This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Vanishing Abelian Integrals on Zero-Dimensional Cycles

A. Álvarez1, J.L. Bravo1, P. Mardešić2
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 f(z)f(z), characterize all rational functions g(z)g(z) and zero-sum integers {ni}\{n_{i}\} such that the function tnig(zi(t))t\mapsto\sum n_{i}g(z_{i}(t)) vanishes identically. Here zi(t)z_{i}(t) are continuously depending roots of f(z)tf(z)-t. We introduce a notion of (un)balanced cycles. Our main result is an inductive solution of the problem of vanishing of Abelian integrals when f,gf,g 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 system
2010 Mathematics Subject Classification:
34C07, 34C08, 34D15, 34M35
The first author was partially supported by Junta de Extremadura and FEDER funds. The second author was partially supported by Junta de Extremadura and a MCYT/FEDER grant number MTM2008-05460. The first two authors are grateful to the Université de Bourgogne, for the hospitality and support during the visit when this work was started. The third author thanks the Universidad de Extremadura for the hospitality and support during the visit when this work was finished.

1. 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 f(z)f\in\mathbb{C}(z) be a rational function of degree m>1m>1. Let (zi(t))1im(z_{i}(t))_{1\leq i\leq m} denote an mm-tuple of analytic preimages zi(t)f1(t)z_{i}(t)\in f^{-1}(t), where t\Σt\in\mathbb{C}\backslash\Sigma and Σ\Sigma is the set of critical values of ff. Define a zero-dimensional cycle (shorter cycle) C(t)C(t) of ff as the sum

(1.1) C(t)=i=1mnizi(t),withi=1mni=0,ni.C(t)=\sum_{i=1}^{m}n_{i}z_{i}(t),\quad\text{with}\quad\sum_{i=1}^{m}n_{i}=0,\quad n_{i}\in\mathbb{Z}.

Zero-dimensional cycles form the reduced 0-th homology group denoted H~0(f1(t))\tilde{H}_{0}(f^{-1}(t)). We say that a cycle C(t)C(t) is simple if it is of the form C(t)=zj(t)zi(t)C(t)=z_{j}(t)-z_{i}(t), and trivial if ni=0n_{i}=0 for every ii.

Let g(z)g\in\mathbb{C}(z) be a rational function. As in [18], define zero-dimensional Abelian integrals of gg along the cycle C(t)C(t) by

(1.2) C(t)g:=i=1mnig(zi(t)).\int_{C(t)}g:=\sum_{i=1}^{m}n_{i}g(z_{i}(t)).

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 ff, gg such that for any family of cycles C(t)C(t) of ff the Abelian integral C(t)g\int_{C(t)}g vanishes identically.

Problem 1.2.

(Tangential center problem) Characterize rational functions ff, gg and cycles C(t)C(t) of ff such that the Abelian integral C(t)g\int_{C(t)}g 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 f,gf,g 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 zi(t)z_{i}(t) as tt 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 C(t)C(t) of ff the Abelian integral C(t)g\int_{C(t)}g vanishes if and only if there exist decompositions f=f0hf=f_{0}\circ h and g=g0hg=g_{0}\circ h such that the projected cycle h(C(t))h(C(t)) is trivial. A simple counter-example was given by Pakovich [27].

This allows the formulation of a weaker composition conjecture: The integral C(t)g\int_{C(t)}g vanishes if and only if there exist decompositions f=fkhkf=f_{k}\circ h_{k}, g=g1h1++gshsg=g_{1}\circ h_{1}+\cdots+g_{s}\circ h_{s} such that the projected cycles hi(C(t))h_{i}(C(t)), i=1,,si=1,\ldots,s 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 f,g(z)f,g\in\mathbb{C}(z), the following conditions are equivalent:

  1. (1)

    C(t)g0\int_{C(t)}g\equiv 0 for every (simple) cycle C(t)C(t).

  2. (2)

    There exists g0(z)g_{0}\in\mathbb{C}(z), such that g=g0fg=g_{0}\circ f.

In order to formulate the main theorem, Theorem 2.2, we must define the notions of unbalanced, totally unbalanced and projected cycles. Let ff be a polynomial of degree mm and Σ\Sigma\subset\mathbb{C} its set of critical values. Then f:f1(Σ)Σf:\mathbb{C}\setminus f^{-1}(\Sigma)\to\mathbb{C}\setminus\Sigma defines a fibration, with fiber f1(t)f^{-1}(t) consisting of m=deg(f)m=\deg(f) points. Take a base point t0Σt_{0}\in\mathbb{C}\setminus\Sigma. Then this fibration defines a mapping from the first homotopy group π1(Σ,t0)\pi_{1}(\mathbb{C}\setminus\Sigma,t_{0}) to the group of automorphisms Aut(f1(t0))Aut(f^{-1}(t_{0})). Its image is called the monodromy group GfG_{f}. It acts transitively on the fiber f1(t0)f^{-1}(t_{0}) (see, e.g., [33]). Changing the base point t0t_{0} conjugates all elements of the group, so the abstract group GfG_{f} does not depend on the choice of the base point t0t_{0}. Consider the loop in π1(Σ,t0)\pi_{1}(\mathbb{C}\setminus\Sigma,t_{0}) winding anti-clockwise around all critical values in Σ\Sigma. It corresponds to a permutation cycle τ\tau_{\infty} of GfG_{f} of order mm. We label the roots zi(t0)z_{i}(t_{0}), i=1,,mi=1,\ldots,m of f(z0)=tf(z_{0})=t so that this permutation shifts the indices of the roots by one. The choice of the first root is arbitrary.

Let Γm(f)Gf\Gamma_{m}(f)\subset G_{f} denote the conjugacy class of τ\tau_{\infty}, that is, the set of all στσ1\sigma\circ\tau_{\infty}\circ\sigma^{-1}, for any σGf\sigma\in G_{f}. Note that Γm(f)\Gamma_{m}(f) is the set of permutations induced by all paths winding counter-clockwise once around infinity.

Definition.

We say that a cycle C(t)C(t) of ff is balanced if

i=1mnpiϵmi=0,for every σ=(p1,p2,,pm)Γm(f),\sum_{i=1}^{m}n_{p_{i}}\epsilon_{m}^{i}=0,\quad\text{for every }\sigma=(p_{1},p_{2},\ldots,p_{m})\in\Gamma_{m}(f),\quad

where ϵm\epsilon_{m} is any primitive mm-th root of unity. If C(t)C(t) is not balanced, we say that C(t)C(t) is unbalanced.

For each f0,h[z]f_{0},h\in\mathbb{C}[z] such that f=f0hf=f_{0}\circ h and each cycle C(t)C(t) of ff we define a cycle h(C(t)){h}({C}(t)) of f0{f}_{0} called the projected cycle by

(2.3) h(C(t))=h(zi(t))(h(zj)=h(zi)nj)h(zi(t))=j=1dn~jwj(t).{h}(C(t))=\sum_{{h}(z_{i}(t))}\left(\sum_{{h}(z_{j})={h}(z_{i})}n_{j}\right){h}(z_{i}(t))=\sum_{j=1}^{d}\tilde{n}_{j}w_{j}(t).

Here w1(t),wd(t)w_{1}(t),\ldots w_{d}(t) are all the different roots h(zi(t)){h}(z_{i}(t)) of f0(z)=tf_{0}(z)=t.

We say that C(t)C(t) is totally unbalanced when for every f0,hf_{0},h such that f=f0hf=f_{0}\circ h, the projected cycle h(C)h({C}) is unbalanced or trivial.

The notion of balanced cycle C(t)H~0(f1(t))C(t)\in\tilde{H}_{0}(f^{-1}(t)) is related to the behaviour of the branches zi(t)z_{i}(t) when tt tends to infinity. It is well defined, i.e., independent on the way how permutation cycles are written. Indeed, if we write σ=(p2,,pm,p1)\sigma=(p_{2},\ldots,p_{m},p_{1}), then

(np2ϵm+np3ϵm2++npmϵmm1+np1)ϵm=i=1mnpiϵmi.\left(n_{p_{2}}\epsilon_{m}+n_{p_{3}}\epsilon_{m}^{2}+\ldots+n_{p_{m}}\epsilon_{m}^{m-1}+n_{p_{1}}\right)\epsilon_{m}=\sum_{i=1}^{m}n_{p_{i}}\epsilon_{m}^{i}.

Thus, i=1mnpiϵmi=0\sum_{i=1}^{m}n_{p_{i}}\epsilon_{m}^{i}=0 if and only if

np2ϵm+np3ϵm2++npmϵmm1+np1=0.n_{p_{2}}\epsilon_{m}+n_{p_{3}}\epsilon_{m}^{2}+\ldots+n_{p_{m}}\epsilon_{m}^{m-1}+n_{p_{1}}=0.

Note that in particular totally unbalanced cycles are unbalanced as seen from the trivial decomposition f=fidf=f\circ\text{id}.

Theorem 2.2.

Let f[z]f\in\mathbb{C}[z].

  1. (1)

    If C(t)C(t) is a totally unbalanced cycle, then

    (2.4) C(t)g0\int_{C(t)}g\equiv 0

    if and only if there exist f1,,fs,g1,,gs,h1,,hs[z]f_{1},\ldots,f_{s},g_{1},\ldots,g_{s},h_{1},\ldots,h_{s}\in\mathbb{C}[z] such that f=fkhkf=f_{k}\circ h_{k}, g=g1h1++gshsg=g_{1}\circ h_{1}+\ldots+g_{s}\circ h_{s} and the projected cycles hk(C)h_{k}(C) are trivial for every k=1,2,,sk=1,2,\ldots,s.

  2. (2)

    If C(t)C(t) is an unbalanced cycle, then g[z]g\in\mathbb{C}[z] is a solution of (2.4) if and only if there exist f1,,fs,g1,,gs,h1,,hs[z]f_{1},\ldots,f_{s},g_{1},\ldots,g_{s},h_{1},\ldots,h_{s}\in\mathbb{C}[z] such that f=fkhkf=f_{k}\circ h_{k}, g=g1h1++gshsg=g_{1}\circ h_{1}+\ldots+g_{s}\circ h_{s}, and for every kk

    hk(C(t))gk0,\int_{{h_{k}}({C}(t))}g_{k}\equiv 0,

    and the projected cycle hk(C){h_{k}}({C}) 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 mm is prime, then there are no balanced cycles. We also completely solve the problem for f(z)=zmf(z)=z^{m}.

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 ff in searching for all polynomials gg such that

01fk(z)g(z)𝑑z=0,for every k.\int_{0}^{1}f^{k}(z)g^{\prime}(z)\,dz=0,\quad\text{for every }k\in\mathbb{N}.

In Section 8 we show that the moment problem is equivalent to solving Problem 1.2 for a concrete cycle Cf(t)C_{f}(t). 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 gg can be written as a sum of at most three reducible solutions (a polynomial gg is a reducible solution if there exists f0,g0,hf_{0},g_{0},h such that h(C(t))g00\int_{h(C(t))}g_{0}\equiv 0, f=f0hf=f_{0}\circ h and g=g0hg=g_{0}\circ h). 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 C(t)g\int_{C(t)}g, C(t)H~0(f1(t))C(t)\in\tilde{H}_{0}(f^{-1}(t)) are defined for a couple ff and gg of polynomials in one variable.

The definition is analogous to the definition of one-dimensional Abelian integrals γ(t)G\int_{\gamma(t)}G, γ(t)H1(F1(t)\gamma(t)\in H_{1}(F^{-1}(t), defined for a polynomial FF and a polynomial differential 11-form η=G1dx+G2dy\eta=G_{1}dx+G_{2}dy in two variables. In this section we push further the analogy between systems depending on one variable and planar systems. Assume that some point p2p\in\mathbb{R}^{2} is a strict local extremum of a real polynomial function F[x,y]F[x,y]. Then the system dF(x,y)=0dF(x,y)=0 has a center at pp, parametrized by a family of one-cycles γ(t)F1(t)\gamma(t)\in F^{-1}(t). Consider the deformation

(3.5) dF+ϵη=0.dF+\epsilon\eta=0.

For most deformations η\eta, the center will be broken. A natural question is for what deformations η\eta, 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 γ(t)\gamma(t) parametrized by FF. The displacement function Dϵ(t)D_{\epsilon}(t) 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 Dϵ0D_{\epsilon}\equiv 0. A classical result is that Abelian integrals give the first order term of the displacement function:

(3.6) Dϵ(t)=ϵγ(t)η+o(ϵ).D_{\epsilon}(t)=-\epsilon\int_{\gamma(t)}\eta+o(\epsilon).

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 γ(t0)F1(t)\gamma(t_{0})\subset F^{-1}(t) to nearby leaves the holonomy associated to γ(t0)\gamma(t_{0}) is defined. By assumption, for ϵ=0\epsilon=0 all holonomies are identity (i.e., the associated displacement functions are zero-functions).

Consider now the problem in one-dimensional space. Let f(z)f\in\mathbb{C}(z) be a function and C(t)H~0(f1(t))C(t)\in\tilde{H}_{0}(f^{-1}(t)) a family of cycles of ff. Consider a perturbation of ff by gg

f+ϵgf+\epsilon g

and the deformed cycle Cϵ(t)=i=1mnizi(t,ϵ)C_{\epsilon}(t)=\sum_{i=1}^{m}n_{i}z_{i}(t,\epsilon), where (f+ϵg)(zi(t,ϵ))=t(f+\epsilon g)(z_{i}(t,\epsilon))=t, for t\Σϵt\in\mathbb{C}\backslash\Sigma_{\epsilon}, where Σϵ\Sigma_{\epsilon} is the set of critical values of f+ϵgf+\epsilon g. We define the displacement function of f+ϵgf+\epsilon g along the cycle C(t)C(t) by

Δϵ(t)=i=1mnif(zi(t,ϵ))=Cϵ(t)f.\Delta_{\epsilon}(t)=\sum_{i=1}^{m}n_{i}f(z_{i}(t,\epsilon))=\int_{C_{\epsilon}(t)}f.

As in the case of one-dimensional loops, zero-dimensional Abelian integrals correspond to the principal part of the displacement function of the perturbation f+ϵgf+\epsilon g along a family of zero-dimensional cycles. Indeed, from (f+ϵg)(zi(t,ϵ))=t(f+\epsilon g)(z_{i}(t,\epsilon))=t, it follows that i=1mni(f+ϵg)(zi(t,ϵ))=0\sum_{i=1}^{m}n_{i}(f+\epsilon g)(z_{i}(t,\epsilon))=0, giving

(3.7) Δϵ(t)=ϵCϵ(t)g=ϵC(t)g+o(ϵ).\Delta_{\epsilon}(t)=-\epsilon\int_{C_{\epsilon}(t)}g=-\epsilon\int_{C(t)}g+o(\epsilon).

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 ff, gg such that for any family of cycles C(t)C(t) of ff the displacement function Δϵ\Delta_{\epsilon} of f+ϵgf+\epsilon g, along any zero cycle C(t)C(t) vanishes identically.

Problem 3.2.

