Home The Poincaré map of degenerate monodromic singularities with Puiseux inverse integrating factor
Article Open Access

The Poincaré map of degenerate monodromic singularities with Puiseux inverse integrating factor

  • Isaac A. García and Jaume Giné EMAIL logo
Published/Copyright: May 25, 2023

Abstract

We consider analytic families of planar vector fields depending analytically on the parameters in Λ that guarantee the existence of a (may be degenerate and with characteristic directions) monodromic singularity. We characterize the structure of the asymptotic Dulac series of the Poincaré map associated to the singularity when the family possesses a Puiseux inverse integrating factor in terms of its multiplicity and index. This characterization is only valid in a restricted monodromic parameter space Λ \ Λ associated to the nonexistence of local curves with zero angular speed. As a byproduct, we are able to study the center-focus problem (under the assumption of the existence of some Cauchy principal values) in very degenerated cases where no other tools are available. We illustrate the theory with several nontrivial examples.

MSC 2010: 34C05; 34A05; 37G15; 37G10

1 Introduction

We consider families of real analytic planar differential systems

(1) x ˙ = P ( x , y ; λ ) , y ˙ = Q ( x , y ; λ )

depending analytically on the parameters λ R p together with its associate family of vector fields X λ = P ( x , y ; λ ) x + Q ( x , y ; λ ) y . Throughout the work, we restrict the family to the parameter space Λ R p so that the origin ( x , y ) = ( 0 , 0 ) is made sure to be a monodromic singularity of the whole family (1). That means that P ( 0 , 0 ; λ ) = Q ( 0 , 0 ; λ ) = 0 and the local orbits of X λ turn around the origin for any λ Λ . The characterization of Λ is usually done by means of the blow-up procedure described by Dumortier in [10], see also Arnold [7] and Medvedeva [30], for example. In [4,5], Algaba et al. give a new algorithmic criterium to determine the parameter constrains defining the set Λ , which is based on the analysis of the Newton diagram of the vector field X λ . This alternative algorithm can be easily checked and improves the previous schemes in some technical aspects.

In this monodromic scenario, Il’yashenko [26] and Écalle [11] show independently that the origin only can be either a center or a focus since X λ is analytic. We recall that a center possesses a punctured neighborhood (period annulus) foliated by periodic orbits of X λ , while in a neighborhood of the focus, the orbits spiral around it. The characterization of the subsets of Λ for which one of these two possible behaviors occurs is called the center-focus problem associated to the origin of X λ .

When the differential matrix D X λ ( 0 , 0 ) of X λ at the origin has nonzero pure imaginary eigenvalues, the center-focus problem was already solved in the Poincaré and Lyapunov works, and the reader may consult the modern text books [12,32]. Later, Moussu [31] solved the center-focus problem in the nilpotent case that consists in the situation in which D X λ ( 0 , 0 ) is not identically zero, but its two eigenvalues are zero.

To give algorithms to know the stability of the monodromic singular points in the degenerate case corresponding to have D X ( 0 , 0 ) identically zero is a challenging problem [21,23,24], and this article tries to advance in this direction under additionally restrictions specified below. So we deal with the very difficult degenerate center-focus problem even though our approach also works in the other less degenerate situations. The desingularization blow-up procedure transforms the origin into a monodromic polycycle Γ with associated Poincaré return map Π on the one side. After Il’yashenko’s work [26] we know that, in general, the Poincaré map Π : Σ ( R + , 0 ) ( R + , 0 ) is no longer differentiable at the origin but it is a semiregular map. In consequence Π can be expressed as a Dulac asymptotic expansion with linear leading term. In this setting the center case is characterized by Π equal to the identity map.

Several degrees of difficulty arise depending on the nature of Γ . The easiest case corresponds to a Γ without singularities because there are no characteristic directions after the first (weighted) polar blow-up, so that Γ is a periodic orbit. We denote by Mo ( p , q ) this monodromic class, see Definition 12 to be more precise. In the class Mo ( p , q ) , the Poincaré map Π is analytic, and therefore, the classical scheme proposed by Poincaré and Lyapunov to solve the center problem works with minor modifications, and it is known as Bautin method. The reader may consult, for instance, see [22] and the references therein.

Due to the analyticity of X λ in λ , an important step was taken by Medvedeva [30] who prove that the Poincaré map Π has a Dulac asymptotic expansion of the form

(2) Π ( x ) = η 1 x + j P j ( log x ) x ν j ,

where η 1 > 0 , the exponents ν j > 1 are independent of λ and grow to infinity, and the coefficients of the P j are polynomials whose coefficients depend analytically on the coefficients of X λ . In particular, Π ( x ) = η 1 x + o ( x ) and Medvedeva in [28] shows how to compute the linear coefficient η 1 after the desingularization of Γ . Under some technical conditions, in [29], Medvedeva and Batcheva shows that when η 1 = 1 , then Π ( x ) = x ( 1 + η 2 x 1 / n + o ( x 1 / n ) ) for some positive integer n . As far as we know, no other results about the structure of Π with specific restrictions on X λ are given in the literature, and in this work, we explore this idea.

Only few works (some of them are [19,20]) give explicit restrictions on X λ to compute η 1 when X λ Mo ( 1 , 1 ) , see also [3] and the references therein. It is also worth to mention that in [20], where Gasull et al. define a kind of monodromic singularities (called the class S k ω in [20]) such that each singular point on the first polar blow-up polycycle Γ desingularizes into two well-posed hyperbolic saddles, and they prove in that case that η 1 can be written as follows:

(3) η 1 ( λ ) = exp PV 0 2 π R k ( θ ; λ ) F k ( θ ; λ ) d θ ,

where R k and F k are homogeneous polynomials of degree k + 1 defined as follows:

R k ( θ ; λ ) = P k ( cos θ , sin θ ; λ ) cos θ + Q k ( cos θ , sin θ ; λ ) sin θ , F k ( θ ; λ ) = Q k ( cos θ , sin θ ; λ ) cos θ P k ( cos θ , sin θ ; λ ) sin θ ,

with X k = P k ( x , y ; λ ) x + Q k ( x , y ; λ ) y being the homogeneous vector field of degree k defining the leading part of X λ , i.e., X λ = X k + , where the dots denote a sum of homogeneous vector fields of degree greater than k . In formula (3), the symbol PV means Cauchy principal value and is defined as follows. Given a continuous function f defined in I [ 0 , 2 π ] \ Ω with Ω = { θ 1 , , θ } , the Cauchy principal value of I f ( θ ) d θ is defined as the following limit (if it exists):

PV I f ( θ ) d θ = lim ε 0 + I ε f ( θ ) d θ ,

where I ε = I \ J ε with J ε = i = 1 ( θ i ε , θ i + ε ) . Improper integrals with more than one singularity are convergent if, splitting the interval of integration, each of the split integrals is convergent. On the other hand, the Cauchy principal value arises when the singularities are handled in a balanced way. Clearly, if an improper integral is convergent, it converges to its principal value, but the latter may exist though the corresponding improper integral does not.

The same expression (3) for η 1 was obtained earlier in [19] for another class of monodromic singularities called the class G in that work. Also Medvedeva in [30] found (3) under different restrictions on X λ .

