take a non-critical value of and joint it to by an arc. Then maps this arc into some arcs close to and others close to . Let denote an element of the monodromy group corresponding to a cycle around . Reorder so that permutes cyclically and and that belongs to the same block of . Now, the blocks reduced to , consist of the congruence classes modulo . Therefore, for every block ,
Conversely, if , since the blocks contain branches where the value of is the same, then every block of the imprimitivity system corresponding to either contains elements of or elements of , and then , and is not trivial.
Therefore, to conclude it only remains to prove that is totally unbalanced. Indeed, let us choose a permutation corresponding to a simple loop around infinity and reorder the numbering of the branches such that . Define
The Monodromy Lemma in [26] states that the convex hulls of the sets and are disjointed when , thus, can be divided into two half-planes such that and .
Now, the center of mass of is the origin (i.e., ) if and only if the center of mass of is equal to the center of mass of . But the center of mass of belongs to and the center of mass of belongs to . Therefore, is unbalanced.
If , then there are two possibilities for the projected cycle : If , then is trivial. If , then we shall prove that is unbalanced. Indeed, the permutation induced on the branches by the preimage of a loop around is a permutation cycle. Therefore, the permutation induced on the branches must also be a permutation cycle. Let us denote by the order of that permutation cycle and by the order of the permutation cycle induced by a loop around on . Then
Applying the same arguments as for , one obtains that is unbalanced. As a consequence, is totally unbalanced with respect to . ∎
Now, consider the trigonometric version of the moment problem. Given a trigonometric polynomial , find all trigonometric polynomials such that
(8.15) |
By the change of variable , the trigonometric moment problem is equivalent to the following problem (see [1]): Given a Laurent polynomial (), find all such that
(8.16) |
Moreover, if , then by integration by parts,
Since , , are the coefficients of the power series at of
then for any fixed Laurent polynomial , a Laurent polynomial such that is a solution of (8.16) if and only if .
Proposition 8.4.
Proof.
Take close to infinity. By the Residue Theorem,
Assume that and take . By analytic continuation,
Since is transitive, if we sum the previous formula over , we obtain
Dividing by and multiplying by gives
Conversely, if is a solution of (2.4), then using the Residues Theorem in the two regions of the complementary of the unit circle,
Now, take . Note that when . Then for close to ,
Therefore, is a polynomial in , and the same holds for . To conclude, note that as , then . ∎
In the general rational case, Pakovich [25] has proved that if is a rational curve and are rational functions, then
if and only if for a finite number of cycles defined in terms of the the “dessin d’enfants” of .
9. The Tangential Center-Focus Problem in the Hyper-elliptic Case
Let us now consider the original center problem in the two-dimensional space. Let and consider the foliation
(9.17) |
deforming the initial foliation , where is a one-form. Let be a regular value of and let be a closed path in the leaf . Take a transversal to the leaves of parameterized by values of and consider the holonomy along and the corresponding displacement map (holonomy minus identity). As the path pushes to nearby closed paths , we have that . We say that the family is a center. We have
(9.18) |
Of course the Abelian integral in (9.18) depends only on the homology class of . We say that the family is a persistent center if and a tangential center if .
We consider more precisely the hyper-elliptic case , . Recall that the tangential center problem in the hyper-elliptic case for vanishing cycles was solved in [12]. Vanishing cycles are a particular class of simple cycles. We want to study the general tangential center problem in the hyper-elliptic case.
It is well-known (see for instance [33]) that any one-form is relatively cohomologous to the form
(9.19) |
for some , . Hence, the tangential center problem reduces to the problem of characterizing the vanishing of the integral
(9.20) |
along some cycle
We represent as a Riemann surface and consider its first homology group . Let be all the roots of . It is well-known that the homology of the fiber is generated by simple cycles going from to on one leaf of the Riemann surface followed by the lift of the same path on the other leaf travelled in the opposite direction. As the value of on the second half of the path is opposite to the value on the first part as well as the sense of travel, the two halves of the integral add up. Set
(9.21) |
the indefinite integral of . Up to the problem of multivaluedness of , this gives that the original tangential center problem in the plane reduces to the zero-dimensional tangential center problem
(9.22) |
where the zero-cycle is composed of the ramification points in the presentation of the class of in . Note that , that is, is indeed a cycle as the class of is generated by simple cycles.