(Infinitesimal center problem) Characterize rational functions ff, gg and cycles C(t)C(t) of ff such that the displacement function Δϵ\Delta_{\epsilon} of f+ϵgf+\epsilon g, along the family of zero-cycles C(t)C(t) vanishes identically.

Solution of Problem 3.2 is a corollary to Theorem 2.1.

Corollary 3.3.

Given f,g(z)f,g\in\mathbb{C}(z), the following conditions are equivalent:

  1. (1)

    Δϵ(t)0\Delta_{\epsilon}(t)\equiv 0 for every (simple) cycle C(t)C(t).

  2. (2)

    There exists g0(z)g_{0}\in\mathbb{C}(z), such that g=g0fg=g_{0}\circ f.

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 f(z)f\in\mathbb{C}(z) of degree mm, let Σ\Sigma\subset\mathbb{C} be the set of critical values of ff. The monodromy group GfG_{f} is defined as in the polynomial case and acts transitively on generic fibers f1(t),f^{-1}(t), tΣt\in\mathbb{C}\setminus\Sigma.

As shown in [15] the monodromy group GfG_{f} is the Galois group of the Galois extension of (t)\mathbb{C}(t) by the mm preimages z1(t),,zm(t)z_{1}(t),\ldots,z_{m}(t) of tt by ff, that is,

Gf=Aut((z1,,zm)/(t)).G_{f}=Aut(\mathbb{C}(z_{1},\ldots,z_{m})/\mathbb{C}(t)).

Let X={1,,m}X=\{1,\ldots,m\} and let GSmG\subseteq S_{m} be a transitive permutation group on XX. A subset BXB\subseteq X is called a block ([32]) of GG if for each gGg\in G the image set g(B)g(B) and BB are either equal or disjoint. Given a block BB, the set ={g(B)|gG}\mathcal{B}=\{g(B)|g\in G\} forms a partition of XX into disjoint blocks of the same cardinality which is called an imprimitivity system (or a complete block system). Each permutation group GSmG\subseteq S_{m} has two trivial imprimitivity systems: {X}\{X\} and {{x}}xX\{\{x\}\}_{x\in X}.

Given f(z)(z)f(z)\in\mathbb{C}(z) we say that two decompositions of ff, f=f1h1f=f_{1}\circ h_{1} and f=f2h2f=f_{2}\circ h_{2}, are equivalent if h1h_{1} and h2h_{2} define the same field over every preimage of tt by f(z)f(z), that is, if (h1(zi))=(h2(zi))\mathbb{C}(h_{1}(z_{i}))=\mathbb{C}(h_{2}(z_{i})) for every i=1,,mi=1,\ldots,m.

Proposition 4.1.

Let f(z)(z)f(z)\in\mathbb{C}(z). There exists a one-to-one correspondence between imprimitivity systems of GfG_{f} and equivalence classes of decompositions of ff.

Moreover, if f=f0hf=f_{0}\circ h, h\mathcal{B}_{h} is the imprimitivity system corresponding to hh and iBhi\in B\in\mathcal{B}_{h}, then B={k:h(zk)=h(zi)}B=\{k\colon h(z_{k})=h(z_{i})\}.

Remark 4.2.

Let us point out that the second statement in the proposition is in particular telling us that given a cycle C(t)=i=1mnizi(t)C(t)=\sum_{i=1}^{m}n_{i}z_{i}(t) of ff, the projected cycle h(C(t))h(C(t)) of f0f_{0} given by the expression (2.3) can be written as

h(C(t))=h(zi(t))(jBinj)h(zi(t)),h(C(t))=\sum_{h(z_{i}(t))}\left(\sum_{j\in B_{i}}n_{j}\right)h(z_{i}(t)),

where BiB_{i} is the block containing ii of the imprimitivity system \mathcal{B} corresponding to the decomposition f=f~hf=\tilde{f}\circ h.

Proof.

Let f(z)f(z) be a rational function of degree mm, and let us denote by z1(t),,zm(t)z_{1}(t),\ldots,z_{m}(t) the preimages of t\Σt\in\mathbb{C}\backslash\Sigma as above.

Let f=f0hf=f_{0}\circ h be a decomposition and let us find the associated imprimitivity system h\mathcal{B}_{h}. To do so, we only have to find a block BB, since by definition, h:={σ(B):σGf}\mathcal{B}_{h}:=\{\sigma(B):\sigma\in G_{f}\} is an imprimitivity system of GfG_{f}. For any preimage ziz_{i} of t\Σt\in\mathbb{C}\backslash\Sigma, we have that

t=f(zi)=f0(h(zi)),t=f(z_{i})=f_{0}(h(z_{i})),

and therefore

(t)(h(zi))(zi)(z1,,zm).\mathbb{C}(t)\subseteq\mathbb{C}(h(z_{i}))\subseteq\mathbb{C}(z_{i})\subseteq\mathbb{C}(z_{1},\ldots,z_{m}).

By the Fundamental Theorem of Galois theory (in Artin’s version, see [2, 31]) there exists a subgroup HH of the monodromy group GfG_{f} containing HiH_{i}, the stabilizer of ii, such that

(h(zi))=(z1,,zm)H.\mathbb{C}(h(z_{i}))=\mathbb{C}(z_{1},\ldots,z_{m})^{H}.

As a consequence, h(zi)=τ(h(zi))=h(τ(zi))h(z_{i})=\tau(h(z_{i}))=h(\tau(z_{i})) for every τH\tau\in H. The orbit BB of ii by the action of this subgroup HH is the block we are looking for:

B:={σ(i):σH}H/Hi.B:=\{\sigma(i):\sigma\in H\}\simeq H/H_{i}.

It is obvious that τ(B)=B\tau(B)=B for every τH\tau\in H, and it is also immediate to see that σ(B)B=\sigma(B)\cap B=\emptyset when σGf\H\sigma\in G_{f}\backslash H. Then, we also have proved that h(zi)h(z_{i}) is constant on iBhi\in B\in\mathcal{B}_{h}. Conversely, if h(zi)=h(zk)h(z_{i})=h(z_{k}), then kBk\in B: since the monodromy group GfG_{f} is transitive, then there exists some σGf\sigma\in G_{f} such that σ(i)=k\sigma(i)=k, and therefore

σ(h(zi))=h(σ(zi))=h(zk)=h(zi),\sigma(h(z_{i}))=h(\sigma(z_{i}))=h(z_{k})=h(z_{i}),

which implies that h(zi)h(z_{i}) is invariant by σ\sigma. Then σ\sigma must belong to HH and kk belongs to BB. For another different block B~=σ(B)\tilde{B}=\sigma(B), a similar argument proves the same statement. Then, the blocks of the imprimitivity system h\mathcal{B}_{h} verify that h(zi)=h(zj)h(z_{i})=h(z_{j}) if and only if i,ji,j belongs to the same block of h\mathcal{B}_{h}, proving that the imprimitivity system we get does not depend on the choice of ziz_{i}.

Now let \mathcal{B} be an imprimitivity system for GfG_{f}. The block BB containing the element ii provides a subgroup of GfG_{f} containing the stabilizer HiH_{i} of ii. Precisely,

GB:={τGf:τ(B)=B}.G_{B}:=\{\tau\in G_{f}:\tau(B)=B\}.

By the Fundamental Theorem, the following inclusions of groups

{Id}HiGBGf\{{\rm Id}\}\subseteq H_{i}\subseteq G_{B}\subseteq G_{f}

yield the following inclusions of fields

(t)LB:=(z1,,zm)GB(zi)(z1,,zm).\mathbb{C}(t)\subseteq L_{B}:=\mathbb{C}(z_{1},\ldots,z_{m})^{G_{B}}\subseteq\mathbb{C}(z_{i})\subseteq\mathbb{C}(z_{1},\ldots,z_{m}).

By Lüroth’s Theorem [31] applied to LB(zi)\mathbb{C}\subset L_{B}\subseteq\mathbb{C}(z_{i}), LB=(h(zi))L_{B}=\mathbb{C}(h(z_{i})) for some rational function hh. Since f(zi)=tLBf(z_{i})=t\in L_{B}, there exists another rational function f0f_{0} such that f=f0hf=f_{0}\circ h. Moreover, h(zi)=h(zj)h(z_{i})=h(z_{j}) for any i,jBi,j\in B.

If we choose an element kk of a different block B~\tilde{B}, then we get another subgroup GB~G_{\tilde{B}} of GfG_{f}, which contains the stabilizer HkH_{k} of kk. Again by the Fundamental Theorem, it provides a field LB~:=(z1,,zm)GB~L_{\tilde{B}}:=\mathbb{C}(z_{1},\ldots,z_{m})^{G_{\tilde{B}}}, which by Lüroth’s Theorem is generated by some rational function h~\tilde{h}, that is, LB~=(h~(zk))L_{\tilde{B}}=\mathbb{C}(\tilde{h}(z_{k})). Therefore, there exists another rational function f~0\tilde{f}_{0} such that f=f~0h~f=\tilde{f}_{0}\circ\tilde{h}. Let us see how different these functions hh and h~\tilde{h} are. Since GfG_{f} is transitive, there exists a permutation τ\tau such that τ(i)=k\tau(i)=k. Therefore, τ(B)=B~\tau(B)=\tilde{B} (or B=τ1(B~)B=\tau^{-1}(\tilde{B})), and it is immediate to check that

(4.8) τGBτ1=GB~.\tau\circ G_{B}\circ\tau^{-1}=G_{\tilde{B}}.

Since by the Fundamental Theorem we know that GBG_{B} and GB~G_{\tilde{B}} are the Galois groups of (z1,,zm)\mathbb{C}(z_{1},\ldots,z_{m}) over (h(zi))\mathbb{C}(h(z_{i})) and (h~(zk))\mathbb{C}(\tilde{h}(z_{k})) respectively, let us see what equality (4.8) gives: an element σGB~\sigma^{\prime}\in G_{\tilde{B}} is an algebra automorphism of (z1,,zm)\mathbb{C}(z_{1},\ldots,z_{m}) such that σ(h~(zk))=h~(zk)\sigma^{\prime}(\tilde{h}(z_{k}))=\tilde{h}(z_{k}). Since σ=τστ1\sigma^{\prime}=\tau\circ\sigma\circ\tau^{-1} for some σGB\sigma\in G_{B}, that means

τ(σ(τ1(h~(zk))))=h~(zk)τ(σ(h~(zi)))τ(h~(zi)),\begin{matrix}\tau(\sigma(\tau^{-1}(\tilde{h}(z_{k}))))&=&\tilde{h}(z_{k})\\ \parallel&&\parallel\\ \tau(\sigma(\tilde{h}(z_{i})))&&\tau(\tilde{h}(z_{i})),\end{matrix}

which implies that σ(h~(zi))=h~(zi)\sigma(\tilde{h}(z_{i}))=\tilde{h}(z_{i}). Also, as σ=τ1στ\sigma=\tau^{-1}\circ\sigma^{\prime}\circ\tau leaves invariant h(zi)h(z_{i}), it holds that

τ1(σ(τ(h(zi))))=h(zi)τ1(σ(h(zk)))τ1(h(zk)),\begin{matrix}\tau^{-1}(\sigma^{\prime}(\tau(h(z_{i}))))&=&h(z_{i})\\ \parallel&&\parallel\\ \tau^{-1}(\sigma^{\prime}(h(z_{k})))&&\tau^{-1}(h(z_{k})),\end{matrix}

which implies that σ(h(zk))=h(zk)\sigma^{\prime}(h(z_{k}))=h(z_{k}). To sum up, equality (4.8) implies that GBG_{B} is not only the Galois group of (z1,,zm)\mathbb{C}(z_{1},\ldots,z_{m}) over (h(zi))\mathbb{C}(h(z_{i})), but also over (h~(zi))\mathbb{C}(\tilde{h}(z_{i})). Similarly, GB~G_{\tilde{B}} is not only the Galois group of (z1,,zm)\mathbb{C}(z_{1},\ldots,z_{m}) over (h~(zk))\mathbb{C}(\tilde{h}(z_{k})), but also over (h(zk))\mathbb{C}(h(z_{k})). By the Fundamental Theorem, we obtain that LB=(h(zi))=(h~(zi))L_{B}=\mathbb{C}(h(z_{i}))=\mathbb{C}(\tilde{h}(z_{i})) and LB~=(h~(zk))=(h(zk))L_{\tilde{B}}=\mathbb{C}(\tilde{h}(z_{k}))=\mathbb{C}(h(z_{k})). Therefore, given an imprimitivity system \mathcal{B}, we get not only one decomposition of ff, 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 Δϵ\Delta_{\epsilon} (and hence also the corresponding Abelian integral) of the family f+ϵgf+\epsilon g, f,g(z)f,g\in\mathbb{C}(z), along a family of cycles C(t)C(t) of ff.

Proposition 5.1.

Assume that f=f0hf=f_{0}\circ h, g=g0hg=g_{0}\circ h, where f0,g0,h(z)f_{0},g_{0},h\in\mathbb{C}(z). Let C(t)=knkzk(t)C(t)=\sum_{k}n_{k}z_{k}(t) be a cycle of ff and Cϵ(t)=knkzk(t,ϵ)C_{\epsilon}(t)=\sum_{k}n_{k}z_{k}(t,\epsilon) its continuation as a cycle of f+ϵgf+\epsilon g. Let

h(C(t))=h(zk(t))(h(zj(t))=h(zk(t))nj)h(zk(t))h(C(t))=\sum_{h(z_{k}(t))}\left(\sum_{h(z_{j}(t))=h(z_{k}(t))}n_{j}\right)h(z_{k}(t))

and

h(Cϵ(t))=h(zk(t,ϵ))(h(zj(t,ϵ))=h(zk(t,ϵ))nj)h(zk(t,ϵ))h(C_{\epsilon}(t))=\sum_{h(z_{k}(t,\epsilon))}\left(\sum_{h(z_{j}(t,\epsilon))=h(z_{k}(t,\epsilon))}n_{j}\right)h(z_{k}(t,\epsilon))

be the cycle of f0f_{0} and f0+ϵg0f_{0}+\epsilon g_{0} obtained by projection of C(t)C(t) and Cϵ(t)C_{\epsilon}(t) by hh respectively. Then,

  1. (1)

    Δϵ(t)=h(Cϵ(t))f0\Delta_{\epsilon}(t)=\int_{h(C_{\epsilon}(t))}f_{0} and C(t)g=h(C(t))g0\int_{C(t)}g=\int_{h(C(t))}g_{0}.

  2. (2)

    In particular, if the projected cycle h(C(t))h(C(t)) is trivial, then h(Cϵ(t))h(C_{\epsilon}(t)) is trivial,

    Δϵ(t)0\Delta_{\epsilon}(t)\equiv 0

    and

    C(t)g0.\int_{C(t)}g\equiv 0.
Remark 5.2.