In this work, we reobtain the same expression (3) under our specific restrictions on X λ . These restrictions are related with the existence of an inverse integrating factor v ( x , y ; λ ) in a neighborhood of the origin for X λ , which can be written in Puiseux series using weighted polar coordinates, see the expression (12). Recall that in this situation the differential 1-form ω = P ( x , y ; λ ) d y Q ( x , y ; λ ) d x satisfies that ω / v is closed (the exterior differential d ( ω / v ) = 0 ) off the set v 1 ( 0 ) . The inverse integrating factor with Laurent expansion was used in [18] to obtain the coefficients of the Taylor expansion of Π in the simplest situation when Γ is a periodic orbit, and therefore, X λ Mo ( p , q ) . In the recent work [16], the center conditions of the degenerate center-focus problem for X λ in Λ was considered using inverse integrating factors whose expression in polar coordinates has the exceptional divisor of the first polar blow-up as isolated zero-set. The present work is a wide generalization of several results included in [18] and [16].

The work is structured as follows. The basic background and the main results are given in Section 2. Section 3 is dedicated to summarize well-known results about the characterization of Λ , while in Section 4, we develop the concepts associated to the weighted polar blow-up. In the next section, we make a brief break to explain what is the algebraic origin of Puiseux inverse integrating factors. In Section 6, we give the proofs of all the results. Section 7 is dedicated to present some nontrivial examples illustrating how the theory works. We also write an Appendix where some technical details that are implicit throughout this work are given.

2 Main results

Hereinafter, we omit the dependency on the parameters λ in all the formulas except when it is relevant.

Developing in Taylor series at ( 0 , 0 ) both components of the analytic system (1), we obtain

(4) x ˙ = P ( x , y ) = i k P i ( x , y ) , y ˙ = Q ( x , y ) = i k Q i ( x , y ) ,

where P i and Q i are homogeneous polynomials of degree i and P k 2 + Q k 2 0 .

Taking polar coordinates ( x , y ) ( θ , r ) with x = r cos θ , y = r sin θ , and rescaling the time (dividing the vector field by r k 1 ) system (4) becomes the polar system

(5) r ˙ = ( θ , r ) = i k R i ( θ ) r i k + 1 , θ ˙ = Θ ( θ , r ) = i k F i ( θ ) r i k ,

with

R i ( θ ) = P i ( cos θ , sin θ ) cos θ + Q i ( cos θ , sin θ ) sin θ , F i ( θ ) = Q i ( cos θ , sin θ ) cos θ P i ( cos θ , sin θ ) sin θ .

We say that θ = θ is a characteristic direction for the origin of system (4) if F k ( θ ) = 0 , and we define Ω as the set of all the characteristic directions:

Ω = { θ S 1 : F k ( θ ) = 0 } = { θ 1 , , θ } .

We recall that we can move a characteristic direction just by taking a linear change of coordinates in system (1).

Considering system (5), and assuming F k ( θ ) 0 , we have Θ ( θ , r ) 0 for 0 < r 1 sufficiently small. The point ( θ , r ) = ( θ , 0 ) with θ Ω is a singularity of the map Θ because, though Θ ( θ , 0 ) = 0 , the hypothesis of the implicit function theorem fail at the singularity since r Θ ( θ , 0 ) = 0 (and also θ Θ ( θ , 0 ) = 0 ) by the first monodromic conditions in Remark 10. We say that a branch of solutions of the equation Θ ( θ , r ) = 0 bifurcates from the characteristic direction θ of (4) if there is a continuous function r ( θ ) 0 , defined for all θ in a half-neighborhood of θ , such that r ( θ ) = 0 and Θ ( θ , r ( θ ) ) 0 . In short, to each characteristic direction θ i Ω , with i = 1 , , , of (4), there can be associated several (or none) branches.

We define the sufficiently small cylinder

(6) C = { ( θ , r ) S 1 × R : 0 r 1 } with S 1 = R / ( 2 π Z )

and the critical set

(7) S = { ( θ , r ) C : Θ ( θ , r ) = 0 } .

Notice that S because Ω although it may happen that

(8) S \ { r = 0 } =

just when no branches appear. In [19], it is proved that the latter condition (8) always occurs when system (4) is polynomial and defined by the sum of two homogeneous vector fields. Of course, S depends on λ .

Remark 1

In general, Θ ( θ , r ) = F k ( θ ) + F k + 1 ( θ ) r + F k + 2 ( θ ) r 2 + O ( r 3 ) , and we may write down sufficient conditions that guarantee (8). For example, (8) holds if one of the following sentences are satisfied for all θ J ε = i = 1 ( θ i ε , θ i + ε ) with ε > 0 sufficiently small:

  • F k + 1 ( θ ) 0 ;

  • F k + 1 2 ( θ ) 4 F k ( θ ) F k + 2 ( θ ) < 0 .

We define the critical parameters as those belonging to the subset Λ Λ of the parameter space of X λ where (8) does not hold, that is,

Λ = { λ Λ : S \ { r = 0 } } .

We consider the ordinary differential equation of the orbits of (5):

(9) d r d θ = ( θ , r ) ,

where ( θ , r ) = ( θ , r ) / Θ ( θ , r ) is a function well defined in C \ S .

We say that a not locally null real C 1 ( C \ S ) function V ( θ , r ) is an inverse integrating factor of (9) if it is a solution of the linear partial differential equation:

(10) V θ ( θ , r ) + V r ( θ , r ) ( θ , r ) = r ( θ , r ) V ( θ , r ) .

Remark 2

If v ( x , y ) is an inverse integrating factor of (1), that is, ( P ( x , y ; λ ) d y Q ( x , y ; λ ) d x ) / v ( x , y ) is closed off the zero-set v 1 ( 0 ) , then

(11) V ( θ , r ) = v ( r cos θ , r sin θ ) r k Θ ( θ , r )

is an inverse integrating factor of (9) in C \ { S { r = 0 } } .

From now on, we will fix our attention on a special class of functions V . We say that an inverse integrating factor V ( r , θ ) of (9) in C \ { S { r = 0 } } is a Puiseux inverse integrating factor if it can be expanded in convergent Puiseux series about r = 0 of the form

(12) V ( r , θ ; λ ) = i m v i ( θ ; λ ) r i / n

whose coefficients, for a fixed λ , are C 1 functions v i : S 1 \ Ω R , the leading coefficient v m ( θ ; λ ) 0 , and ( m , n ) Z × N are fixed numbers called multiplicity and index. Here, we have used the notation N = N \ { 0 } . The particular case n = 1 leads to a Laurent inverse integrating factor, while n = 1 and m 0 produces an analytic inverse integrating factor in case that the v i are all analytic. It may happen that ( m , n ) are dependent of λ , and in this case, our results are no longer valid for all the family X λ with λ Λ rather they are still valid for specific vector fields X λ ˆ with λ ˆ Λ ˆ being Λ ˆ = { λ Λ : ( m ( λ ) , n ( λ ) ) Z × N } . In the particular case that ( m , n ) is independent of λ , then Λ ˆ = Λ .