The explicit dependence of on is not a problem. We must however be more careful here as the function is multivalued in due to the multivaluedness of the square root. It has the same (two-sheeted) Riemann surface as the function We will now be more precise in the choices we make in representing the cycle . We take big, order cyclically the roots and suppose for simplicity that is even. We present the Riemann surface of as a two-sheeted surface with cuts joining to , next to etc. When crossing any cut a path changes the sheet. Let be the cycle going first from to in the upper sheet and then along the same projection from to on the lower sheet. We suppose that the cycle avoids cuts from exterior (that is, it belongs to the complement of the convex hull of the cuts except for the ramification points ).
If is odd, we have to add a cut from to infinity, but we will stick rather to the case when is even for simplicity. The space provides a basis of . Any cycle can hence be written in a unique way as
(9.23) |
We associate the zero-cycle
(9.24) |
to the one-cycle . In fact, we have constructed an isomorphism
(9.25) |
Then
(9.26) |
We say that a cycle in is balanced (unbalanced, totally unbalanced) if the corresponding zero-cycle in is balanced (unbalanced, totally unbalanced).
Let denote the usual action of the monodromy group of on and let denote the conjugated group:
Then acts on .
The action of the group is very much related to the action of the monodromy group of , but in general is more complicated due to cuts in the Riemann surface of .
We believe that the following conjecture, analogous to Theorem 2.2, holds:
Conjecture 9.1.
Let , and let be given by (9.21).
-
(1)
If is a totally unbalanced cycle of , then
(9.27) if and only if there exist and analytic functions in a neighborhood of infinity such that , and the cycles are trivial for every , where
-
(2)
If is unbalanced, then (9.27) holds if and only if there exist , analytic in a neighborhood of infinity such that , , and for every , either the projected cycle is trivial or it is balanced and
The proof should go similarly as the proof of Theorem 2.2. The delicate point is the use of Lemma 6.1. If in Lemma 6.1 we could replace -irreducible and -invariant by -irreducible and -invariant, then we could prove the conjecture by using the isomorphism and then following the proof of Theorem 2.2. Here it is essential that the function is polynomial. The inductive argument is on the divisors of the degree of . The fact that the function is only analytic is not a problem. The difficulty is that is not a permutation of the roots as . In fact it acts on cycles and not on individual roots.
Example 9.2.
Let and assume that is even. Let be a balanced cycle of . Let be the polynomial associated to the cycle . Note that here the group coincides with because there is only one critical value and hence only one generator (corresponding to winding around zero). Hence Proposition 7.3 applies and vanishes if and only if , whenever does not divide , where and is the development of given in (9.21) in a neighborhood of infinity. Following Example 7.4, for , and the balanced cycle , with , we have that if and only if the function given by (9.21) is of the form , where the functions , for fixed are analytic in a neighborhood of infinity in .
Recall the definition of . We use the development in a neighborhood of infinity. There develops as times a function of . This gives that the powers of appearing in are by two higher than the powers in . One extra shift in powers comes from integration. This gives all together that the powers appearing in are by modulo higher than the powers in . Finally we get that if and only if the polynomial is of the form , with .
10. Tangential canard centers for generalized Van der Pol’s equations
Consider the singular perturbation of the generalized Van der Pol’s equation
(10.28) |
for certain rational functions .
As in [13] we blow up the singular point in the family and study the blown up system. A natural question is when after perturbing with the blow up parameter we obtain a center. The tangential version of this problem is to study when the perturbed equation has a center “to first order” in the parameter .
In this section on the example of Van der Pol’s singular perturbation, we show the relevance of the zero-dimensional tangential center problem for studying persistence of centers in slow-fast systems.