In other words, the assumption (2) of the proposition can be reformulated by saying that the cycle C(t)C(t) projects by hh to a trivial cycle

h(C(t))=f0(wi(t))=t(h(zk(t))=wi(t)nk)wi(t)h(C(t))=\sum_{f_{0}(w_{i}(t))=t}\left(\sum_{h(z_{k}(t))=w_{i}(t)}n_{k}\right)w_{i}(t)

of f0f_{0} and, moreover, the function gg factors through the same hh, so a change of coordinates leading to integration on the trivial cycle is possible.

Proof.

(1) Let C(t)=knkzk(t)C(t)=\sum_{k}n_{k}z_{k}(t) be a cycle and assume that f=f0hf=f_{0}\circ h, g=g0hg=g_{0}\circ h. Set wi(t,ϵ)=h(zi(t,ϵ))w_{i}(t,\epsilon)=h(z_{i}(t,\epsilon)). Then

Δϵ(t)=i=1mnif(zi(t,ϵ))=h(zi(t,ϵ))(h(zj(t,ϵ))=h(zi(t,ϵ))ni)f0(wj(t,ϵ))=h(Cϵ(t))f0.\begin{split}\Delta_{\epsilon}(t)&=\sum_{i=1}^{m}n_{i}f(z_{i}(t,\epsilon))=\sum_{h(z_{i}(t,\epsilon))}\left(\sum_{h(z_{j}(t,\epsilon))=h(z_{i}(t,\epsilon))}n_{i}\right)f_{0}(w_{j}(t,\epsilon))\\ &=\int_{h(C_{\epsilon}(t))}f_{0}.\end{split}

The second claim of (1) follows the same way.

(2) If h(C(t))h(C(t)) is a trivial cycle, then all its coefficients are equal to zero. The dependence with respect to ϵ\epsilon being continuous, it follows that the cycle h(Cϵ(t))h(C_{\epsilon}(t)) 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 f(z)(z)f(z)\in\mathbb{C}(z), and consider a simple cycle C(t)=zi(t)zj(t)C(t)=z_{i}(t)-z_{j}(t). Then

C(t)g0\int_{C(t)}g\equiv 0

if and only if there exist f0,g0,h(z)f_{0},g_{0},h\in\mathbb{C}(z) such that f=f0hf=f_{0}\circ h, g=g0hg=g_{0}\circ h and h(zi(t))=h(zj(t))h(z_{i}(t))=h(z_{j}(t)).

Remark 5.4.

Let us observe that since h(zi(t))=h(zj(t))h(z_{i}(t))=h(z_{j}(t)), by Proposition 4.1 the set {i,j}\{i,j\} is included in some block BB of the imprimitivity system h\mathcal{B}_{h} corresponding to the decomposition f=f0hf=f_{0}\circ h.

Proposition 5.3 essentially follows from Theorem 4 of [18] and Lüroth’s Theorem, but we include the proof for completeness.

Proof.

The sufficient condition is a consequence of Proposition 5.1.

For the necessary condition, let us define

𝒲={h(z):h(zi(t))=h(zj(t))}.\mathcal{W}=\{h\in\mathbb{C}(z)\colon h(z_{i}(t))=h(z_{j}(t))\}.

Since

𝒲(z),\mathbb{C}\subset\mathcal{W}\subset\mathbb{C}(z),

by Lüroth’s Theorem 𝒲=(h(z))\mathcal{W}=\mathbb{C}(h(z)) for some rational function hh. Since f,g𝒲f,g\in\mathcal{W}, then f=f0hf=f_{0}\circ h, g=g0hg=g_{0}\circ h for some f0,g0(z)f_{0},g_{0}\in\mathbb{C}(z). ∎

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 (2)(2) implies (1)(1), similarly in Corollary 3.3 (2)(2) implies (1)(1). Therefore, to conclude it is sufficient to prove that in Theorem 2.1 (1)(1) implies (2)(2).

Let g(z)g\in\mathbb{C}(z) be such that C(t)g0\int_{C(t)}g\equiv 0 for every (simple) cycle C(t)C(t). First, let us observe that the condition C(t)g0\int_{C(t)}g\equiv 0 for every (simple) cycle C(t)C(t) is equivalent to

g(z1(t))=g(z2(t))==g(zm(t)).g(z_{1}(t))=g(z_{2}(t))=\ldots=g(z_{m}(t)).

Therefore, g(zi)(z1,,zm)g(z_{i})\in\mathbb{C}(z_{1},\ldots,z_{m}) is invariant under the action of the whole monodromy group GfG_{f}, that is, g(zi)(z1,,zm)Gf=(t)g(z_{i})\in\mathbb{C}(z_{1},\ldots,z_{m})^{G_{f}}=\mathbb{C}(t), which means that there exists g0(z)(z)g_{0}(z)\in\mathbb{C}(z) such that g(zi(t))=g0(t)=g0(f(zi(t)))g(z_{i}(t))=g_{0}(t)=g_{0}(f(z_{i}(t))). Then g=g0fg=g_{0}\circ f. ∎

Under a generic condition on ff we show that the existence of a persistent center or a tangential center for any non-trivial cycle C(t)C(t) is equivalent to the integrability of the system as characterized in Theorem 2.1. Recall that a point z0z_{0} is of Morse type if f(z)=f(z0)+a(zz0)2+(zz0)2f(z)=f(z_{0})+a(z-z_{0})^{2}+\circ(z-z_{0})^{2}, aa\in\mathbb{C}. Generically all critical points are of Morse type and all critical values are different.

Proposition 5.5.

Given f(z)f\in\mathbb{C}(z), 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 C(t)C(t) the following conditions are equivalent:

  1. (1)

    C(t)g0\int_{C(t)}g\equiv 0 .

  2. (2)

    Δϵ(t)0\Delta_{\epsilon}(t)\equiv 0 along C(t)C(t).

  3. (3)

    There exists g0(z)g_{0}\in\mathbb{C}(z) such that g=g0fg=g_{0}\circ f.

Proof.

We prove that (1)(1) implies (3)(3) for a given cycle C(t)C(t), since, by Theorem 2.1, (3)(3) implies (1)(1) and (2)(2), and trivially (2)(2) implies (1)(1).

The permutation in the monodromy group associated with a critical value with a Morse point is a transposition. Let the cycle C(t)C(t) be given by (1.1).

First, we shall prove that there is a transposition τs\tau_{s} corresponding to a critical value ss and k{1,2,,m}k\in\{1,2,\ldots,m\} such that nknτs(k)n_{k}\neq n_{\tau_{s}(k)}. Note that otherwise, nk=nτs(k)n_{k}=n_{\tau_{s}(k)} for every k{1,2,,m}k\in\{1,2,\ldots,m\} and every critical value ss. Since the monodromy group is transitive and it is generated by the permutations corresponding to critical values, we would have nk=an_{k}=a for every k{1,2,,m}k\in\{1,2,\ldots,m\}, and certain aa\in\mathbb{Z}. Now, since C(t)C(t) is a cycle, 0=nk=ma0=\sum n_{k}=ma and the cycle C(t)C(t) would be trivial.

Take ss and k{1,2,,m}k\in\{1,2,\ldots,m\} such that nknτs(k)n_{k}\neq n_{\tau_{s}(k)}. By analytic continuation,

i=1mnτs(i)g(zi(t))0.\sum_{i=1}^{m}n_{\tau_{s}(i)}g(z_{i}(t))\equiv 0.

Therefore,

0=i=1mnig(zi(t))i=1mnτs(i)g(zi(t))=(nknτs(k))(g(zk(t))g(zτs(k)(t))),0=\sum_{i=1}^{m}n_{i}g(z_{i}(t))-\sum_{i=1}^{m}n_{\tau_{s}(i)}g(z_{i}(t))=\left(n_{k}-n_{\tau_{s}(k)}\right)\left(g(z_{k}(t))-g(z_{\tau_{s}(k)}(t))\right),

so the integral zk(t)zτs(k)(t)g\int_{z_{k}(t)-z_{\tau_{s}(k)}(t)}g vanishes. By Proposition 5.3 and Remark 5.4, there exists f0,g0,hf_{0},g_{0},h such that {k,τs(k)}Bh\{k,\tau_{s}(k)\}\subset B\in\mathcal{B}_{h}, f=f0hf=f_{0}\circ h and g=g0hg=g_{0}\circ h.

If τ=(i,j)Gf\tau=(i,j)\in G_{f} and \mathcal{B} is a non-trivial imprimitivity system of GfG_{f}, then {i,j}B\{i,j\}\subset B for some block BB\in\mathcal{B}. Since GfG_{f} is generated by the transpositions corresponding to the critical values and it is transitive, then for each {i,j}{1,,m}\{i,j\}\subset\{1,\ldots,m\}, there is a chain of transpositions (i,i1)(i,i_{1}), (i1,i2)(i_{1},i_{2}), \ldots, (ir,j)(i_{r},j). Therefore, there is only one block B={1,,m}B=\{1,\ldots,m\}, and the unique imprimitivity systems of GfG_{f} are the trivial ones. Thus, h\mathcal{B}_{h} must be the trivial imprimitivity system with a unique block and, in consequence, h=fh=f. ∎

If the polynomial ff is not generic, then there are some solutions gg which are not of the form g=g0fg=g_{0}\circ f. Moreover, there exists solutions that are not of the form g=g0hg=g_{0}\circ h with f=f0hf=f_{0}\circ h.

Example 5.6.

Let f(x)=x4x2f(x)=x^{4}-x^{2}. Let x1(t)<x2(t)<x3(t)<x4(t)x_{1}(t)<x_{2}(t)<x_{3}(t)<x_{4}(t), t(1/4,0)t\in(-1/4,0), be the real roots of ff. Consider the cycle

C(t)=x1(t)x2(t)x3(t)+x4(t).C(t)=x_{1}(t)-x_{2}(t)-x_{3}(t)+x_{4}(t).

Here tt can be analytically extended to {1/4,0}\mathbb{C}\setminus\{-1/4,0\}. Then for any odd function gg it follows C(t)g0\int_{C(t)}g\equiv 0, because f(x)=tf(x)=t is a biquadratic equation and therefore x1(t)=x4(t)x_{1}(t)=-x_{4}(t), x2(t)=x3(t)x_{2}(t)=-x_{3}(t). Take, for instance, g(x)=x3g(x)=x^{3}.

Let us show that there not exist f0,g0,hf_{0},g_{0},h with deg(h)>1deg(h)>1 such that f=f0hf=f_{0}\circ h, g=g0hg=g_{0}\circ h. Indeed, assume that there exist such decompositions and let h\mathcal{B}_{h} be the associated imprimitivity system, and BhB\in\mathcal{B}_{h}. Since the imprimitivity system is not trivial, then B={i,j}B=\{i,j\} and h(xi)=h(xj)h(x_{i})=h(x_{j}). On the other hand, gg does not admit decompositions, so g=hg=h, but g(xi)g(xj)g(x_{i})\neq g(x_{j}) for every iji\neq j.

6. Tangential center problem for polynomials

In this section we focus on the case when f,gf,g are polynomials. We shall prove Theorem 2.2 and some other results under the assumption that the cycle C(t)C(t) given by (1.1) is totally unbalanced.

Consider a cycle C(t)C(t) given by (1.1). It is represented by the vector n¯=(n1,,nm)\bar{n}=(n_{1},\ldots,n_{m}). Let σGf\sigma\in G_{f} be an element of the monodromy group of ff. It transforms the cycle C(t)C(t) into the cycle

(6.9) σ(C(t))=i=1mnizσ(i)(t)=i=1mnσ1(i)zi(t).\sigma(C(t))=\sum_{i=1}^{m}n_{i}z_{\sigma(i)}(t)=\sum_{i=1}^{m}n_{\sigma^{-1}(i)}z_{i}(t).

We consider the action of GfG_{f} on m\mathbb{C}^{m} defined by

σ(v)=(vσ(1),,vσ(m)),for every v=(v1,,vm)m,σGf.\sigma(v)=(v_{\sigma(1)},\ldots,v_{\sigma(m)}),\quad\text{for every }v=(v_{1},\ldots,v_{m})\in\mathbb{C}^{m},\ \sigma\in G_{f}.

Replacing σ1\sigma^{-1} by σ\sigma, we see that the orbit of the cycle C(t)C(t) by the action of the monodromy group GfG_{f} is represented by the space VmV\subset\mathbb{C}^{m} defined by

V=<σ(n¯):σGf>.V=<\sigma(\bar{n})\colon\sigma\in G_{f}>.

Let us choose τΓm(f)\tau\in\Gamma_{m}(f) and number the branches in such a way that τ=(1,2,,m)\tau=(1,2,\ldots,m).

Let \mathcal{B} be an imprimitivity system of GfG_{f}. By Proposition 4.1 it corresponds to a class of decompositions f=f0hf=f_{0}\circ h, where f0,hf_{0},h can be chosen polynomials (see Lemma 3.5 of [12]). Let d=deghd=\deg h. Then d|md|m and m/d=degf0m/d=\deg f_{0}. Since (1,2,,m)Gf(1,2,\ldots,m)\in G_{f}, then each BiB_{i}\in\mathcal{B}, i=1,,m/di=1,\ldots,m/d consists of the congruence class of ii modulo dd in {1,2,,m}\{1,2,\ldots,m\}. For each decomposition as above, the corresponding projected cycle is of the form

C~d(t)=1im/d(jdinj)wi(t),\tilde{C}_{d}(t)=\sum_{1\leq i\leq m/d}\left(\sum_{j\equiv_{d}i}n_{j}\right)w_{i}(t),

where wi(t)w_{i}(t) are all the different values of h(zj(t))h(z_{j}(t)).

Let D(f)D(f) denote the set of divisors rr of m=degfm=\deg f such that there is a decomposition f=fkhkf=f_{k}\circ h_{k} with deg(hk)=r\deg(h_{k})=r. Note that in D(f)D(f) we have the partial order induced by divisibility. We say that r1,,rlr_{1},\ldots,r_{l} is a is a complete set of elements of D(f)D(f) covered by rr if rk|rr_{k}|r and r1,r2,,rlr_{1},r_{2},\ldots,r_{l} are maximal among divisors of rr in D(f)\{r}D(f)\backslash\{r\}.

For every rD(f)r\in D(f), let us denote by VrV_{r} the set of rr-periodic vectors.

Lemma 6.1 ([28]).

Each GfG_{f}-irreducible subspace of m\mathbb{Q}^{m} has the form

Ur:=Vr(Vr1Vrl),U_{r}:=V_{r}\cap\left(V_{r_{1}}^{\bot}\cap\ldots\cap V_{r_{l}}^{\bot}\right),

where rD(f)r\in D(f) and r1,,rlr_{1},\ldots,r_{l} is a complete set of divisors of D(f)D(f) covered by rr. The subspaces UrU_{r} are mutually orthogonal and every GfG_{f}-invariant subspace of m\mathbb{Q}^{m} is a direct sum of some UrU_{r} as above.

Lemma 6.2 ([28]).