Remark 3

Equation (9) given by d r / d θ = r 3 possesses the inverse integrating factor

(13) V ( θ , r ) = r 3 sin ( 2 θ + r 2 )

which is of class C 1 in C \ { r = 0 } . We note that (13) cannot be expanded in Puiseux series about r = 0 and hence does not have the form (12).

The following first result is valid in all the monodromic parameter space Λ and shows a wide class of systems (1) having a center at the origin when a Puiseux inverse integrating factor with nonpositive multiplicity exists.

Theorem 4

Let the origin be a monodromic singularity of system (1). If the associated differential equation (9) possesses a Puiseux inverse integrating factor (12) with multiplicity m 0 , then the origin is a center of (1).

Our second main result is only valid in the restricted parameter space Λ ˆ \ Λ . It characterizes the structure of the asymptotic Dulac expansion of Π in Λ ˆ \ Λ under the assumption that family (1) possesses a Puiseux inverse integrating factor in terms of its multiplicity m and index n . In particular, its statement (iii) extends (in Λ ˆ \ Λ ) the scope of Theorem 4.

Theorem 5

Let the origin be a monodromic singularity of family (1) with parameters in Λ , 0 Ω and Poincaré map Π ( r 0 ; λ ) = η 1 ( λ ) r 0 + o ( r 0 ) with η 1 > 0 . Assume the existence of a Puiseux inverse integrating factor V ( r , θ ) as in (12) of equation (9) in Λ ˆ Λ . Then, restricting the family to the parameter space Λ ˆ \ Λ , the following holds:

  1. If η 1 1 , then m = n ;

  2. If η 1 = 1 and the origin is a focus, then m > n and the asymptotic Dulac expansion of Π is

    Π ( r 0 ; λ ) = r 0 + η ( λ ) r 0 m / n log j ( r 0 ) + o ( r 0 m / n log j ( r 0 ) )

    with η 0 and j N { 0 } .

  3. If m < n , then the origin is a center, that is, Π ( r 0 ; λ ) = r 0 .

From Theorem 5(ii) and expression (2), it follows the next important corollary.

Corollary 6

Let the origin be a monodromic singularity of family (1) with parameters in Λ , 0 Ω and having a Puiseux inverse integrating factor (12) of (9) in Λ ˆ Λ . If the Poincaré map of system (1) is Π ( r 0 ; λ ) = r 0 + o ( r 0 ) then, restricting the family to the parameter space Λ ˆ \ Λ , its Dulac asymptotic expansion (2) is

(14) Π ( r 0 ; λ ) = r 0 + P 0 ( log r 0 ; λ ) r 0 m / n + i = 1 P i ( log r 0 ; λ ) r 0 ν i ,

where the exponents ν i > m / n grow to infinity. Moreover, the origin is a center if and only if P 0 ( log r 0 ; λ ) 0 , and hence, the center variety is an analytic variety in the parameter space of codimension less than or equal to the degree of P 0 plus 1.

We end the main results of this work focusing in families of vector fields in the monodromic class Mo ( 1 , 1 ) . For such families, we define the quantity ξ as follows:

(15) ξ = 0 2 π R k ( θ ) F k ( θ ) d θ ,

and it will play a fundamental role in the results presented below.

From the Puiseux series ( r , θ ) / V ( r , θ ) = r ( n m ) / n i 0 g i ( θ ) r i / n , we also define the following relevant quantities

ξ i = 0 2 π g i ( θ ) d θ .

The explicit expressions of the ξ i in terms of the coefficients of system (5) and the Puiseux inverse integrating factor (12) are rather complicated. As a sample, in the particular case n = 1 , after some extensive but routine calculations, we obtain

ξ 1 = 0 2 π F k + 1 ( θ ) R k ( θ ) v m ( θ ) + F k ( θ ) R k + 1 ( θ ) v m ( θ ) F k ( θ ) R k ( θ ) v m + 1 ( θ ) F k 2 ( θ ) v m 2 ( θ ) d θ ,

while

ξ 2 = 0 2 π 1 F k 2 ( θ ) v 2 3 ( θ ) [ F k + 1 ( θ ) R k ( θ ) v 2 2 ( θ ) + F k ( θ ) R k + 1 ( θ ) v 2 2 ( θ ) + F k ( θ ) R k ( θ ) v 3 2 ( θ ) F k ( θ ) R k ( θ ) v 2 ( θ ) v 4 ( θ ) ] d θ ,

when n = m = 2 .

The next main result generalizes the following well-known fact about centers in monodromic singularities of homogeneous vector fields (indeed, it can be stated in a more general form for quasi-homogeneous vector fields). The homogeneous vector field X k = P k ( x , y ) x + Q k ( x , y ) y with a monodromic singularity at the origin ( F k ( θ ) 0 ) has the inverse integrating factor x Q k y P k , which is transformed into the Puiseux inverse integrating factor r 2 of the corresponding differential equation (9) with m = 2 and n = 1 . Moreover, ξ 0 = 0 2 π R k ( θ ) / F k ( θ ) d θ and ξ 1 = 0 , and the origin is a center of X k if and only if ξ 0 = 0 .

Theorem 7

Let the origin be a monodromic singularity of family (1) with parameters in Λ with the flow rotating counterclockwise around it, with 0 Ω after a rotation of coordinates if necessary, and Poincaré map Π ( r 0 ; λ ) = η 1 ( λ ) r 0 + o ( r 0 ) with η 1 > 0 . We assume that the associated differential equation (9) possesses a Puiseux inverse integrating factor (12) with m n in Λ ˆ Λ . If the origin of (1) belongs to the class Mo ( 1 , 1 ) , then η 1 ( λ ) = exp ( ξ ( λ ) ) and the following holds:

  1. If the origin is a center of system (1), then ξ 0 ( λ ) = ξ 1 ( λ ) = = ξ m n ( λ ) = 0 .

  2. When m > n , then η 1 = 1 and, if ξ 0 ( λ ) = ξ 1 ( λ ) = = ξ m n ( λ ) = 0 , then the origin is a center of system (1).

Remark 8

The case F k ( θ ) 0 in which the flow of (4) rotates clockwise around the origin can be reduced to the opposite monodromic case F k ( θ ) 0 of counterclockwise rotation just by changing the sign of the independent variable t of (4). The outcome is that Theorem 7 still works but for example with η 1 = exp ( ξ ) instead of the original formula for η 1 in Theorem 7.

Remark 9

It is very important to emphasize here that all the main results (Theorems 4, 5, and 7) presented in this work using polar coordinates can be generalized to the weighted polar coordinates (17) explained in section 4 in a straightforward way. Indeed, this idea is applied in most of the examples presented in this work.

3 Monodromic conditions

In this section, we summarize some well-known results about necessary and some sufficient conditions for the origin of (1) be monodromic. In other words, we will present some restriction on the parameters of family (1) in order that they lie in Λ .

The characteristic direction for the origin of system (4) appears from the linear factors in R [ x , y ] of the homogeneous polynomial x Q k ( x , y ) y P k ( x , y ) . Therefore, unless that polynomial be identically zero, the number of characteristic directions is less than or equal to k + 1 . It is well known that necessary monodromy conditions for the origin of system (4) are that k be odd, F k 0 , and F k ( θ ) does not change sign, hence F k ( θ ) 0 reversing the time if necessary. Therefore, without loss of generality, we focus in the case F k ( θ ) 0 , so that the flow of (4) rotates counterclockwise around the origin, and when we use the term monodromic, it is implicit the above direction of rotation.