First, following [13], (10.28) can be transformed into the slow-fast system
where are rational functions and . We shall assume in addition that is a Morse function . Note that for every close enough to zero, has exactly two solutions also close to zero. Let denote those two solutions. Also, assume that
with for close to zero.
Now the system rewrites as
(10.29) |
We blow-up the family. Here we consider only one chart:
(10.30) |
Dividing by , the blown up vector field, has a center at on the divisor , for . We are interested in the persistence (or rather first order persistence of this center). With this notation, the second part of Theorem 1 of [14] can be formulated:
Proposition 10.1 ([14]).
Under above assumptions, for close to there is a -function defined on a transversal such that the closed orbits are given by .
Moreover,
and
(10.31) |
where
If for some and all , then we have that (10.29) has a center to first order in . Rewriting (10.31) as
where is a primitive of , the problem of characterizing is equivalent to the zero-dimensional tangential center problem for the cycle .
Proposition 5.3 states that is a solution iff there exist with such that , and . Therefore,
Now, replacing we get
for any and any such that and . Thus, given , it is possible to obtain all such that the perturbed system has a center at first order in .
11. Concluding remarks and open problems
The zero-dimensional tangential center problem (Problem 1.2) is a kind of toy example for studying the tangential center problem for the two-dimensional phase space (vanishing of Abelian integrals on one-dimensional cycles). In Section 9 we see how the initial problem in the hyper-elliptic case reduces to a problem in the zero-dimensional case. A first problem is to generalize our results and prove an analogue of Theorem 2.2 for general cycles on hyper-elliptic systems.
To completely solve the zero-dimensional tangential center problem it remains to solve the problem for balanced cycles. We believe that the existence of powers or Chebyshev polynomials as composition factors of plays a center role in balanced tangential centers. The study of balanced centers of was easy. The Chebyshev case should be only slightly more complicated. Example 7.4 shows that even the weak composition conjecture is not valid for the zero-dimensional tangential center problem. It would be very interesting to formulate a new plausible conjecture for necessary and sufficient condition of a tangential center.
Recall that Abelian integrals represent the linear term of the displacement function. If they vanish, one searches for the first non-zero term in the development of the displacement function. In the one-dimensional cycle case it is known that this first term is given by iterated integrals (see [8], [17], [20] and [21]). It would be interesting to define and study iterated integrals in the zero-dimensional case, too. It seems that they cannot exist for unbalanced cycles, but probably for balanced cycles it makes sense.
When studying the center problem one often develops the displacement function at the center. The center condition is the condition of vanishing of all coefficients in this development. This gives a growing sequence of ideals which stabilize. The order at which they stabilize is called the Bautin index. Bautin index was calculated by Bautin for quadratic vector fields [3]. Its calculation for general vector fields is an important unsolved problem. Zero-dimensional problem being more accessible could help to develop techniques for studying this problem.
Here we solved the polynomial tangential center problem for totally unbalanced cycles. It would be interesting to solve it for rational or Laurent systems. The Laurent systems case corresponds to trigonometric polynomials.
References
- [1] A. Álvarez, J.L. Bravo, C. Christopher, Some results on the trigonometric moment problem, preprint.
- [2] E. Artin, Galois theory, University of Notre Dame Press, Indiana, 1942.
- [3] N.N. Bautin, On the number of limit cycles appearing with variation of the coefficients from an equilibrium state of the type of a focus or a center, Mat. Sbornik N.S., 30(72) (1952) 181–196.
- [4] P. Bonnet, A. Dimca, Relative differential forms and complex polynomials, Bull. Sci. math., 124-7 (2000) 557–571.
- [5] M. Briskin, J.P. Françoise, Y. Yomdin, Center conditions, compositions of polynomials and moments on algebraic curve, Ergodic Theory Dyn. Syst., 19-5 (1999) 1201-1220.
- [6] M. Briskin, J.P. Françoise, Y. Yomdin, Center conditions II: Parametric and model center problems Isr. J. Math., 118 (2000) 61–82.