Let gg be a polynomial and let

g(z1(t))=kdeg(g)sktk/mg(z_{1}(t))=\sum_{k\geq-\deg(g)}s_{k}t^{-k/m}

denote the Puiseaux expansion at infinity of g(z1(t))g(z_{1}(t)), where z1(t)z_{1}(t) is one of the branches of the preimage of the polynomial ff. For any cD(f)c\in D(f), cmc\neq m, if we define

ψc(t)=km/c0,kdeg(g)sktk/m,\psi_{c}(t)=\sum_{k\equiv_{m/c}0,k\geq-\deg(g)}s_{k}t^{-k/m},

there exist w(z)[z]w(z)\in\mathbb{C}[z] such that

ψc(t)=w(z1(t)).\psi_{c}(t)=w(z_{1}(t)).

Moreover, there exist f0,g0,h[z]f_{0},g_{0},h\in\mathbb{C}[z], with deg(h)>1deg(h)>1, such that

f=f0h,w=g0h.f=f_{0}\circ h,\quad w=g_{0}\circ h.

Now we are in conditions of proving Theorem 2.2.

Proof of Theorem 2.2.

First we prove (2)(2).

Assume that there exist f1,,fs,g1,,gs,h1,,hs[z]f_{1},\ldots,f_{s},g_{1},\ldots,g_{s},h_{1},\ldots,h_{s}\in\mathbb{C}[z] such that f=fkhkf=f_{k}\circ h_{k}, k=1,,sk=1,\ldots,s, g=g1h1++gshsg=g_{1}\circ h_{1}+\ldots+g_{s}\circ h_{s}. Set dk=deg(hk)d_{k}=\deg(h_{k}) and let hk(C(t))h_{k}(C(t)) denote the projected cycle

hk(C(t))=1im/dk(jdkinj)wi(t),k=1,,s.h_{k}(C(t))=\sum_{1\leq i\leq m/d_{k}}\left(\sum_{j\equiv_{d_{k}}i}n_{j}\right)w_{i}(t),\quad k=1,\ldots,s.

Assume that for every kk, either the projected cycle is trivial or balanced, and

hk(C(t))gk0.\int_{h_{k}(C(t))}g_{k}\equiv 0.

Then by Proposition 5.1, gkhkg_{k}\circ h_{k} is a solution of (2.4) for every kk. By linearity, gg is also a solution of (2.4).


Now, assume that gg is a solution of (2.4). By analytic continuation this means that

(6.10) i=1mnσ(i)g(zi(t))0.\sum_{i=1}^{m}n_{\sigma(i)}g(z_{i}(t))\equiv 0.

Assume that C(t)C(t) is unbalanced. Thus, there exists a permutation cycle of order mm, τ=(p1,p2,,pm)Γm(f)\tau=(p_{1},p_{2},\ldots,p_{m})\in\Gamma_{m}(f), such that

i=1mnpiϵmi0.\sum_{i=1}^{m}n_{p_{i}}\epsilon_{m}^{i}\neq 0.

There is no loss of generality in assuming that τ=(1,2,,m)\tau=(1,2,\ldots,m). Consider the orbit of the cycle

V=<σ(n¯):σGf>.V=<\sigma(\bar{n})\colon\sigma\in G_{f}>.

Let us define the vectors

wk=(1,ϵmk,ϵm2k,,ϵm(m1)k)w_{k}=(1,\epsilon_{m}^{k},\epsilon_{m}^{2k},\ldots,\epsilon_{m}^{(m-1)k})

for every natural kk, where ϵm=e2πi/m\epsilon_{m}=e^{2\pi i/m}. Note that these vectors for k=1,,mk=1,\ldots,m form an orthogonal basis of m\mathbb{C}^{m}. Moreover, for every rD(f)r\in D(f), the (complexified) subspace VrV^{\mathbb{C}}_{r} is generated by the vectors wkw_{k} such that (m/r)|k(m/r)|k.

By Lemma 6.1, for r=mr=m we have that

W=Um=Vr1VrlW=U_{m}=V_{r_{1}}^{\bot}\cap\ldots\cap V_{r_{l}}^{\bot}

is a GfG_{f}-irreducible subspace of m\mathbb{Q}^{m}, where r1,,rlr_{1},\ldots,r_{l} are the maximal elements of D(f){m}D(f)\setminus\{m\}.

Let us prove that WVW\subset V. Since V,WV,W are invariant by the action of GfG_{f} and WW is irreducible, then either WVW\subset V or WVW\subset V^{\perp}. We show that w1Ww_{1}\in W^{\mathbb{C}} and w1Vw_{1}\not\in V^{\mathbb{C}\perp}, thus WVW^{\mathbb{C}}\not\subset V^{\mathbb{C}\perp}. Indeed, for every ii, VriV^{\mathbb{C}}_{r_{i}} is generated by the vectors wkw_{k} where (m/ri)|k(m/r_{i})|k. Since rimr_{i}\neq m, then k>1k>1 and these wkw_{k} are orthogonal to w1w_{1}. Therefore w1Ww_{1}\in W^{\mathbb{C}}, but since C(t)C(t) is unbalanced, w1w_{1} is not orthogonal to n¯V\bar{n}\in V. Then VV and WW are not orthogonal, thus showing that WVW\subset V.

Now, take the Puiseux expansions of g(zi(t))g(z_{i}(t)) at infinity. Applying Newton’s Polygon Method to f(z)=tf(z)=t (see, e.g., [10, p. 45]) zi(t)z_{i}(t) can be written as an analytic function in ϵmi1t1/m\epsilon_{m}^{i-1}t^{-1/m}, and composing with gg we get

g(zi(t))=k=deg(g)skϵm(i1)ktk/m.g(z_{i}(t))=\sum_{k=-\deg(g)}^{\infty}s_{k}\epsilon_{m}^{(i-1)k}t^{-k/m}.

Now, (6.10) yields

(6.11) ski=1mnσ(i)ϵm(i1)k=0,kdeg(g),σGf.s_{k}\sum_{i=1}^{m}n_{\sigma(i)}\epsilon_{m}^{(i-1)k}=0,\quad k\geq-\deg(g),\ \sigma\in G_{f}.

Assume that sk00s_{k_{0}}\neq 0 for some k0deg(g){k_{0}}\geq-\deg(g). Then wk0w_{k_{0}} is orthogonal to VV, and therefore, to WW. Since WW^{\mathbb{C}\perp} is generated by the vectors wkw_{k}, (m/r)|k(m/r)|k, rr maximal in D(f){m}D(f)\setminus\{m\}, this implies that wk0w_{k_{0}} is a linear combination of these vectors. But {wi}1,,m\{w_{i}\}_{1,\ldots,m} are linearly independent, therefore wk0w_{k_{0}} coincides with one of them. Consequently, (m/c)|k0(m/c)|k_{0} for some cc maximal element of D(f)\{m}D(f)\backslash\{m\}.

Let r=m/cr=m/c. Define

ψc(t)=kr0kdeg(g)sktk/m.\psi_{c}(t)=\underset{k\geq-\deg(g)}{\underset{k\equiv_{r}0}{\sum}}s_{k}t^{-k/m}.

By Lemma 6.2, there exist f0,g0,h,w[z]f_{0},g_{0},h,w\in\mathbb{C}[z], such that ψc=wz1\psi_{c}=w\circ z_{1} and f=f0hf=f_{0}\circ h, w=g0hw=g_{0}\circ h, with deg(h)>1\deg(h)>1.

The Puiseux expansion of w(zi(t))w(z_{i}(t)) at infinity is

w(zi(t))=kr0kdeg(g)skϵm(i1)ktk/m.w(z_{i}(t))=\underset{k\geq-\deg(g)}{\underset{k\equiv_{r}0}{\sum}}s_{k}\epsilon_{m}^{(i-1)k}t^{-k/m}.

Therefore, ww satisfies (6.11) and as a consequence

C(t)w=i=1mniw(zi(t))0.\int_{C(t)}w=\sum_{i=1}^{m}n_{i}w(z_{i}(t))\equiv 0.

Now, since w=g0hw=g_{0}\circ h, projecting the cycle C(t)C(t) by hh,

h(C(t))g0=1im/d(jdinj)g0(wi(t))0,\int_{h(C(t))}g_{0}=\sum_{1\leq i\leq m/d}\left(\sum_{j\equiv_{d}i}n_{j}\right)g_{0}(w_{i}(t))\equiv 0,

where wi(t)=h(zi)w_{i}(t)=h(z_{i}), and dd is the number of different wi(t)w_{i}(t).

Since gwg-w has sim/c=0s_{im/c}=0 for every ideg(g)i\geq-\deg(g), by induction on the number of maximal elements c¯\bar{c} of D(f)\{m}D(f)\backslash\{m\} such that sim/c¯0s_{im/\bar{c}}\neq 0 for some ii, we obtain

g=kgk,g=\sum_{k}g_{k},

where the Abelian integral of gkg_{k} along the cycle

Ck(t):=1im/dk(jdkinj)wj(t)C_{k}(t):=\sum_{1\leq i\leq m/d_{k}}\left(\sum_{j\equiv_{d_{k}}i}n_{j}\right)w_{j}(t)

vanishes for every kk. It only remains to prove that we can assume that either Ck(t)C_{k}(t) is trivial or balanced. If it is none of them, then it is unbalanced, and applying again the arguments above, gk=jgkjg_{k}=\sum_{j}g_{k_{j}}, where now the cycles associated with each gkjg_{k_{j}} are projections of the cycle Ck(t)C_{k}(t). Repeating this argument recursively, one obtains that either Ck(t)C_{k}(t) is trivial or balanced.

In particular, if C(t)C(t) is totally unbalanced, then every Ck(t)C_{k}(t) must be trivial, which proves (1)(1). ∎

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 f(z)f(z) be a polynomial of degree mm such that f1(t)f^{-1}(t) has mm different branches z1(t),,zm(t)z_{1}(t),\ldots,z_{m}(t) for every tt\in\mathbb{C}. Let C(t)C(t) be a cycle given by (1.1), which can be seen as a vector of the \mathbb{Q}-vector space E=z1zmE=\mathbb{Q}z_{1}\oplus\ldots\oplus\mathbb{Q}z_{m}. Note that given a vector of EE, we can always get a multiple with integer coordinates.

As before, let us choose τΓm(f)\tau\in\Gamma_{m}(f) and choose the numbering of the branches such that τ=(1,2,,m)\tau=(1,2,\ldots,m). Then, for every positive integer l|ml|m the candidates to imprimitivity systems of GfG_{f} are

l={Br}r=1,,d,\mathcal{B}_{l}=\{B_{r}\}_{r=1,\ldots,d},

where

Br={r,r+d,,r+(l1)d}B_{r}=\{r,r+d,\ldots,r+(l-1)d\}

for d=m/ld=m/l. By Proposition 4.1, if l\mathcal{B}_{l} is an imprimitivity system, we get a decomposition (up to an equivalence) f=f0hf=f_{0}\circ h. Then, by Remark 4.2 the projected cycle h(C(t))h(C(t)) is trivial if and only if jBinj=0\sum_{j\in B_{i}}n_{j}=0 for i=1,,di=1,\ldots,d. Similarly, we shall say that jBinj=0\sum_{j\in B_{i}}n_{j}=0 for i=1,,di=1,\ldots,d is the projection of the cycle C(t)C(t) by l\mathcal{B}_{l} even when l\mathcal{B}_{l} is not an imprimitivity system (note that if f(z)=zmf(z)=z^{m}, then l\mathcal{B}_{l} is always an imprimitivity system). These dd equations form a homogeneous linear system in mm unknowns njn_{j}. As every column in the associated matrix has just one 11 and the rest of the coefficients zero, while every row has at least one 11, we can assure that the rank of the system is dd. In particular, a non-trivial (integer) solution of the system is a cycle, since the sum of all equations is the condition for C(t)C(t) being a cycle.

Let us check the dimension of the \mathbb{Q}-vector space of the cycles with trivial projection for every l\mathcal{B}_{l}.

Proposition 6.3.

Under the above conditions, the dimension of the \mathbb{Q}-vector space TT of the cycles with trivial projection for every l\mathcal{B}_{l} is ϕ(m)\phi(m), where ϕ(m)\phi(m) is Euler’s totient function, defined as the number of positive integers less than mm that are relatively prime with mm.

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 \mathbb{Q}-vector subspaces from the \mathbb{Q}-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 p\mathcal{B}_{p}, where pp is a prime divisor of mm.

Let m=p1n1prnrm=p_{1}^{n_{1}}\cdot\ldots\cdot p_{r}^{n_{r}} be the decomposition of mm as product of primes p1,,prp_{1},\ldots,p_{r}. For every prime pip_{i} we get a homogeneous linear system of di=m/pid_{i}=m/p_{i} linearly independent equations in pip_{i} unknowns. Every one of those equations gives a did_{i}-periodic vector vkv^{k} of m\mathbb{Q}^{m}:

vjk={1 if jBk, i.e., if jdik0 otherwise v^{k}_{j}=\left\{\begin{matrix}1&\mbox{ if }j\in B_{k},\mbox{ i.e., if }j\equiv_{d_{i}}k\\ &\\ 0&\mbox{ otherwise }\end{matrix}\right.

So for every prime pi|mp_{i}|m we have did_{i} linearly independent vectors of VdiV_{d_{i}}, that is, we get a basis for VdiV_{d_{i}}.

The subspace TT of EE is the solution of the homogeneous linear system of d1++drd_{1}+\ldots+d_{r} equations we get for the primes p1,,prp_{1},\ldots,p_{r}. So, the codimension of TT is the rank of the matrix associated with this system of equations.

If we consider the equations associated with primes pip_{i} and pjp_{j}, we have a set of vectors which consists of a basis for VdiV_{d_{i}} and a basis for VdjV_{d_{j}}. Then, we get a system of generators for Vdi+VdjV_{d_{i}}+V_{d_{j}} and, in consequence, the rank of the matrix associated with those di+djd_{i}+d_{j} equations is the dimension of Vdi+VdjV_{d_{i}}+V_{d_{j}}. That is,

dim(Vdi+Vdj)=dimVdi+dimVdjdim(VdiVdj).\dim\,(V_{d_{i}}+V_{d_{j}})=\dim V_{d_{i}}+\dim V_{d_{j}}-\dim\,(V_{d_{i}}\cap V_{d_{j}}).

Now let us observe that VdiVdj=Vm/pipjV_{d_{i}}\cap V_{d_{j}}=V_{m/p_{i}p_{j}}, since g.c.d.(di,dj)=m/pipjg.c.d.(d_{i},d_{j})=m/p_{i}p_{j} because pip_{i} and pjp_{j} are relatively prime. Therefore,

dim(Vdi+Vdj)=di+djm/pipj.\dim\,(V_{d_{i}}+V_{d_{j}})=d_{i}+d_{j}-m/p_{i}p_{j}.