Remark 10

For each θ Ω , the following necessary monodromic conditions can be found in [17]:

  1. F k ( θ ) = 0 ,

  2. R k ( θ ) = 0 ,

  3. F k + 1 ( θ ) = 0 ,

  4. F k ( θ ) 0 ,

  5. F k ( θ ) 2 R k ( θ ) 0 ,

  6. [ F k + 1 ( θ ) R k + 1 ( θ ) ] 2 2 [ F k ( θ ) 2 R k ( θ ) ] F n + 2 ( θ ) 0 ,

  7. F k + 2 ( θ ) 0 ,

  8. the minimum s such that F s ( θ ) 0 is odd,

  9. if F k + 2 ( θ ) = 0 then F k + 1 ( θ ) = R k + 1 ( θ ) and either F k ( θ ) = R k ( θ ) = F k + 1 ( θ ) = R k + 1 ( θ ) = 0 , or F k ( θ ) 2 R k ( θ ) > 0 .

We have that the invariant circle Γ = { r = 0 } is a polycycle of system (5) and that ( r , θ ) = ( 0 , θ ) is a critical point of (5) for any θ Ω whose linear part is identically zero, and hence, { r = 0 } is a nonelementary polycycle.

Moreover, if θ is a characteristic direction, then θ + π is too because ( ( θ + π , r ) , Θ ( θ + π , r ) ) = ( ( θ , r ) , Θ ( θ , r ) ) . Moreover, since ( ( θ + π , r ) , Θ ( θ + π , r ) ) = ( 1 ) k ( ( θ , r ) , Θ ( θ , r ) ) , it suffices to consider only the characteristic direction in [ 0 , π ) to know the topological behavior of the orbits of (5) around the rest of characteristic direction.

We will assume that x Q k ( x , y ) y P k ( x , y ) 0 so that the characteristic directions are isolated, that is, the set Ω ¯ [ 0 , π ) of half characteristic directions has cardinality # Ω ¯ ( k + 1 ) / 2 .

The main tool in the analysis of the monodromy of a degenerate singularity is the blow-up technique, see, for example, [7]. Dumortier proved in 1977 that after a finite number of blow-ups, any planar real analytic vector with real isolated singularity can be transformed into an analytic field of directions on a smooth manifold where the singular point becomes an elementary polycycle, that is, a polycycle with a finite number of elementary (hyperbolic or degenerate elementary) singularities different from a focus or a center.

System (4) becomes (5) after performing the first polar blow-up. If # Ω ¯ = 0 , there appear no critical points on Γ and then the Poincaré-Lyapunov theory can be reproduced in a straightforward way, in particular, the Poincaré map Π is analytic. In this work, we assume that # Ω ¯ 1 so that Γ = { r = 0 } becomes a polycycle.

Remark 11

The previous work [20] introduced the so-called monodromic class S k ω as the set of vector fields (4) with F k 0 , F k ( θ ) 0 (the reversing inequality is transformed in the above just changing the time sign), and such that for any θ Ω the following holds:

  1. R k ( θ ) = 0 ;

  2. F k + 1 ( θ ) = 0 ;

  3. [ ( F k 2 R k ) F k ] ( θ ) > 0 ;

  4. [ ( F k + 1 R k + 1 ) 2 2 ( F k 2 R k ) F k + 2 ] ( θ ) < 0 .

The invariant circle { r = 0 } is a polycycle of system (5) whose singular points ( r , θ ) = ( 0 , θ ) for any θ Ω have linear part

R k ( θ ) 0 F k + 1 ( θ ) 0 = 0 0 0 0

when X λ S k ω . Therefore, { r = 0 } is a nonelementary polycycle, and we may try to use the Bendixson-Seidenberg-Dumortier desingularization theorem [10] for analytic vector fields to reduce the problem to the case of an elementary polycycles. For any X λ S k ω in [20], it is proved that the Poincaré return map has the shape Π ( x ) = η 1 x + o ( x ) with leading coefficient (first generalized Poincaré-Lyapunov quantity) given by (3).

In this work, when we need to characterize Λ , that is, when we need necessary and sufficient monodromic conditions, we use Theorems 3 and 4 of [4] or Theorem 2 of [5].

4 The weighted polar blow-up

We take an analytic vector field X = P ( x , y ) x + Q ( x , y ) y with

P ( x , y ) = ( i , j ) N 2 a i j x i y j 1 , Q ( x , y ) = ( i , j ) N 2 b i j x i 1 y j

and define the support of X by supp ( X ) = { ( i , j ) N 2 : ( a i j , b i j ) ( 0 , 0 ) } . Then the boundary of the convex hull of the set

( i , j ) supp ( X ) { ( i , j ) + R + 2 } ,

where R + denote the set positive real numbers, is made up of two open rays and a polygon. Then the Newton diagram N ( X ) of the vector field X is composed by that polygon together with the rays that do not lie on a coordinate axis, if they exist. It is easy to see that selecting the weights ( p , q ) N 2 , with p and q coprimes, given by the tangent q / p of the angle between an edge of N ( X ) and the ordinate axis, one obtains an expansion

(16) X = j r X j ,

where r 1 and X j = P p + j ( x , y ) x + Q q + j ( x , y ) y are ( p , q ) -quasihomogeneous vector fields of degree j . We call X r the leading part of X , which clearly depends on the chosen weights ( p , q ) . We define the set W ( N ( X ) ) N 2 whose elements are all the weights in N ( X ) .

The weighted polar blow-up ( x , y ) ( ρ , φ ) given by

(17) x = ρ p cos φ , y = ρ q sin φ ,

wit Jacobian J ( φ , ρ ) = ρ p + q 1 ( p cos 2 φ + q sin 2 φ ) , brings the ( p , q ) -quasihomogeneous vector field X j = P p + j ( x , y ) x + Q q + j ( x , y ) y of degree j into the system

ρ ˙ = ρ j + 1 F ˆ j ( φ ) D ( φ ) , φ ˙ = ρ j G ˆ j ( φ ) D ( φ ) ,

where

F ˆ j ( φ ) = P p + j ( cos φ , sin φ ) cos φ + Q q + j ( cos φ , sin φ ) sin φ , G ˆ j ( φ ) = p Q p + j ( cos φ , sin φ ) cos φ q P q + j ( cos φ , sin φ ) sin φ , D ( φ ) = p cos 2 φ + q sin 2 φ > 0 .

Therefore, after removing the common factor ρ r / D ( φ ) > 0 , X is transformed into the following form

(18) ρ ˙ = R ˆ ( φ , ρ ) = j r ρ j r + 1 F ˆ j ( φ ) , φ ˙ = Θ ˆ ( φ , ρ ) = j r ρ j r G ˆ j ( φ ) .