- [7] M. Briskin, N. Roytvarf, Y. Yomdin, Center conditions at infinity for Abel differential equation, Annals of Math., 172 (2010) 437–483
- [8] K.-T.Chen, Algebras of iterated path integrals and fundamental groups, Trans. AMS, 156 (1971) 359–379.
- [9] K.-T.Chen, Iterated Path Integrals, Bull. AMS, 83 (1977) 831–879.
- [10] S.N. Chow, J.K. Hale, Methods of Bifurcation Theory Springer-Verlag, New York, 1982.
- [11] C. Christopher, Abel equations: composition conjectures and the model problem, Bull. Lond. Math. Soc. 32-3, (2000) 332–338.
- [12] C. Christopher, P. Mardešić, The monodromy problem and the tangential center problem, Funkts. Anal. Prilozh., 44-1 (2010) 27–43.
- [13] F. Dumortier, R. Roussarie, Canard Cycles and Center Manifolds, Mem. Amer. Math. Soc., 121 (1996).
- [14] F. Dumortier, R. Roussarie, Multiple canard cycles in generalized Liénard equations, J. Differential Equations, 174 (2001) 1–29.
- [15] O. Forster, Lectures on Riemann surfaces, Graduate Texts in Mathematics 81, Springer-Verlag, New York-Berlin, 1981.
- [16] J.P. Françoise, F. Pakovich, Y. Yomdin, W. Zhao, Moment vanishing problem and positivity: some examples, to appear in Bul. Sci. Math.
- [17] L. Gavrilov, Petrov modules and zeros of Abelian integrals, Bull. Sci. Math., 122-8 (1998) 571–584.
- [18] L. Gavrilov, H. Movasati, The infinitesimal 16th Hilbert problem in dimension zero, Bull. Sci. math. 131 (2007) 242–257.
- [19] Yu. Ilyashenko, Appearance of limit cycles by perturbation of the equation , where is a polynomial, Mat. Sbornik (New Series) 78-3 (1969) 360–373.
- [20] A. Jebrane, P. Mardešić, M. Pelletier, A generalization of Françoise’s algorithm for calculating higher order Melnikov functions, Bull. Sci. Math., 126-9 (2002) 705–732.
- [21] A. Jebrane, P. Mardešić, M. Pelletier, A note on a generalization of Françoise’s algorithm for calculating higher order Melnikov functions, Bull. Sci. Math., 128-9 (2004) 749–760.
- [22] J. Muciño-Raymundo, Deformations of holomorphic foliations having a meromorphic first integral, J. Reine Angew. Math., 461 (1995) 189–219.
- [23] F. Pakovich, Counterexample to the Composition Conjecture, Proc. of the Amer. Math. Soc, 130-12 (2002) 3747-3749
- [24] F. Pakovich, Generalized “second Ritt theorem” and explicit solutions of the polynomial moment problem, preprint.
- [25] F. Pakovich, On rational functions orthogonal to all powers of a given rational function on a curve, preprint.
- [26] F. Pakovich, On polynomials orthogonal to all powers of a given polynomial on a segment, Isr. J. Math., 142 (2004) 273–283.
- [27] F. Pakovich, Prime and Composite Laurent Polynomials, Bull. Sci. Math., 133-7 (2009) 693–732.
- [28] F. Pakovich, M. Muzychuk, Solution of the polynomial moment problem, Proc. Lond. Math. Soc., 99-3 (2009) 633–657.
- [29] F. Pakovich, C. Pech, A.K. Zvonkin, Lauren polynomial problem: a case study, Contem. Math., 532 (2010) 177–194.
- [30] J.F. Ritt, Prime and composite polynomials, Trans. Amer. Math. Soc., 23 (1922) 51–66.
- [31] B. L. van der Waerden, Algebra, Vol. 1. Frederick Ungar Publishing Co., New York, 1970.
- [32] H. Wielandt, Finite Permutation Groups, Academic Press, New York and London, 1964.
- [33] H. Zoladek, The Monodromy Group, Mathematics Institute of the Polish Academy of Science. Mathematical Monographs (New Series), 67, Birkhäuser Verlag, Basel, 2006.