In the general case, when we consider all the equations we get:

codimT=dim(Vd1++Vdr)=i=1rdimVdi1i1<i2rdim(Vdi1Vdi2)+1i1<i2<i3rdim(Vdi1Vdi2Vdi3)++(1)r+1dim(Vd1Vdr)=i=1rmpi1i1<i2rmpi1pi2+1i1<i2<i3rmpi1pi2pi3++(1)r+1mp1pr=m(i=1r1pi1i1<i2r1pi1pi2+1i1<i2<i3r1pi1pi2pi3++(1)r+11p1pr)=m(1i=1r(11pi))=mmi=1r(11pi)=mϕ(m).\begin{split}{\rm codim}\,T=&\dim\,(V_{d_{1}}+\ldots+V_{d_{r}})=\sum_{i=1}^{r}\dim V_{d_{i}}-\sum_{1\leq i_{1}<i_{2}\leq r}\dim\,(V_{d_{i_{1}}}\cap V_{d_{i_{2}}})\\ &+\sum_{1\leq i_{1}<i_{2}<i_{3}\leq r}\dim\,(V_{d_{i_{1}}}\cap V_{d_{i_{2}}}\cap V_{d_{i_{3}}})+\ldots\\ &+(-1)^{r+1}\dim\,(V_{d_{1}}\cap\ldots\cap V_{d_{r}})\\ =&\sum_{i=1}^{r}\frac{m}{p_{i}}-\sum_{1\leq i_{1}<i_{2}\leq r}\frac{m}{p_{i_{1}}p_{i_{2}}}+\sum_{1\leq i_{1}<i_{2}<i_{3}\leq r}\frac{m}{p_{i_{1}}p_{i_{2}}p_{i_{3}}}\\ &+\ldots+(-1)^{r+1}\frac{m}{p_{1}\cdot\ldots\cdot p_{r}}=m\left(\sum_{i=1}^{r}\frac{1}{p_{i}}-\sum_{1\leq i_{1}<i_{2}\leq r}\frac{1}{p_{i_{1}}p_{i_{2}}}\right.\\ &\left.+\sum_{1\leq i_{1}<i_{2}<i_{3}\leq r}\frac{1}{p_{i_{1}}p_{i_{2}}p_{i_{3}}}+\ldots+(-1)^{r+1}\frac{1}{p_{1}\cdot\ldots\cdot p_{r}}\right)\\ =&\,m\left(1-\prod_{i=1}^{r}\left(1-\frac{1}{p_{i}}\right)\right)=m-m\prod_{i=1}^{r}\left(1-\frac{1}{p_{i}}\right)=m-\phi(m).\end{split}

Therefore, TT is a \mathbb{Q}-vector subspace of EE of dimension ϕ(m)\phi(m). Consequently, for every mm we have ϕ(m)\phi(m) (up to a multiple) linearly independent cycles with trivial projections. ∎

Example 6.5.

Given m=210=2357m=210=2\cdot 3\cdot 5\cdot 7, for the polynomial f(z)=z210f(z)=z^{210} we can find at least ϕ(210)1=47\phi(210)-1=47 linearly independent totally unbalanced cycles such that g(z)=z2+z3+z5+z7g(z)=z^{2}+z^{3}+z^{5}+z^{7} 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 m=p1psm=p_{1}\cdot\ldots\cdot p_{s}, for the polynomial f(z)=zmf(z)=z^{m} we can find at least ϕ(m)1\phi(m)-1 linearly independent totally unbalanced cycles such that g(z)=zp1+zp2++zpsg(z)=z^{p_{1}}+z^{p_{2}}+\ldots+z^{p_{s}} is a solution of (2.4), for ss 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 kLk\to L and an algebraic element αL\alpha\in L over kk, the minimal polynomial of α\alpha is the monic polynomial p(z)k[z]p(z)\in k[z] of the least degree such that p(α)=0p(\alpha)=0. The minimal polynomial is irreducible over kk and any other non-zero polynomial q(z)k[z]q(z)\in k[z] with q(α)=0q(\alpha)=0 is a multiple of p(z)p(z) in k[z]k[z].

The mm-th roots of unity, that is, the roots of the polynomial zm1k[z]z^{m}-1\in k[z] form an abelian cyclic group denoted by μm\mu_{m}. If the characteristic of kk is zero, then μm\mu_{m} has order mm and

μm={e2πi/m=ϵm,ϵm2,,ϵmm=1},\mu_{m}=\{e^{2\pi i/m}=\epsilon_{m},\epsilon_{m}^{2},\ldots,\epsilon_{m}^{m}=1\},

where ϵm\epsilon_{m} is called a primitive mm-th root of unity because it generates μm\mu_{m}. When k=k=\mathbb{Q}, then there are exactly ϕ(m)\phi(m) primitive mm-th roots of unity.

The mm-th cyclotomic polynomial is defined to be the minimal polynomial Φm(z)\Phi_{m}(z) that has the primitive mm-th roots of unity as simple roots:

Φm(z)=1k<m,gcd(k,m)=1(ze2πimk).\Phi_{m}(z)=\underset{\begin{subarray}{c}\quad 1\leq k<m,\\ gcd(k,m)=1\end{subarray}}{\prod}(z-e^{\frac{2\pi i}{m}k}).

Obviously, ϕ(m)\phi(m) is the degree of Φm(z)\Phi_{m}(z). Cyclotomic polynomials are irreducible polynomials with integer coefficients and Φm(z)\Phi_{m}(z) is the minimal polynomial of e2πi/me^{2\pi i/m} over \mathbb{Q}.

Given a divisor dd of mm, the primitive dd-th roots of unity are the elements of μm\mu_{m} of order dd. As the order of any element of μm\mu_{m} is a divisor of mm, we can conclude that

μm=d|m{ primitive d-th roots of unity }.\mu_{m}=\underset{d|m}{\coprod}\{\text{ primitive $d$-th roots of unity }\}.

Then

zm1=αμm(zα)=d|mΦd(z).z^{m}-1=\underset{\alpha\in\mu_{m}}{\prod}(z-\alpha)=\underset{d|m}{\prod}\Phi_{d}(z).

Given a cycle C(t)C(t) by (1.1) and a permutation cycle σ=(p1,,pm)Γm(f)Gf\sigma=(p_{1},\ldots,p_{m})\in\Gamma_{m}(f)\subset G_{f}, we associate the principal ideal

IC,σ=([np1z+np2z2++npmzm])[z]/(zm1).I_{C,\sigma}=([n_{p_{1}}z+n_{p_{2}}z^{2}+\ldots+n_{p_{m}}z^{m}])\subset\mathbb{C}[z]/(z^{m}-1).

The definition above does not depend on the way we represent σ\sigma. Indeed, if we write σ=(p2,p3,,pm,p1)\sigma=(p_{2},p_{3},\ldots,p_{m},p_{1}), then

np1z+np2z2++npmzm=(np2z+np3z2++npmzm1+np1zm)znp1z(zm1),\begin{split}n_{p_{1}}z+n_{p_{2}}z^{2}+\ldots+n_{p_{m}}z^{m}=&(n_{p_{2}}z+n_{p_{3}}z^{2}+\ldots+n_{p_{m}}z^{m-1}+n_{p_{1}}z^{m})z\\ &-n_{p_{1}}z(z^{m}-1),\end{split}

and consequently

[np1z+np2z2++npmzm]=[(np2z+np3z2++npmzm1+np1zm)][z],[n_{p_{1}}z+n_{p_{2}}z^{2}+\ldots+n_{p_{m}}z^{m}]=[(n_{p_{2}}z+n_{p_{3}}z^{2}+\ldots+n_{p_{m}}z^{m-1}+n_{p_{1}}z^{m})][z],

where [z][z] is invertible in [z]/(zm1)\mathbb{C}[z]/(z^{m}-1).

We will define PC,σ(z)P_{C,\sigma}(z) to be any generator of IC,σI_{C,\sigma}, that is,

(7.12) PC,σ(z)=[np1+np2z++npmzm1].P_{C,\sigma}(z)=[n_{p_{1}}+n_{p_{2}}z+\ldots+n_{p_{m}}z^{m-1}].

Note that the cycle C(t)C(t) is balanced if and only if ϵm\epsilon_{m} is a root of every polynomial of IC,σI_{C,\sigma} for every σΓm(f)\sigma\in\Gamma_{m}(f), which is equivalent to ϵm\epsilon_{m} being a root of PC,σ(z)P_{C,\sigma}(z) for every σΓm(f)\sigma\in\Gamma_{m}(f).

Proposition 7.1.

The cycle C(t)C(t) is balanced if and only if

PC,σ(z)=Φ1(z)Φm(z)dσ(z)σΓm(f),P_{C,\sigma}(z)=\Phi_{1}(z)\Phi_{m}(z)d_{\sigma}(z)\quad\forall\sigma\in\Gamma_{m}(f),

where Φm(z)\Phi_{m}(z) is the mm-th cyclotomic polynomial and dσ(z)[z]d_{\sigma}(z)\in\mathbb{C}[z] .

Proof.

Let C(t)=i=1mnizi(t)C(t)=\sum_{i=1}^{m}n_{i}z_{i}(t) be a cycle, σΓm(f)\sigma\in\Gamma_{m}(f) and PC,σ(z)P_{C,\sigma}(z) be the polynomial associated with C(t)C(t) and σ\sigma by (7.12). Since C(t)C(t) is a cycle, it follows that i=1mni=0\sum_{i=1}^{m}n_{i}=0, that is, 11 is a root of PC,σ(z)P_{C,\sigma}(z), thus giving that Φ1(z)=z1\Phi_{1}(z)=z-1 is a divisor of PC,σ(z)P_{C,\sigma}(z).

Now the cycle is balanced if and only if PC,σ(ϵm)=0P_{C,\sigma}(\epsilon_{m})=0 for every σΓm(f)\sigma\in\Gamma_{m}(f), that is, if and only if ϵm\epsilon_{m} is a root of the polynomial PC,σ(z)P_{C,\sigma}(z). As Φm(z)\Phi_{m}(z) is the minimal polynomial of ϵm\epsilon_{m}, it is equivalent to Φm(z)\Phi_{m}(z) dividing PC,σ(z)P_{C,\sigma}(z). The claim follows as Φ1(z)\Phi_{1}(z) and Φm(z)\Phi_{m}(z) are relatively prime. ∎

Corollary 7.2.

If mm is prime, then every cycle is totally unbalanced.

Proof.

First, if mm is prime, then Φm(z)=1+z++zm1\Phi_{m}(z)=1+z+\ldots+z^{m-1}. So it is not a factor of PC,σ(z)P_{C,\sigma}(z), and the cycle is unbalanced. Moreover, since mm is prime, then the unique imprimitivity systems are the trivial ones, and in consequence, C(t)C(t) 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 ff.

The first example is f(z)=zmf(z)=z^{m}, which can be completely solved. In this case PC,σ(z)P_{C,\sigma}(z) does not depend on σΓm(f)\sigma\in\Gamma_{m}(f), so to simplify we write PC(z)P_{C}(z).

Proposition 7.3.

Assume that f(z)=zmf(z)=z^{m} and that C(t)C(t) given by (1.1) is balanced. Then

g(z)=j=0bjzjg(z)=\sum_{j=0}^{\ell}b_{j}z^{j}

is a solution of (2.4) if and only if bj=0b_{j}=0 whenever the mk\frac{m}{k}-th cyclotomic polynomial Φm/k(z)\Phi_{m/k}(z) does not divide PC(z)P_{C}(z), where k=gcd(m,j)k=gcd(m,j), j>1j>1.

Proof.

The mm-th roots of f(z)=tf(z)=t are zi(t)=ϵmi1sz_{i}(t)=\epsilon_{m}^{i-1}s for i=1,,mi=1,\ldots,m, where ss is an mm-th root of tt and ϵm\epsilon_{m} is a primitive mm-th root of unity.

Let us compute the value of the expression C(t)g\int_{C(t)}g:

C(t)g=i=1mnig(zi)=i=1mni(j=0bjzi(t)j)=i=1mni(j=0bj(ϵmi1s)j)=j=0bjsj(i=1mniϵm(i1)j)==j=0bjsjPC(ϵmj).\begin{split}\int_{C(t)}g&=\sum_{i=1}^{m}n_{i}g(z_{i})=\sum_{i=1}^{m}n_{i}\left(\sum_{j=0}^{\ell}b_{j}z_{i}(t)^{j}\right)\\ &=\sum_{i=1}^{m}n_{i}\left(\sum_{j=0}^{\ell}b_{j}(\epsilon_{m}^{i-1}s)^{j}\right)=\sum_{j=0}^{\ell}b_{j}s^{j}\left(\sum_{i=1}^{m}n_{i}\epsilon_{m}^{(i-1)j}\right)=\\ &=\sum_{j=0}^{\ell}b_{j}s^{j}P_{C}(\epsilon_{m}^{j}).\end{split}

Hence, C(t)g0\int_{C(t)}g\equiv 0 if and only if bj=0b_{j}=0 whenever PC(ϵmj)0P_{C}(\epsilon_{m}^{j})\neq 0, j=0,,j=0,\ldots,\ell. For j=0j=0 and j=1j=1, PC(ϵmj)=0P_{C}(\epsilon_{m}^{j})=0 because C(t)C(t) is a cycle and because it is balanced respectively. Then, we can suppose j>1j>1. For j|mj|m, ϵmj\epsilon_{m}^{j} is a primitive mj\frac{m}{j}-th root of unity. As we saw in Proposition 7.1, ϵn\epsilon_{n} is a root of PC(z)P_{C}(z) if and only if Φn(z)\Phi_{n}(z) divides PC(z)P_{C}(z). Now PC(ϵmj)=PC(ϵm/kj/k)0P_{C}(\epsilon_{m}^{j})=P_{C}(\epsilon_{m/k}^{j/k})\neq 0 is equivalent to the condition that the mk\frac{m}{k}-th cyclotomic polynomial Φm/k(z)\Phi_{m/k}(z) does not divide PC(z)P_{C}(z), where k=gcd(m,j)k=gcd(m,j). ∎

Example 7.4.

Let us study some degree four examples. First note that since Φ1(z)Φ4(z)\Phi_{1}(z)\Phi_{4}(z) =(z1)(z2+1)=z3z2+z1=(z-1)(z^{2}+1)=z^{3}-z^{2}+z-1 has degree three, by Proposition 7.1, PC(z)=a(z3z2+z1)P_{C}(z)=a(z^{3}-z^{2}+z-1), a0a\neq 0. Then, the only candidates for balanced cycles are

C(t)=a(z1(t)z2(t)+z3(t)z4(t)),a0.C(t)=a\left(z_{1}(t)-z_{2}(t)+z_{3}(t)-z_{4}(t)\right),\quad a\neq 0.

Take f(z)=z4f(z)=z^{4}. Then, by Proposition 7.3, gg is a solution if and only if it is of the form

g(z)=g0(z4)+zg1(z4)+z3g3(z4),g(z)=g_{0}(z^{4})+zg_{1}(z^{4})+z^{3}g_{3}(z^{4}),

with g0,g1,g3g_{0},g_{1},g_{3} polynomials. Indeed, the only condition corresponds to coefficients bjb_{j}, with j42j\equiv_{4}2. Then k=gcd(4,j)=2k=gcd(4,j)=2, Φ2(z)=z+1\Phi_{2}(z)=z+1 does not divide PC(z)P_{C}(z) and hence there are no terms of powers 42\equiv_{4}2 in gg.