If X r has not a monodromic singularity at the origin, then G ˆ r ( φ ) has real roots in [ 0 , 2 π ) . Taking into account that ρ ˙ = O ( ρ ) , φ ˙ = G ˆ r ( φ ) + O ( ρ ) , we see that { ρ = 0 } is a polycycle.

Definition 12

The vector field (16) belongs to the class Mo ( p , q ) if its leading part X r with respect to some weights ( p , q ) W ( N ( X ) ) has a monodromic singularity at the origin.

We consider the ordinary differential equation of the orbits of (18), that is,

(19) d ρ d φ = ˆ ( φ , ρ ) = i 1 ˆ i ( φ ) ρ i ,

where ˆ 1 ( φ ) = F ˆ r ( φ ) / G ˆ r ( φ ) . Then ˆ is a function well defined in C \ S p q being the critical set

S p q = { ( φ , ρ ) C : Θ ˆ ( φ , ρ ) = 0 } ,

so that S p q if Ω p q = { φ S 1 : G ˆ r ( φ ) = 0 } . Analogously as in (8), it may happen that

(20) S p q { ρ = 0 } = .

With this notation, the set of characteristic directions Ω and the critical set S defined in (7) correspond to Ω 11 and S 11 , respectively.

For each weight ( p , q ) W ( N ( X ) ) , we define the subset Λ p q Λ of the parameter space of X as follows:

Λ p q = { λ Λ : S p q \ { ρ = 0 } }

and the critical parameters Λ Λ as the intersection

Λ = ( p , q ) W ( N ( X ) ) Λ p q .

We recall that X Mo ( p , q ) if and only if G ˆ r ( φ ) has no real roots in [ 0 , 2 π ) . Therefore, in particular, S p q = when X Mo ( p , q ) . Moreover, the origin is a center of X r if and only if, additionally, 0 2 π ˆ 1 ( φ ) d φ = 0 , see [21]. It can be checked (using the Bautin method or just by the reciprocal of Theorem 5 in [1]) that if the origin is a center of X Mo ( p , q ) , then it is also a center of X r . In other words, 0 2 π ˆ 1 ( φ ) d φ = 0 is a necessary center condition at the origin of X Mo ( p , q ) .

When X Mo \ Mo ( p , q ) we still may define, like it was done in (15) for the particular case ( p , q ) = ( 1 , 1 ) , the principal value

ξ ˆ = PV 0 2 π ˆ 1 ( φ ) d φ ,

which may exist or not.

Remark 13

We can generalize Remark 2 in the following way. If v ( x , y ) is an inverse integrating factor of the vector field X λ associated to (1) and ( p , q ) W ( N ( X λ ) ) , then

(21) V ( φ , ρ ) = v ( ρ p cos φ , ρ q sin φ ) J ( φ , ρ ) Θ ˆ ( φ , ρ ) ρ r / D ( φ )

is an inverse integrating factor of (19) in C \ { S p q { ρ = 0 } } .

Analogously to the case ( p , q ) = ( 1 , 1 ) , when V ( φ , ρ ) is a Puiseux inverse integrating factor with multiplicity m and index n , from the Puiseux series ˆ ( φ , ρ ) / V ( φ , ρ ) = ρ ( n m ) / n i 0 g ˆ i ( φ ) ρ i / n , we also define the following relevant quantities

ξ ˆ i = PV 0 2 π g ˆ i ( φ ) d φ ,

when they exist.

5 The origin of Puiseux inverse integrating factors

Let V ˆ ( θ , r ) be an inverse integrating factor of the polar vector field X ˆ = Θ ( θ , r ) θ + ( θ , r ) r associated to system (5). Then V ˆ ( θ , r ) satisfies X ˆ ( V ˆ ) = V ˆ div ( X ˆ ) , that is, the linear partial differential equation

(22) V ˆ θ Θ ( θ , r ) + V ˆ r ( θ , r ) = V ˆ ( θ , r ) D ˆ ( θ , r ) ,

where D ˆ ( θ , r ) = div ( X ˆ ) = θ Θ ( θ , r ) + r ( θ , r ) .

The equation of the characteristics of (22) is

d θ Θ ( θ , r ) = d r ( θ , r ) = d v v D ˆ ( θ , r ) ,

and the associated vector field in S 1 × R 2 is

(23) Z = X ˆ + v D ˆ ( θ , r ) v .

By construction, { v = 0 } and { r = 0 } are invariant cylinders of Z and both contain the singularities ( θ , r , v ) = ( θ , 0 , 0 ) for any θ Ω . By the monodromic conditions in Remark 10, the linearizations of Z at these points are as follows:

θ Θ ( θ , 0 ) r Θ ( θ , 0 ) 0 θ ( θ , 0 ) r ( θ , 0 ) 0 0 0 D ˆ ( θ , 0 ) = 0 0 0 0 0 0 0 0 0 .

Let F : S 1 × R 2 R be an analytic function with F ( θ , 0 , v ) = 0 for any θ S 1 and some fixed v R . We say that the equation F ( θ , r , v ) = 0 has a local branch emanating from ( r , v ) = ( 0 , v ) if there is a function f ( θ , r ) with f ( θ , 0 ) = v such that F ( θ , r , f ( θ , r ) ) = 0 for any θ S 1 and all r in a positive half-neighborhood of 0.

Theorem 14

Let F : S 1 × R 2 R and F 1 ( 0 ) an analytic invariant surface of the characteristic vector field (23). If the function f ( θ , r ) is a local C 1 branch of the surface, then it is a Puiseux inverse integrating factor of the polar vector field X ˆ .

If F ( θ , r , v ) = i = 0 s a i ( θ , r ) v i is a polynomial in v of degree s , then Z ( F ) is too. Therefore, the cofactor K such that Z ( F ) = K F must be a function K ( θ , r ) , and the coefficients of F satisfy X ˆ ( a i ) = ( K i D ˆ ) a i , and hence, { a i = 0 } are invariant curves of X ˆ with cofactors K i D ˆ .

The simplest case arises from a s 1 and a i 0 for i = 1 , , s 1 , which corresponds with the specific F ( θ , r , v ) = a 0 ( θ , r ) + v s . Assume that a 0 ( θ , 0 ) = 0 so that F ( θ , 0 , 0 ) = 0 , and hence, the branches bifurcate from ( r , v ) = ( 0 , 0 ) and the function a 0 0 is characterized by X ˆ ( a 0 ) = s D ˆ a 0 with a 0 ( θ , 0 ) = 0 . Recall that r = 0 is always invariant for X ˆ . This behavior is exhibited by the following example in Mo ( p , q ) analyzed in [2]. There it is proved that the system

x ˙ = y 3 + 2 a x 3 y + 2 x ( α x 4 + β x y 2 ) , y ˙ = x 5 3 a x 2 y 2 + 3 y ( α x 4 + β x y 2 )

with α β 0 is not orbitally reversible [25] nor formally integrable, but there are values of ( α , β , a ) with a 0 and a < 1 / 6 (i.e., just the monodromy condition) such that the origin is a center. Taking polar coordinates it follows that the origin has the characteristic directions { 0 , π } . The vector field Z possesses the analytic invariant manifold F 1 ( 0 ) with F ( θ , r , v ) = a 0 ( θ , r ) + v 12 and a 0 ( θ , r ) = r 16 ( 2 r 2 cos 6 θ + 12 a r cos 3 θ sin 2 θ + 3 sin 4 θ ) 13 . The branch of F given by V ˆ ( θ , r ) = ( a 0 ( θ , r ) ) 1 / 12 = v 4 ( θ ) r 4 / 3 + o ( r 4 / 3 ) is a Puiseux inverse integrating factor of X ˆ with index 3 and leading coefficient v 4 ( θ ) = 3 13 / 12 ( sin 4 θ ) 13 / 12 .

6 The proofs

Following the ideas of the recent work [16] and adapting them to our framework, we can prove the next result that is slightly more general than the one presented in [16]. We omit the proof since it is almost equal to that presented in [16].

Theorem 15

Let the origin be a monodromic singularity of system (1) and V ˆ ( θ , r ) be an inverse integrating factor in C \ { r = 0 } of the polar system (5). Consider the inverse integrating factor V ( r , θ ) defined by (A1) of the associated differential equation (9) and assume that V 1 ( 0 ) \ { r = 0 } = . Taking C = I × S 1 the cylinder (6) with I = { 0 } I + , where I + is a sufficiently small positive half-neighborhood of the origin, it follows that the function G : I + R given by

(24) G ( r ) = 0 2 π ( r , θ ) V ( r , θ ) d θ

is well defined, and the origin is a center of (1) if and only if G ( r ) 0 .

A main step in the proof of Theorem 15 is the formula

(25) G ( r 0 ) = r 0 Π ( r 0 ) d r V ( r , 0 ) ,

proved in [16], which will be a key in the proof of Theorem 7.

Before presenting the proofs of the main theorems, we need some preliminary technical results. In the following result, the quantity ξ defined in (15) will be the key to analyze the leading starting term of any Puiseux inverse integrating factor of (9) , independently of the nature (monodromic or not) of the singularity at the origin.

Lemma 16

Let V ( r , θ ) be a Puiseux inverse integrating factor (12) of (9) with analytic leading coefficient v m ( θ ) and consider the quantity ξ defined in (15). Then the following holds:

  1. The real zeros of v m ( θ ) are characteristic directions of the origin of (1).

  2. If m n , then ξ = 0 .

  3. m = n provided that ξ exists and is different from zero. In this case, v m ( θ ) is a constant for all θ S 1 .

Proof

Introducing the expansions V ( θ , r ) = v m ( θ ) r m / n + O ( r ( m + 1 ) / n ) and

( r , θ ) = ( θ , r ) Θ ( θ , r ) = R k ( θ ) r + O ( r ) F k ( θ ) + O ( r ) = r R k ( θ ) F k ( θ ) + O ( r )

into the partial differential equation (10) that defines V and equating the coefficients of the power r m n , we obtain that v m ( θ ) satisfies the linear ordinary differential equation:

(26) n F k ( θ ) v m ( θ ) = ( n m ) R k ( θ ) v m ( θ ) .

A straightforward consequence of (26) is that the simple real zeros of v m ( θ ) lie in Ω . But, since any multiple zero of v m ( θ ) is indeed a simple zero of v m ( θ ) / v m ( θ ) , again from (26), it follows that all the real zeros of v m ( θ ) belong to Ω . This proves statement (i) and also the final part of statement (iii) trivially since F k ( θ ) 0 .

Again from (26), we can write

(27) PV 0 2 π v m ( θ ) v m ( θ ) d θ = n m n ξ .

By definition

PV 0 2 π v m ( θ ) v m ( θ ) d θ = lim ε 0 + I ε v m ( θ ) v m ( θ ) d θ ,

where I ε = S 1 \ i = 1 ( θ i ε , θ i + ε ) . Then we have two cases according to whether 0 lies in Ω :

PV 0 2 π v m ( θ ) v m ( θ ) d θ = log v m ( 2 π ) v m ( 0 ) + i = 1 lim ε 0 + log v m ( θ i ε ) v m ( θ i + ε ) if 0 Ω , i = 1 lim ε 0 + log v m ( θ i ε ) v m ( θ i + ε ) if 0 Ω

If v m ( θ i ) = 0 for some i , by the analyticity of v m , θ i is a finite order zero so that, repeating l’Hopital rule just this order number of times, we see that

lim ε 0 + v m ( θ i ε ) v m ( θ i + ε ) = 1 .

Therefore, since v m ( 2 π ) = v m ( 0 ) , we obtain that

PV 0 2 π v m ( θ ) v m ( θ ) d θ = 0 .

Going back to (27), we have

0 = n m n ξ ,

and the proof is done.□

Remark 17

In the application of Lemma 16, we need to compute the value of ξ defined in (15). We recall that in [20], it is studied when this Cauchy principal value is algebraically computable. There it is also proved that the former principal value exists for the subset of vector fields (4) satisfying F k ( θ j ) = F k ( θ j ) = R k ( θ j ) = 0 and F k ( θ j ) 0 for all θ j Ω and they give a method to compute it via a nonsingular integral. Observe also that this subset contains the class S k ω , but that condition F k ( θ j ) 0 does not hold in general for the monodromic points.

The statement (iii) of the following lemma will be used each time we apply Theorem 15 with a Puiseux inverse integrating factor (12).

Lemma 18

Let V ( r , θ ) be a Puiseux inverse integrating factor (12) of (9). Then V 1 ( 0 ) \ { r = 0 } = in the following particular cases:

  1. The leading coefficient of V is v m ( θ ) 0 for any θ S 1 ;

  2. m = n ;

  3. The origin is a monodromic singularity of system (1).

Proof

Statement (i) is a straightforward consequence of the factorization

V ( r , θ ) = r m / n ( v m ( θ ) + O ( r 1 / n ) )

since the function inside the brackets in this expression is a unit in a neighborhood of r = 0 . The part (ii) follows by Lemma 16(iii).

Assume that the origin is a monodromic singularity of system (4), so that { r = 0 } is a polycycle of (9). Taking into account that V 1 ( 0 ) is an invariant set, if V 1 ( 0 ) \ { r = 0 } , the only possibility is that V 1 ( 0 ) contains a separatrix in C of some singularity ( r , θ ) = ( 0 , θ ) on the polycycle. But this situations cannot occur by monodromy. This proves statement (iii).□

6.1 Proof of Theorem 4

Proof

Let ω = d r ( r , θ ) d θ be the Pfaffian 1-form associated to equation (9). Then ω / V is a closed 1-form having the expression

ω V = d r ( r , θ ) d θ V ( θ , r ) = Θ ( θ , r ) d r ( θ , r ) d θ V ˆ ( θ , r ) ,

where V ˆ ( θ , r ) = V ( θ , r ) Θ ( θ , r ) , see (A1). Since V ( θ , r ) = v m ( θ ) r m / n + o ( r m / n ) is a Puiseux series with index n and Θ ( θ , r ) = F k ( θ ) + O ( r ) is a power series in r , it follows that V ˆ ( θ , r ) is also a Puiseux series with index n . More precisely V ˆ ( θ , r ) = r m / n ( v ˆ m ( θ ) + O ( r 1 / n ) ) with v ˆ m ( θ ) = v m ( θ ) F k ( θ ) , which only vanishes at Ω .