This gives examples of solutions gg of (2.4) which can not be written as a sum of gkg_{k} such that f=f0hf=f_{0}\circ h and gk=gk,0hg_{k}=g_{k,0}\circ h. For instance, g(z)=z3g(z)=z^{3} is a solution of (2.4), while the unique non-trivial decomposition of z4z^{4} (up to Moebius transformations) is z4=(z2)2z^{4}=\left(z^{2}\right)^{2}.

Example 7.5.

Now let f(z)=T4(z)=8z48z2+1f(z)=T_{4}(z)=8z^{4}-8z^{2}+1 be the fourth Chebyshev polynomial and again assume that C(t)C(t) is a balanced cycle, thus,

C(t)=a(z1(t)z2(t)+z3(t)z4(t)),a0.C(t)=a\left(z_{1}(t)-z_{2}(t)+z_{3}(t)-z_{4}(t)\right),\quad a\neq 0.

By direct computation,

z1(t)=1/2+1+t/(22),z2(t)=1/21+t/(22),z3(t)=1/2+1+t/(22),z4(t)=1/21+t/(22).\begin{split}z_{1}(t)=\sqrt{1/2+\sqrt{1+t}/(2\sqrt{2})},\quad&z_{2}(t)=\sqrt{1/2-\sqrt{1+t}/(2\sqrt{2})},\\ z_{3}(t)=-\sqrt{1/2+\sqrt{1+t}/(2\sqrt{2})},\quad&z_{4}(t)=-\sqrt{1/2-\sqrt{1+t}/(2\sqrt{2})}.\end{split}

Rewrite g=i=03aizigi(T4(z))g=\sum_{i=0}^{3}a_{i}z^{i}g_{i}(T_{4}(z)), for certain g0,g1,g2,g3[z]g_{0},g_{1},g_{2},g_{3}\in\mathbb{C}[z]. Then the Abelian integral along the cycle C(t)C(t) reduces to

k=14nkg(zk(t))=a2a2+2tg2(t).\sum_{k=1}^{4}n_{k}g(z_{k}(t))=a_{2}a\sqrt{2+2t}g_{2}(t).

Therefore gg is a solution if and only if there exist polynomials g0,g1,g3g_{0},g_{1},g_{3} such that

g(z)=g0(T4(z))+zg1(T4(z))+z3g3(T4(z)).g(z)=g_{0}(T_{4}(z))+zg_{1}(T_{4}(z))+z^{3}g_{3}(T_{4}(z)).

Observe that the weak composition conjecture does not hold in this case: take, for instance, the solution g(z)=z3g(z)=z^{3}, which does not share any non-trivial compositional factor with T4(z)=T2(T2(z))T_{4}(z)=T_{2}(T_{2}(z)).

Example 7.6.

Consider now f(z)=z6f(z)=z^{6} and a balanced cycle C(t)C(t). Note that this implies that the associated polynomial PC(z)P_{C}(z) must have the factors Φ1(z)\Phi_{1}(z) and Φ6(z)\Phi_{6}(z). By Proposition 7.3, the solutions depend on which cyclotomic polynomials divide PC(z)P_{C}(z). The unique candidates are Φ2(z)\Phi_{2}(z) and Φ3(z)\Phi_{3}(z). Moreover, it is not possible that both Φ2(z)\Phi_{2}(z), Φ3(z)\Phi_{3}(z) divide PC(z)P_{C}(z), since z61=Φ1(z)Φ2(z)Φ3(z)Φ6(z)z^{6}-1=\Phi_{1}(z)\Phi_{2}(z)\Phi_{3}(z)\Phi_{6}(z). Therefore there are three possibilities: neither Φ2(z)\Phi_{2}(z) nor Φ3(z)\Phi_{3}(z) divide PC(z)P_{C}(z) (e.g., C=C1C=C_{1}), Φ2(z)\Phi_{2}(z) divide PC(z)P_{C}(z) (e.g., C=C2C=C_{2}), or Φ3(z)\Phi_{3}(z) divide PC(z)P_{C}(z) (e.g., C=C3C=C_{3}).

\psfrag{n1}{$-1$}\psfrag{b1}{$\hskip-5.69046ptz_{1}(t)$}\psfrag{n2}{$2$}\psfrag{b2}{$\hskip-2.84544ptz_{2}(t)$}\psfrag{n3}{$-2$}\psfrag{b3}{$z_{3}(t)$}\psfrag{n4}{$\hskip-5.69046pt\phantom{-}1$}\psfrag{b4}{$z_{4}(t)$}\psfrag{n5}{$0$}\psfrag{b5}{$z_{5}(t)$}\psfrag{n6}{$0$}\psfrag{b6}{$\hskip-2.84544ptz_{6}(t)$}\includegraphics[height=85.35826pt]{base.eps}
(a) Cycle C1(t)C_{1}(t)
\psfrag{n1}{$2$}\psfrag{b1}{$\hskip-5.69046ptz_{1}(t)$}\psfrag{n2}{$-1$}\psfrag{b2}{$\hskip-2.84544ptz_{2}(t)$}\psfrag{n3}{$\hskip-2.84544pt-1$}\psfrag{b3}{$z_{3}(t)$}\psfrag{n4}{$\hskip 5.69046pt2$}\psfrag{b4}{$z_{4}(t)$}\psfrag{n5}{$\hskip-2.84544pt-1$}\psfrag{b5}{$z_{5}(t)$}\psfrag{n6}{$-1$}\psfrag{b6}{$\hskip-2.84544ptz_{6}(t)$}\includegraphics[height=85.35826pt]{base2.eps}
(b) Cycle C2(t)C_{2}(t)
\psfrag{n1}{$1$}\psfrag{b1}{$\hskip-5.69046ptz_{1}(t)$}\psfrag{n2}{$-1$}\psfrag{b2}{$\hskip-2.84544ptz_{2}(t)$}\psfrag{n3}{$1$}\psfrag{b3}{$z_{3}(t)$}\psfrag{n4}{$\hskip-5.69046pt-1$}\psfrag{b4}{$z_{4}(t)$}\psfrag{n5}{$1$}\psfrag{b5}{$z_{5}(t)$}\psfrag{n6}{$-1$}\psfrag{b6}{$\hskip-2.84544ptz_{6}(t)$}\includegraphics[height=85.35826pt]{base.eps}
(c) Cycle C3(t)C_{3}(t)

     

Figure 1. Some balanced cycles with six points

Now, Proposition 7.3 gives us all the solutions:

  1. (1)

    For the cycle

    C1(t)=z1(t)+2z2(t)2z3(t)+z4(t),C_{1}(t)=-z_{1}(t)+2z_{2}(t)-2z_{3}(t)+z_{4}(t),

    g(z)=k0bkzkg(z)=\sum_{k\geq 0}b_{k}z^{k} satisfies C1(t)g0\int_{C_{1}(t)}g\equiv 0 if and only if bk=0b_{k}=0 for every k62,3,4k\equiv_{6}2,3,4.


  2. (2)

    For the cycle

    C2(t)=2z1(t)z2(t)z3(t)+2z4(t)z5(t)z6(t),C_{2}(t)=2z_{1}(t)-z_{2}(t)-z_{3}(t)+2z_{4}(t)-z_{5}(t)-z_{6}(t),

    g(z)=k0bkzkg(z)=\sum_{k\geq 0}b_{k}z^{k} satisfies C1(t)g0\int_{C_{1}(t)}g\equiv 0 if and only if bk=0b_{k}=0 for every k62,4k\equiv_{6}2,4.

  3. (3)

    For the cycle

    C3(t)=z1(t)z2(t)+z3(t)z4(t)+z5(t)z6(t),C_{3}(t)=z_{1}(t)-z_{2}(t)+z_{3}(t)-z_{4}(t)+z_{5}(t)-z_{6}(t),

    g(z)=k0bkzkg(z)=\sum_{k\geq 0}b_{k}z^{k} satisfies C1(t)g0\int_{C_{1}(t)}g\equiv 0 if and only if bk=0b_{k}=0 for every k63k\equiv_{6}3.

Example 7.7.

To end with this section, let us show an example of an unbalanced cycle which is not totally unbalanced.

Take f(z)=(z2+az+b)6f(z)=(z^{2}+az+b)^{6}, with a,ba,b\in\mathbb{C}. Let τ\tau 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 τ=(1,2,,12)\tau=(1,2,\ldots,12).

Consider the cycle C(t)=2z1(t)z2(t)z3(t)+2z4(t)z5(t)z6(t)C(t)=2z_{1}(t)-z_{2}(t)-z_{3}(t)+2z_{4}(t)-z_{5}(t)-z_{6}(t). It is easy to check that this cycle is unbalanced, and that the non-trivial decompositions of ff are f(z)=z2(z2+az+b)3f(z)=z^{2}\circ(z^{2}+az+b)^{3}, f(z)=z3(z2+az+b)2f(z)=z^{3}\circ(z^{2}+az+b)^{2}, and f(z)=z6(z2+az+b)f(z)=z^{6}\circ(z^{2}+az+b).

Let us denote

h1=(z2+az+b)3,h2=(z2+az+b)2,h3=z2+az+b.h_{1}=(z^{2}+az+b)^{3},\quad h_{2}=(z^{2}+az+b)^{2},\quad h_{3}=z^{2}+az+b.

Then h1(C(t))h_{1}(C(t)) is trivial, h2(C(t))=4w¯1(t)2w¯2(t)2w¯3(t)h_{2}(C(t))=4\bar{w}_{1}(t)-2\bar{w}_{2}(t)-2\bar{w}_{3}(t) is unbalanced, and h3(C(t))=2w1(t)w2(t)w3(t)+2w4(t)w5(t)w6(t)h_{3}(C(t))=2w_{1}(t)-w_{2}(t)-w_{3}(t)+2w_{4}(t)-w_{5}(t)-w_{6}(t) is balanced.

By Theorem 2.2, for any g[z]g\in\mathbb{C}[z], Cg0\int_{C}g\equiv 0 iff there exist g1,g3[z]g_{1},g_{3}\in\mathbb{C}[z] such that g=g1h1+g3h3g=g_{1}\circ h_{1}+g_{3}\circ h_{3}, and

h3(C(t))g30.\int_{h_{3}(C(t))}g_{3}\equiv 0.

By Example 7.6, this is equivalent to g3(z)=k0bkzkg_{3}(z)=\sum_{k\geq 0}b_{k}z^{k} and bk=0b_{k}=0 for every k62,4k\equiv_{6}2,4.

8. Moment Problem

Let us consider the following Abel equation

(8.13) z=p(w)z2+ϵq(w)z3,z^{\prime}=p(w)z^{2}+\epsilon q(w)z^{3},

where pp, qq are analytic functions. We say that a solution z(w)z(w) of (8.13) is closed if z(0)=z(1)z(0)=z(1). Let us denote

f(w)=0wp(t)𝑑t,g(w)=0wq(t)𝑑t.f(w)=\int_{0}^{w}p(t)\,dt,\quad g(w)=\int_{0}^{w}q(t)\,dt.

Assume that we have a center at the origin for ϵ=0\epsilon=0 (thus, every bounded solution of the system is closed), which happens iff f(0)=f(1)f(0)=f(1), and we want to study when the center persists after perturbation (i.e., for every ϵ\epsilon 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 p,qp,q trigonometric polynomials. The case when p,q[w]p,q\in\mathbb{C}[w] 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 ϵ\epsilon, the center is persistent if and only if (z/ϵ)(1,z0,0)=0(\partial z/\partial\epsilon)(1,z_{0},0)=0 for every z0z_{0} close to zero, where z(t,z0,ϵ)z(t,z_{0},\epsilon) is the solution of (8.13) determined by z(0,z0,ϵ)=z0z(0,z_{0},\epsilon)=z_{0}.

Differentiating z1(t,z0,ϵ)z^{-1}(t,z_{0},\epsilon) with respect to ϵ\epsilon and evaluating at t=1t=1, ϵ=0\epsilon=0, one has that (z/ϵ)(1,z0,0)=0(\partial z/\partial\epsilon)(1,z_{0},0)=0 for every z0z_{0} close to zero if and only if

(8.14) 0=01q(w)dw1z0f(w)=k=0(01fk(w)q(w)𝑑w)z0k,0=\int_{0}^{1}\frac{q(w)dw}{1-z_{0}f(w)}=\sum_{k=0}^{\infty}\left(\int_{0}^{1}f^{k}(w)q(w)dw\right)z_{0}^{k},

for every z0z_{0} close to zero (see [6]).

The moment problem [5] asks for a given pp, to find all qq 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 pp and qq are polynomials, recently solved by Pakovich and Muzychuk [28]. More precisely, they prove:

Theorem 8.1 ([28]).

Take a polynomial pp such that f(0)=f(1)f(0)=f(1). A polynomial qq is a solution of the polynomial moment problem if and only if there exist polynomials f1,,fr,g1,,gr,h1,hrf_{1},\ldots,f_{r},g_{1},\ldots,g_{r},h_{1},\ldots h_{r} such that hk(0)=hk(1)h_{k}(0)=h_{k}(1), f=fkhkf=f_{k}\circ h_{k} for every kk and

g=k=1rgkhk.g=\sum_{k=1}^{r}g_{k}\circ h_{k}.

We shall show that the polynomial moment problem is equivalent to a zero-dimensional tangential center problem for ff and gg defined as in Theorem 8.1, and for the cycle

Cf(t)=i=1n0n1zai(t)i=1n1n0zbi(t),C_{f}(t)=\sum_{i=1}^{n_{0}}n_{1}z_{a_{i}}(t)-\sum_{i=1}^{n_{1}}n_{0}z_{b_{i}}(t),

where zai(t)z_{a_{i}}(t) are all the solutions of f(zai(t))=tf(z_{a_{i}}(t))=t close to 0 for tt close to f(0)f(0), and analogously for zbi(t)z_{b_{i}}(t) and 11. We shall prove that Cf(t)C_{f}(t) 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 p,qp,q such that f(0)=f(1)f(0)=f(1), the polynomial qq is a solution of (8.14) if and only if

Cf(t)g0.\int_{C_{f}(t)}g\equiv 0.

Moreover, Cf(t)C_{f}(t) is totally unbalanced.

Remark 8.3.

As a consequence of Proposition 8.2, if Cf(t)g0\int_{C_{f}(t)}g\equiv 0, then gg is totally determined by Theorem 2.2.

Proof.

By Theorem 1.1 of [28], qq is a solution of the polynomial moment problem if and only if there exists polynomials f1,,fr,g1,,gr,h1,hrf_{1},\ldots,f_{r},g_{1},\ldots,g_{r},h_{1},\ldots h_{r} such that hk(0)=hk(1)h_{k}(0)=h_{k}(1), f=fkhkf=f_{k}\circ h_{k} for every kk and

g=k=1rgkhk.g=\sum_{k=1}^{r}g_{k}\circ h_{k}.

On the other hand, if Cf(t)C_{f}(t) is totally unbalanced, then Theorem 2.2 implies that gg is a solution of (8.14) for the cycle Cf(t)C_{f}(t) if and only if there exist polynomials f1,,fs,g1,,gs,h1,hsf_{1},\ldots,f_{s},g_{1},\ldots,g_{s},h_{1},\ldots h_{s} such that f=fkhkf=f_{k}\circ h_{k} for every kk,

g=k=1rgkhk,g=\sum_{k=1}^{r}g_{k}\circ h_{k},

and hk(Cf(t))h_{k}(C_{f}(t)) is trivial.

First, let us prove that hk(0)=hk(1)h_{k}(0)=h_{k}(1) is equivalent to hk(Cf(t))h_{k}(C_{f}(t)) being trivial. To this end, let \mathcal{B} be the imprimitivity system corresponding to some hkh_{k} such that f=fkhkf=f_{k}\circ h_{k}.

If hk(0)=hk(1)h_{k}(0)=h_{k}(1), then the blocks of \mathcal{B} contain elements of both {ai}\{a_{i}\} and {bi}\{b_{i}\}: take tt a non-critical value of hkh_{k} and joint it to hk(0)=hk(1)h_{k}(0)=h_{k}(1) by an arc. Then hk1h_{k}^{-1} maps this arc into some arcs close to 0 and others close to 11. Let σGf\sigma\in G_{f} denote an element of the monodromy group corresponding to a cycle around hk(0)=hk(1)h_{k}(0)=h_{k}(1). Reorder ai,bia_{i},b_{i} so that σ\sigma permutes cyclically a1,,an0a_{1},\ldots,a_{n_{0}} and b1,,bn1b_{1},\ldots,b_{n_{1}} and that a1,b1a_{1},b_{1} belongs to the same block of \mathcal{B}. Now, the blocks reduced to a1,,an0a_{1},\ldots,a_{n_{0}}, b1,,bn1b_{1},\ldots,b_{n_{1}} consist of the congruence classes modulo dd. Therefore, for every block BB\in\mathcal{B},

kBnk=i=1n0/dn1i=1n1/dn0=0.\sum_{k\in B}n_{k}=\sum_{i=1}^{n_{0}/d}n_{1}-\sum_{i=1}^{n_{1}/d}n_{0}=0.

Conversely, if hk(0)hk(1)h_{k}(0)\neq h_{k}(1), since the blocks contain branches where the value of hkh_{k} is the same, then every block BB of the imprimitivity system \mathcal{B} corresponding to hkh_{k} either contains elements of {ai}\{a_{i}\} or elements of {bi}\{b_{i}\}, and then kBnk0\sum_{k\in B}n_{k}\neq 0, and hk(Cf(t))h_{k}(C_{f}(t)) is not trivial.

Therefore, to conclude it only remains to prove that Cf(t)C_{f}(t) is totally unbalanced. Indeed, let us choose τ\tau a permutation corresponding to a simple loop around infinity and reorder the numbering of the branches such that τ=(1,2,,m)\tau=(1,2,\ldots,m). Define

V(0)={ϵma1,,ϵman0},V(1)={ϵmb1,,ϵmbn1}.V(0)=\{\epsilon_{m}^{a_{1}},\ldots,\epsilon_{m}^{a_{n_{0}}}\},\quad V(1)=\{\epsilon_{m}^{b_{1}},\ldots,\epsilon_{m}^{b_{n_{1}}}\}.

The Monodromy Lemma in [26] states that the convex hulls of the sets V(0)V(0) and V(1)V(1) are disjointed when f(0)=f(1)f(0)=f(1), thus, \mathbb{C} can be divided into two half-planes P1,P2P^{1},P^{2} such that V(0)P1V(0)\in P^{1} and V(1)P2V(1)\in P^{2}.

Now, the center of mass of Cf(t)C_{f}(t) is the origin (i.e., PCf,τ(ϵm)=0P_{C_{f},\tau}(\epsilon_{m})=0) if and only if the center of mass of V(0)V(0) is equal to the center of mass of V(1)V(1). But the center of mass of V(0)V(0) belongs to P1P^{1} and the center of mass of V(1)V(1) belongs to P2P^{2}. Therefore, Cf(t)C_{f}(t) is unbalanced.

If f=fkhkf=f_{k}\circ h_{k}, then there are two possibilities for the projected cycle hk(Cf(t))h_{k}(C_{f}(t)): If hk(0)=hk(1)h_{k}(0)=h_{k}(1), then hk(Cf(t))h_{k}(C_{f}(t)) is trivial. If hk(0)hk(1)h_{k}(0)\neq h_{k}(1), then we shall prove that hk(Cf(t))h_{k}(C_{f}(t)) is unbalanced. Indeed, the permutation induced on the branches za1,za2,,z_{a_{1}},z_{a_{2}},\ldots, zan0z_{a_{n_{0}}} by the preimage of a loop around f(0)f(0) is a permutation cycle. Therefore, the permutation induced on the branches hk(za1),hk(za2),hk(zan0)h_{k}(z_{a_{1}}),h_{k}(z_{a_{2}}),\ldots h_{k}(z_{a_{n_{0}}}) must also be a permutation cycle. Let us denote by dad_{a} the order of that permutation cycle and by dbd_{b} the order of the permutation cycle induced by a loop around f(1)f(1) on hk(zb1),hk(zb2),hk(zbn1)h_{k}(z_{b_{1}}),h_{k}(z_{b_{2}}),\ldots h_{k}(z_{b_{n_{1}}}). Then

hk(Cf(t))=j=1n0/dadan1hk(zaj(t))j=1n1/dbdbn0hk(zbj(t)).h_{k}(C_{f}(t))=\sum_{j=1}^{n_{0}/d_{a}}d_{a}n_{1}h_{k}(z_{a_{j}}(t))-\sum_{j=1}^{n_{1}/d_{b}}d_{b}n_{0}h_{k}(z_{b_{j}}(t)).

Applying the same arguments as for Cf(t)C_{f}(t), one obtains that hk(Cf(t))h_{k}(C_{f}(t)) is unbalanced. As a consequence, Cf(t)C_{f}(t) is totally unbalanced with respect to ff. ∎

Now, consider the trigonometric version of the moment problem. Given a trigonometric polynomial ff, find all trigonometric polynomials gg such that

(8.15) 02πfk(w)g(w)𝑑w=0,k0,\int_{0}^{2\pi}f^{k}(w)g^{\prime}(w)dw=0,\quad k\geq 0,

By the change of variable zeizz\to e^{iz}, the trigonometric moment problem is equivalent to the following problem (see [1]): Given a Laurent polynomial ff (f[z,z1]f\in\mathbb{C}[z,z^{-1}]), find all g[z,z1]g\in\mathbb{C}[z,z^{-1}] such that

(8.16) |z|=1fk(w)g(w)𝑑w=0,k0.\int_{|z|=1}f^{k}(w)g^{\prime}(w)dw=0,\quad k\geq 0.

Moreover, if |z|=1g(w)𝑑w=0\int_{|z|=1}g^{\prime}(w)\,dw=0, then by integration by parts,

|z|=1fk(w)g(w)𝑑w=k|z|=1fk1(w)f(w)g(w)𝑑w,k1.\int_{|z|=1}f^{k}(w)g^{\prime}(w)dw=-k\int_{|z|=1}f^{k-1}(w)f^{\prime}(w)g(w)dw,\quad k\geq 1.

Since |z|=1fk(w)f(w)g(w)𝑑w\int_{|z|=1}f^{k}(w)f^{\prime}(w)g(w)dw, k0k\geq 0, are the coefficients of the power series at t=t=\infty of

I(t)=|z|=1g(w)f(w)dwtf(w).I(t)=\int_{|z|=1}\frac{g(w)f^{\prime}(w)\,dw}{t-f(w)}.

then for any fixed Laurent polynomial ff, a Laurent polynomial gg such that |z|=1g(w)𝑑w=0\int_{|z|=1}g^{\prime}(w)\,dw=0 is a solution of (8.16) if and only if I(t)0I(t)\equiv 0.

Proposition 8.4.

Let

f(z)=k=nmckzkf(z)=\sum_{k=-n}^{m}c_{k}z^{k}

be a proper Laurent polynomial, that is, n,m1n,m\geq 1. A Laurent polynomial gg such that |z|=1g(w)𝑑w=0\int_{|z|=1}g^{\prime}(w)\,dw=0 is a solution of (8.16) if and only if gg is a solution of (2.4) for the cycle

Cf(t)=k=1nmzak(t)k=n+1n+mnzbk(t),C_{f}(t)=\sum_{k=1}^{n}mz_{a_{k}}(t)-\sum_{k=n+1}^{n+m}nz_{b_{k}}(t),

where zai(t)z_{a_{i}}(t) (resp. zbi(t)z_{b_{i}}(t)) are the branches close to 0 (resp. \infty) of a tt close to infinity.

Proof.

Take tt close to infinity. By the Residue Theorem,

I(t)=2πi(k=1ng(zak(t))+Res(gf/(tf),0))).I(t)=2\pi i\left(-\sum_{k=1}^{n}g(z_{a_{k}}(t))+Res(gf^{\prime}/(t-f),0))\right).

Assume that I(t)0I(t)\equiv 0 and take σGf\sigma\in G_{f}. By analytic continuation,

k=1ng(zak(t))k=1ng(zσ(ak)(t))=0\sum_{k=1}^{n}g(z_{a_{k}}(t))-\sum_{k=1}^{n}g(z_{\sigma(a_{k})}(t))=0

Since GfG_{f} is transitive, if we sum the previous formula over GfG_{f}, we obtain

k=1n|Gf|g(zak(t))k=1n+m|Gf|nn+mg(zk(t))=0.\sum_{k=1}^{n}|G_{f}|g(z_{a_{k}}(t))-\sum_{k=1}^{n+m}\frac{|G_{f}|n}{n+m}g(z_{k}(t))=0.

Dividing by |Gf||G_{f}| and multiplying by n+mn+m gives

Cf(t)g=k=1nmg(zak(t))k=n+11n+mng(zbk(t))=0.\int_{C_{f}(t)}g=\sum_{k=1}^{n}mg(z_{a_{k}}(t))-\sum_{k=n+11}^{n+m}ng(z_{b_{k}}(t))=0.

Conversely, if gg is a solution of (2.4), then using the Residues Theorem in the two regions of the complementary of the unit circle,

I(t)=2πin+m(mRes(gf/(tf),0))nRes(gf/(tf),))).I(t)=\frac{2\pi i}{n+m}(mRes(gf^{\prime}/(t-f),0))-nRes(gf^{\prime}/(t-f),\infty))).

Now, take α=1(tznf(z)zn)/cn\alpha=1-(tz^{n}-f(z)z^{n})/c_{n}. Note that α0\alpha\to 0 when z0z\to 0. Then for zz close to 0,

1tf(z)=zncn(1α)=zncnk=0αk=zncnk=0(1tznf(z)zncn)k.\frac{1}{t-f(z)}=\frac{z^{n}}{c_{n}(1-\alpha)}=\frac{z^{n}}{c_{n}}\sum_{k=0}^{\infty}\alpha^{k}=\frac{z^{n}}{c_{n}}\sum_{k=0}^{\infty}\left(1-\frac{tz^{n}-f(z)z^{n}}{c_{n}}\right)^{k}.

Therefore, Res(gf/(tf),0)Res(gf^{\prime}/(t-f),0) is a polynomial in tt, and the same holds for Res(gf/(tf),)Res(gf^{\prime}/(t-f),\infty). To conclude, note that I(t)0I(t)\to 0 as tt\to\infty, then I(t)0I(t)\equiv 0. ∎

In the general rational case, Pakovich [25] has proved that if γ\gamma is a rational curve and f,gf,g are rational functions, then

γfk(z)𝑑g(z)=0,for every k0,\int_{\gamma}f^{k}(z)\,dg(z)=0,\quad\text{for every }k\geq 0,

if and only if Cf,k(t)g0\int_{C_{f,k}(t)}g\equiv 0 for a finite number of cycles Cf,k(t)C_{f,k}(t) defined in terms of the the “dessin d’enfants” of ff.

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 F(x,y)[x,y]F(x,y)\in\mathbb{C}[x,y] and consider the foliation

(9.17) dF+ϵω=0dF+\epsilon\omega=0

deforming the initial foliation F=tF=t, where ω\omega is a one-form. Let t0t_{0} be a regular value of FF and let γ(t0)F1(t0)\gamma(t_{0})\subset F^{-1}(t_{0}) be a closed path in the leaf F1(t0)F^{-1}(t_{0}). Take a transversal TT to the leaves of FF parameterized by values tt of FF and consider the holonomy along γ(t0)\gamma(t_{0}) and the corresponding displacement map Δ(t,ϵ)\Delta(t,\epsilon) (holonomy minus identity). As the path γ(t0)\gamma(t_{0}) pushes to nearby closed paths γ(t)F1(t)\gamma(t)\subset F^{-1}(t), we have that Δ(t,0)0\Delta(t,0)\equiv 0. We say that the family γ(t)\gamma(t) is a center. We have

(9.18) Δ(t,ϵ)=ϵγ(t)ω+o(ϵ).\Delta(t,\epsilon)=-\epsilon\int_{\gamma(t)}\omega+o(\epsilon).

Of course the Abelian integral in (9.18) depends only on the homology class of γ(t)H1(F1(t))\gamma(t)\in H_{1}(F^{-1}(t)). We say that the family γ(t)\gamma(t) is a persistent center if Δ(t,ϵ)0\Delta(t,\epsilon)\equiv 0 and a tangential center if γ(t)ω0\int_{\gamma(t)}\omega\equiv 0.

We consider more precisely the hyper-elliptic case F(x,y)=y2+f(x)F(x,y)=y^{2}+f(x), f(x)(x),f(x)\in\mathbb{C}(x), deg(f)=m\deg(f)=m. 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 ω\omega is relatively cohomologous to the form

(9.19) ω=g(x,y)dF+dR(x,y)+κ(x)ydx,\omega=g(x,y)dF+dR(x,y)+\kappa(x)ydx,

for some g(x,y),R(x,y)[x,y]g(x,y),R(x,y)\in\mathbb{C}[x,y], κ(x)[x]\kappa(x)\in\mathbb{C}[x]. Hence, the tangential center problem reduces to the problem of characterizing the vanishing of the integral

(9.20) γ(t)κ(x)y𝑑x\int_{\gamma(t)}\kappa(x)ydx

along some cycle γ(t)H1(F1(t)).\gamma(t)\in H_{1}(F^{-1}(t)).