We continue assuming that m 0 so that ω / V is well-defined in C = C \ { ( r , θ ) = ( 0 , θ ) : θ Ω } .

Since ω / V is closed, using De Rham’s theorem (see [13]) with the topology of the cylinder C , we know that ω / V is exact if and only if its line integral along any noncontractible cycle in C vanishes. We take the following family of noncontractible loops γ ε C , which are defined as segments of r = 0 with consecutive endpoints joined by semicircles of radius ε > 0 sufficiently small with center at each point ( 0 , θ ) with θ Ω . Since the oval r = 0 is an orbit of the equation Θ ( θ , r ) d r ( θ , r ) d θ = 0 , it is clear that

0 = lim ε 0 γ ε ω V .

Therefore, by virtue of De Rham’s theorem, we have that the closed 1-form ω / V is indeed exact on C , and therefore,

0 = { r = r 0 } ω V

for any r 0 > 0 sufficiently small. This condition can be rewritten as follows:

0 = 0 2 π ( r 0 , θ ) V ( r 0 , θ ) d θ = G ( r 0 ) ,

hence, the function G defined in Theorem 15 is G ( r 0 ) 0 . Since V 1 ( 0 ) \ { r = 0 } = holds by Lemma 18(iii), we conclude that the origin is a center of (4) applying Theorem 15.□

6.2 Proof of Theorem 5

Adapting Theorem 3 of the work [15] to our context, it follows the next fundamental relation between the Poincaré map Π of the monodromic singularity and the inverse integrating factors of equation (9):

(28) V ( Π ( r 0 ) , 0 ) = V ( r 0 , 0 ) Π ( r 0 ) .

Remark 19

Equation (28) is proved under the assumption that the function V C 1 ( C \ { r = 0 } ) . Therefore (28) only can be used with a Puiseux inverse integrating factor (12) only when (8) holds. In other words, we must exclude from the monodromic parameter space Λ the critical parameters subset Λ .

Proof

We shall use the following properties using the Landau little-o notation when r 0 0 :

  • r 0 ( r 0 ν log i ( r 0 ) ) = ν r 0 ν 1 log i ( r 0 ) + o ( r 0 ν 1 log i ( r 0 ) ) with ν > 1 ;

  • r 0 ν 2 ( log ( r 0 ) ) i = o ( r 0 ν 1 ( log ( r 0 ) ) j ) when ν 2 > ν 1 > 1 and ( i , j ) N 2 ;

  • r 0 ν log i ( r 0 ) = o ( r 0 ν log j ( r 0 ) ) when ν > 1 and i < j .

The Poincaré map is Π ( r 0 ) = η 1 r 0 + o ( r 0 ν log j r 0 ) with ν > 1 and j 0 , and hence, Π ( r 0 ) = η 1 + o ( r 0 ν 1 log j r 0 ) . From (12), we know that V ( r , θ ) = v m ( θ ) r m / n + o ( r m / n ) . Inserting the expressions of Π and V in equation (28), the lowest degree terms in powers of r 0 correspond to the power r 0 m / n and equating the coefficients gives

v m ( 0 ) ( η 1 m / n η 1 ) = 0 ,

and hence, either v m ( 0 ) = 0 or m = n when η 1 1 . Then part (i) of Theorem 5 is proved since v m ( 0 ) 0 when 0 Ω by Lemma 16(i).

Now we assume that η 1 = 1 . If the origin is a focus, from expansion (2), we know that

Π ( r 0 ) = r 0 + η r 0 ν log j ( r 0 ) + o ( r 0 ν log j ( r 0 ) )

with η 0 , ν > 1 , and j N { 0 } . Putting this expression of Π and the Puiseux series (12) of V into (28) gives

i 0 v m + i ( 0 ) ( r 0 + η r 0 ν log j ( r 0 ) + o ( r 0 ν log j ( r 0 ) ) ) ( m + i ) / n = i 0 v m + i ( 0 ) r 0 ( m + i ) / n ( 1 + η ν r 0 ν 1 log j ( r 0 ) + o ( r 0 ν 1 log j ( r 0 ) ) ) .

The nonvanishing term of minimum order (which is chosen first with minimum exponent in r 0 and among them with the maximum exponent in log ( r 0 ) ) is r 0 m / n + ν 1 log j ( r 0 ) and equating their coefficients produces the following relations:

m n ν v m ( 0 ) η = 0

whose unique solution is ν = m / n and statement (ii) of the lemma follows.

Finally, part (iii) is just a direct consequence of (i) and (ii) because by the reciprocal of (ii), if m n , then either the origin is a center or a focus (with η 1 1 ) and m = n by (i).□

6.3 Proof of Theorem 7

Proof

First, if 0 is a characteristic direction, then we make a rotation of coordinates (or we simply may interchange the variables x and y ) and we can assume that 0 Ω . Recall that a change of coordinates may change also the Newton diagram of the transformed vector field.

The fact that inside the class Mo ( 1 , 1 ) one has η 1 ( λ ) = exp ( ξ ( λ ) ) follows easily just by integrating the first variational equation of equation (9) with respect to the invariant set { r = 0 } .

Taking into account that ( r , θ ) = r R k ( θ ) / F k ( θ ) + O ( r 2 ) and the expression (A2) of V , we obtain the Puiseux series

( r , θ ) V ( r , θ ) = r ( n m ) / n α 0 g α ( θ ) r α / n

with leading coefficient

g 0 ( θ ) = R k ( θ ) F k ( θ ) v m ( θ ) .

We are under the hypothesis of Theorem 15 so that the function G ( r 0 ) defined in (24) is

G ( r 0 ) = r 0 ( n m ) / n 0 2 π i 0 g i ( θ ) r 0 i / n d θ .

Without any assumption on uniform convergence of the involved series, we cannot integrate term by term the above series, but we will do it only for the first terms associated to the nonpositive powers of r 0 . Recalling that m n ,

(29) G ( r 0 ) = r 0 ( n m ) / n i = 0 m n ξ i r 0 i / n + G 1 ( r 0 ) ,

where G 1 ( 0 ) = 0 . Statement (i) follows recalling that lim r 0 0 + G ( r 0 ) = 0 is a necessary center condition.

On the other hand, by (25), the function G ( r 0 ) can be expressed as follows:

(30) G ( r 0 ) = r 0 Π ( r 0 ) d r V ( r , 0 ) = r 0 Π ( r 0 ) d r i m v i ( 0 ) r i / n = r 0 Π ( r 0 ) i m v ˜ i r i / n d r ,

where v ˜ m = 1 / v m ( 0 ) and v m ( 0 ) 0 by Lemma 16(i) and the condition 0 Ω . We may split G into two parts

(31) G ( r 0 ) = i = m n v ˜ i α ( r 0 , i / n ) + G 2 ( r 0 ) ,

where we have defined

α ( r 0 , γ ) = r 0 Π ( r 0 ) r γ d r

which satisfies α ( 0 , γ ) = 0 if γ < 1 . Therefore, G 2 ( 0 ) = 0 because Π ( 0 ) = 0 .