We represent F1(t0)F^{-1}(t_{0}) as a Riemann surface y=t0f(x)y=\sqrt{t_{0}-f(x)} and consider its first homology group H1(F1(t0))H_{1}(F^{-1}(t_{0})). Let x1(t0),,xm(t0)x_{1}(t_{0}),\ldots,x_{m}(t_{0}) be all the roots of f1(t0)f^{-1}(t_{0}). It is well-known that the homology of the fiber F1(t0){F^{-1}(t_{0})} is generated by simple cycles Cij(t0)C_{ij}(t_{0}) going from xi(t0)x_{i}(t_{0}) to xj(t0)x_{j}(t_{0}) 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 yy 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) g(x,t)=κ(x)tf(x)𝑑xg(x,t)=\int\kappa(x)\sqrt{t-f(x)}dx

the indefinite integral of κ(x)tf(x)\kappa(x)\sqrt{t-f(x)}. Up to the problem of multivaluedness of gg, this gives that the original tangential center problem in the plane reduces to the zero-dimensional tangential center problem

(9.22) C(t)g(x,t)=i=1mnig(xi(t),t)0,\int_{C(t)}g(x,t)=\sum_{i=1}^{m}n_{i}g(x_{i}(t),t)\equiv 0,

where the zero-cycle C(t)=i=1mnixi(t)C(t)=\sum_{i=1}^{m}n_{i}x_{i}(t) is composed of the ramification points xi(t)f1(t)x_{i}(t)\in f^{-1}(t) in the presentation of the class of γ(t)\gamma(t) in H1(f1(t))H_{1}(f^{-1}(t)). Note that i=1mni=0\sum_{i=1}^{m}n_{i}=0, that is, C(t)C(t) is indeed a cycle as the class of γ(t)\gamma(t) is generated by simple cycles.

The explicit dependence of g(x(t),t)g(x(t),t) on tt is not a problem. We must however be more careful here as the function g(x,t)g(x,t) is multivalued in xx due to the multivaluedness of the square root. It has the same (two-sheeted) Riemann surface as the function y=tf(x).y=\sqrt{t-f(x)}. We will now be more precise in the choices we make in representing the cycle γ(t)\gamma(t). We take t0t_{0} big, order cyclically the roots xi(t0)x_{i}(t_{0}) and suppose for simplicity that mm is even. We present the Riemann surface of gg as a two-sheeted surface with cuts joining x1(t0)x_{1}(t_{0}) to x2(t0)x_{2}(t_{0}), next x3(t0)x_{3}(t_{0}) to x4(t0)x_{4}(t_{0}) etc. When crossing any cut a path changes the sheet. Let γij(t0)\gamma_{ij}(t_{0}) be the cycle going first from xi(t0)x_{i}(t_{0}) to xj(t0)x_{j}(t_{0}) in the upper sheet and then along the same projection from xj(t0)x_{j}(t_{0}) to xi(t0)x_{i}(t_{0}) 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 xi(t0)x_{i}(t_{0})).

If mm is odd, we have to add a cut from xmx_{m} to infinity, but we will stick rather to the case when mm is even for simplicity. The space ={γi,i+1(t0):i=1,,m1}\mathcal{B}=\{\gamma_{i,i+1}(t_{0}):i=1,\ldots,m-1\} provides a basis of H1(F1(t0))H_{1}(F^{-1}(t_{0})). Any cycle γ(t0)\gamma(t_{0}) can hence be written in a unique way as

(9.23) γ(t0)=i=1m1niγi,i+1(t0).\gamma(t_{0})=\sum_{i=1}^{m-1}n_{i}\gamma_{i,i+1}(t_{0}).

We associate the zero-cycle

(9.24) ϕ(γ(t0))=C(t0)=i=1m1ni(xi+1(t0)xi(t0))\phi(\gamma(t_{0}))=C(t_{0})=\sum_{i=1}^{m-1}n_{i}(x_{i+1}(t_{0})-x_{i}(t_{0}))

to the one-cycle γ(t0)\gamma(t_{0}). In fact, we have constructed an isomorphism

(9.25) ϕ:H1(F1(t0))H~0(f1(t0)).\phi:H_{1}(F^{-1}(t_{0}))\to\tilde{H}_{0}(f^{-1}(t_{0})).

Then

(9.26) γ(t0)κ(x)y𝑑x=2C(t0)g.\int_{\gamma(t_{0})}\kappa(x)ydx=2\int_{C(t_{0})}g.

We say that a cycle γ(t)\gamma(t) in F1(t)F^{-1}(t) is balanced (unbalanced, totally unbalanced) if the corresponding zero-cycle ϕ(γ(t))\phi(\gamma(t)) in f1(t)f^{-1}(t) is balanced (unbalanced, totally unbalanced).

Let GFG_{F} denote the usual action of the monodromy group of FF on H1(F1(t0))H_{1}(F^{-1}(t_{0})) and let GF0G_{F}^{0} denote the conjugated group:

GF0={ϕσϕ1:σGF}.G^{0}_{F}=\{\phi\circ\sigma\circ\phi^{-1}:\sigma\in G_{F}\}.

Then GF0G^{0}_{F} acts on H~0(f1(t0))\tilde{H}_{0}(f^{-1}(t_{0})).

The action of the group GF0G^{0}_{F} is very much related to the action of the monodromy group GfG_{f} of ff, but in general is more complicated due to cuts in the Riemann surface of gg.

We believe that the following conjecture, analogous to Theorem 2.2, holds:

Conjecture 9.1.

Let F(x,y)=y2+f(x)[x,y]F(x,y)=y^{2}+f(x)\in\mathbb{C}[x,y], κ[x]\kappa\in\mathbb{C}[x] and let g(x)g(x) be given by (9.21).

  1. (1)

    If γ(t)\gamma(t) is a totally unbalanced cycle of FF, then

    (9.27) γ(t)κ(x)y𝑑x0\int_{\gamma(t)}\kappa(x)ydx\equiv 0

    if and only if there exist f1,,fs,h1,,hs[x]f_{1},\ldots,f_{s},h_{1},\ldots,h_{s}\in\mathbb{C}[x] and analytic functions in a neighborhood of infinity g1,,gsg_{1},\ldots,g_{s} such that f=fkhkf=f_{k}\circ h_{k}, g=g1h1++gshsg=g_{1}\circ h_{1}+\ldots+g_{s}\circ h_{s} and the cycles Hk(γ)H_{k}(\gamma) are trivial for every k=1,,sk=1,\ldots,s, where Hk(x,y)=(hk(x),y).H_{k}(x,y)=(h_{k}(x),y).

  2. (2)

    If γ(t)\gamma(t) is unbalanced, then (9.27) holds if and only if there exist f1,,fs,h1,,f_{1},\ldots,f_{s},h_{1},\ldots, hs[x]h_{s}\in\mathbb{C}[x], g1,,gsg_{1},\ldots,g_{s} analytic in a neighborhood of infinity such that f=fkhkf=f_{k}\circ h_{k}, g=g1h1++gshsg=g_{1}\circ h_{1}+\ldots+g_{s}\circ h_{s}, and for every kk, either the projected cycle Hk(γ){H_{k}}({\gamma}) is trivial or it is balanced and

    hk(ϕ(γ(t)))gk0.\int_{{h_{k}}({\phi(\gamma(t))})}g_{k}\equiv 0.

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 GfG_{f}-irreducible and GfG_{f}-invariant by GF0G_{F}^{0}-irreducible and GF0G_{F}^{0}-invariant, then we could prove the conjecture by using the isomorphism ϕ\phi and then following the proof of Theorem 2.2. Here it is essential that the function ff is polynomial. The inductive argument is on the divisors of the degree of ff. The fact that the function gg is only analytic is not a problem. The difficulty is that GF0G_{F}^{0} is not a permutation of the roots as GfG_{f}. In fact it acts on cycles and not on individual roots.

Example 9.2.

Let F(x,y)=y2+xmF(x,y)=y^{2}+x^{m} and assume that mm is even. Let γ(t)\gamma(t) be a balanced cycle of FF. Let PγP_{\gamma} be the polynomial associated to the cycle γ\gamma. Note that here the group GF0G_{F}^{0} coincides with GfG_{f} because there is only one critical value and hence only one generator (corresponding to winding around zero). Hence Proposition 7.3 applies and γ(t)κ(x)y𝑑x\int_{\gamma(t)}\kappa(x)ydx vanishes if and only if bj(t)=0b_{j}(t)=0, whenever Φm/k(x)\Phi_{m/k}(x) does not divide Pγ(x)P_{\gamma}(x), where k=gcd(m,j)k=gcd(m,j) and g(x,t)=bj(t)xjg(x,t)=\sum b_{j}(t)x^{j} is the development of gg given in (9.21) in a neighborhood of infinity. Following Example 7.4, for F(x,y)=y2+x4F(x,y)=y^{2}+x^{4}, and the balanced cycle γ(t)=ϕ1(C(t))\gamma(t)=\phi^{-1}(C(t)), with C(t)=a(x1(t)x2(t)+x3(t)x4(t))C(t)=a(x_{1}(t)-x_{2}(t)+x_{3}(t)-x_{4}(t)), we have that γ(t)κ(x)y𝑑x0\int_{\gamma(t)}\kappa(x)ydx\equiv 0 if and only if the function g(x,t)g(x,t) given by (9.21) is of the form g(x,t)=g0(x4,t)+xg1(x4,t)+x3g3(x4,t)g(x,t)=g_{0}(x^{4},t)+xg_{1}(x^{4},t)+x^{3}g_{3}(x^{4},t), where the functions gig_{i}, for fixed tt are analytic in a neighborhood of infinity in xx.

Recall the definition of gg. We use the development in a neighborhood of infinity. There tx4\sqrt{t-x^{4}} develops as x2x^{2} times a function of (x4,t)(x^{4},t). This gives that the powers of xx appearing in κ(x)tx4\kappa(x)\sqrt{t-x^{4}} are by two higher than the powers in κ(x)\kappa(x). One extra shift in powers comes from integration. This gives all together that the powers appearing in κ(x)\kappa(x) are by 33 modulo 44 higher than the powers in gg. Finally we get that γ(t)κ(x)y𝑑x0\int_{\gamma(t)}\kappa(x)ydx\equiv 0 if and only if the polynomial κ(x)\kappa(x) is of the form κ(x)=κ1(x4)+x2κ2(x4)+x3κ3(x4)\kappa(x)=\kappa_{1}(x^{4})+x^{2}\kappa_{2}(x^{4})+x^{3}\kappa_{3}(x^{4}), with κi[x]\kappa_{i}\in\mathbb{C}[x].

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) ϵx′′+α(x)x+β(x)=0,\epsilon x^{\prime\prime}+\alpha(x)x^{\prime}+\beta(x)=0,

for certain rational functions α,β\alpha,\beta.

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 uu 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 uu.

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

x=yf(x),y=ϵG(x),\begin{split}x^{\prime}&=y-f(x),\\ y^{\prime}&=\epsilon G(x),\end{split}

where f,Gf,G are rational functions and ϵ(0,+)\epsilon\in(0,+\infty). We shall assume in addition that ff is a Morse function f(x)=12x2+o(x2)f(x)=\frac{1}{2}x^{2}+o(x^{2}). Note that for every t>0t>0 close enough to zero, f(x)=tf(x)=t has exactly two solutions also close to zero. Let z1(t)<0<z2(t)z_{1}(t)<0<z_{2}(t) denote those two solutions. Also, assume that

G(x)=a0x(1+O(x)),G(x)=a_{0}-x(1+O(x)),

with 1+O(x)>01+O(x)>0 for xx close to zero.

Now the system rewrites as

(10.29) x=yf(x),y=ϵ(a0xG¯(x)).\begin{split}x^{\prime}&=y-f(x),\\ y^{\prime}&=\epsilon(a_{0}-x\bar{G}(x)).\end{split}

We blow-up the family. Here we consider only one chart:

(10.30) (x,y,a0,ϵ)=(ux¯,u2y¯,uA0,u2).(x,y,a_{0},\epsilon)=(u\bar{x},u^{2}\bar{y},uA_{0},u^{2}).

Dividing by uu, the blown up vector field, has a center at (x¯,y¯)=(0,0)(\bar{x},\bar{y})=(0,0) on the divisor u=0u=0, for A0=0A_{0}=0. 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 (A0,u)(A_{0},u) close to (0,0)(0,0) u>0u>0 there is a 𝒞\mathcal{C}^{\infty}-function Δ(y¯,A0,u)\Delta(\bar{y},A_{0},u) defined on a transversal x¯=0\bar{x}=0 such that the closed orbits are given by Δ(y¯,A0,u)=0\Delta(\bar{y},A_{0},u)=0.

Moreover,

Δy¯=I(y¯)+O(u),\frac{\partial\Delta}{\partial\bar{y}}=-I(\bar{y})+O(u),

and

(10.31) I(t)=z1(t)z2(t)x¯w(x¯)F2(x¯)𝑑x¯,I(t)=\int_{z_{1}(t)}^{z_{2}(t)}\bar{x}w(\bar{x})F^{2}(\bar{x})\,d\bar{x},

where

w(x¯)=x¯G¯(x¯),F(x¯)=1x¯fx¯(x¯).w(\bar{x})=-\dfrac{\bar{x}}{\bar{G}(\bar{x})},\quad F(\bar{x})=\frac{1}{\bar{x}}\frac{\partial f}{\partial\bar{x}}(\bar{x}).

If I(t)0I(t)\equiv 0 for some w(x)w(x) and all t>0t>0, then we have that (10.29) has a center to first order in uu. Rewriting (10.31) as

I(t)=g(z1(t))g(z2(t)),I(t)=g(z_{1}(t))-g(z_{2}(t)),

where gg is a primitive of x¯w(x¯)F2(x¯)\bar{x}w(\bar{x})F^{2}(\bar{x}), the problem of characterizing I(t)0I(t)\equiv 0 is equivalent to the zero-dimensional tangential center problem for the cycle C(t)=z1(t)z2(t)C(t)=z_{1}(t)-z_{2}(t).

Proposition 5.3 states that gg is a solution iff there exist f0,g0,hf_{0},g_{0},h with h(z1(t))=h(z2(t))h(z_{1}(t))=h(z_{2}(t)) such that f=f0hf=f_{0}\circ h, and g=g0hg=g_{0}\circ h. Therefore,

g0(h(x¯))h(x¯)=x¯w(x¯)F2(x¯).g_{0}^{\prime}(h(\bar{x}))h^{\prime}(\bar{x})=\bar{x}w(\bar{x})F^{2}(\bar{x}).

Now, replacing we get

G¯(x¯)=(fx¯(x¯))2g0(h(x¯))h(x¯)\bar{G}(\bar{x})=-\frac{\left(\frac{\partial f}{\partial\bar{x}}(\bar{x})\right)^{2}}{g_{0}^{\prime}(h(\bar{x}))h^{\prime}(\bar{x})}

for any g0g_{0}^{\prime} and any hh such that h(z1(t,c))=h(z2(t,c))h(z_{1}(t,c))=h(z_{2}(t,c)) and f=f0hf=f_{0}\circ h. Thus, given ff, it is possible to obtain all G¯(z)\bar{G}\in\mathbb{C}(z) such that the perturbed system has a center at first order in uu.

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 fi(x)=xif_{i}(x)=x^{i} or Chebyshev polynomials Ti(x)=cos(iarccos(x))T_{i}(x)=\cos(i\arccos(x)) as composition factors of ff plays a center role in balanced tangential centers. The study of balanced centers of fi(x)=xif_{i}(x)=x^{i} 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 dw/dz=Rz/Rwdw/dz=-R_{z}/R_{w}, where R(z;w)R(z;w) 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.