Taking into account that Π ( r 0 ) = η 1 r 0 + o ( r 0 ) , we obtain

(32) α ( r 0 , 1 ) = r 0 Π ( r 0 ) r 1 d r = log Π ( r 0 ) r 0 = log ( η 1 + o ( r 0 ) / r 0 ) .

We continue assuming that m = n . In this particular case, (29) becomes G ( r 0 ) = ξ 0 + G 1 ( r 0 ) , and (30) is transformed into

G ( r 0 ) = v ˜ n α ( r 0 , 1 ) + G 2 ( r 0 ) .

Comparing the value of lim r 0 0 + G ( r 0 ) in the two former expressions of G ( r 0 ) and using (32) leads to ξ 0 = 1 v m ( 0 ) log ( η 1 ) , that is,

η 1 = exp ( v m ( 0 ) ξ 0 ) .

Recall now that v m ( θ ) = v m ( 0 ) is a constant for all θ S 1 by Lemma 16(iii). Then the relation η 1 = exp ( ξ ) is confirmed again.

From now, we take m > n , and hence, η 1 = 1 by the converse of Theorem 5(i). We need to compute the order at r 0 = 0 of the integrals α ( r 0 , i / n ) appearing in (31). So we are going to calculate α ( r 0 , γ ) with γ { m i n : i = 0 , , m n } , hence with γ 1 . When γ = 1 , we already know by (32) that

α ( r 0 , 1 ) = log ( 1 + o ( r 0 ) / r 0 ) ,

and therefore,

lim r 0 0 + α ( r 0 , 1 ) = 0 ,

while if 1 < γ < m / n , then

α ( r 0 , γ ) = ( Π ( r 0 ) ) 1 γ r 0 1 γ 1 γ = 1 1 γ 1 ( r 0 + o ( r 0 ) ) γ 1 1 r 0 γ 1 = 1 1 γ 1 r 0 γ 1 1 1 + ( γ 1 ) r 0 1 o ( r 0 ) + o ( r 0 1 o ( r 0 ) ) 1 = 1 1 γ 1 r 0 γ 1 ( ( γ 1 ) r 0 1 o ( r 0 ) + o ( r 0 1 o ( r 0 ) ) ) .

Using Theorem 5(ii) we have indeed that, in the previous expansion, o ( r 0 ) = η r 0 m / n log j ( r 0 ) + o ( r 0 m / n log j ( r 0 ) ) . Therefore,

α ( r 0 , γ ) = η r 0 m / n γ log j ( r 0 ) + o ( r 0 m / n γ log j ( r 0 ) )

and

lim r 0 0 + α ( r 0 , γ ) = 0

for the range 1 < γ < m / n .

In summary, assuming that ξ 0 = ξ 1 = = ξ m n = 0 , equation (29) reduces to G ( r 0 ) = G 1 ( r 0 ) , and hence, lim r 0 0 + G ( r 0 ) = 0 . In addition, taking the limit when r 0 goes to 0 in the expression (31) produces that

lim r 0 0 + G ( r 0 ) = v ˜ m lim r 0 0 + α ( r 0 , m / n ) .

By using the expression (14), that is, Π ( r 0 ) = r 0 + P 0 ( log r 0 ) r 0 m / n + o ( r 0 m / n ) , we obtain that

lim r 0 0 + G ( r 0 ) = v ˜ m lim r 0 0 + P 0 ( log r 0 ) ,

which is zero only when P 0 ( log r 0 ) 0 , and hence, by Corollary 6, the origin is a center of system (4) proving statement (ii).□

Remark 20

We are going to explain why the Mo ( 1 , 1 ) assumption cannot be removed from the hypothesis of Theorem 7. If X Mo ( 1 , 1 ) , then the ξ i are improper integrals than perhaps cannot do not exist. We claim that even if the Cauchy principal values of the ξ i exists, equation (29) is wrong. This is because we have

G ( r 0 ) = r 0 ( n m ) / n 0 2 π i 0 g i ( θ ) r 0 i / n d θ = r 0 ( n m ) / n lim ε 0 + I ε i 0 g i ( θ ) r 0 i / n d θ ,

where I ε = [ 0 , 2 π ] \ J ε with J ε = i = 1 ( θ i ε , θ i + ε ) and Ω = { θ 1 , , θ } . But we cannot split the last limit into the sum of two limits as follows:

G ( r 0 ) = r 0 ( n m ) / n i = 0 m n ξ i r 0 i / n + lim ε 0 + I ε i m n + 1 g i ( θ ) r 0 i / n d θ

because we do not know if the second one exists.

6.4 Proof of Theorem 14

Proof

Since F ( θ , r , v ) = 0 is an invariant surface of the characteristic vector field Z , we know that

(33) Z ( F ) F = 0 = 0 .

We calculate

Z ( F ) ( θ , r , v ) = Θ ( θ , r ) F θ ( θ , r , v ) + ( θ , r ) F r ( θ , r , v ) + v D ˆ ( θ , r ) F v ( θ , r , v )

and evaluate it on v = f ( θ , r ) . Since f ( θ , r ) is a branch of F = 0 , by condition (33), we know that Z ( F ) ( θ , r , f ( θ , r ) ) 0 so that

(34) 0 = Θ ( θ , r ) F θ ( θ , r , f ( θ , r ) ) + ( θ , r ) F r ( θ , r , f ( θ , r ) ) + f ( θ , r ) D ˆ ( θ , r ) F v ( θ , r , f ( θ , r ) ) .

On the other hand, since f C 1 , taking implicit derivatives in the identity F ( θ , r , f ( θ , r ) ) = 0 and using the chain rule yield

F θ ( θ , r , f ( θ , r ) ) + F v ( θ , r , f ( θ , r ) ) f θ ( θ , r ) = 0 , F r ( θ , r , f ( θ , r ) ) + F v ( θ , r , f ( θ , r ) ) f r ( θ , r ) = 0 .

Introducing these relations into (34) gives X ˆ ( f ) = f D ˆ , and hence, f is an inverse integrating factor of X ˆ .

It only remains to prove that f can be expressed as a convergent Puiseux series in the variable r . We take the bifurcating point ( r , v ) = ( 0 , v ) . Since F ( θ , r , v ) is analytic with F ( θ , 0 , v ) = 0 , by Newton-Puiseux theorem, see [9], for instance, there exists a finite factorization

F ( θ , r , v ) = U ( θ , r , v ) r m ( v f ( θ , r ) ) κ i ( v f i ( θ , r ) ) κ i ,

where U is an analytic unit U ( θ , 0 , v ) 0 for any θ S 1 , m 0 is natural, and the branches f and f i ( θ , . ) are convergent Puiseux series with index n 1 (that is analytic functions of r 1 / n ) with f ( θ , 0 ) = f i ( θ , 0 ) = v , and the multiplicities κ and κ i are some positive integers. Now the proof is done.□

7 Examples

7.1 Example 1 (nilpotent in Mo ( p , q ) )

We solve the center-focus problem at the origin for the full family

(35) x ˙ = y + α x 2 + x ( a x 2 + b y ) , y ˙ = 2 α x y + β x 3 + 2 y ( a x 2 + b y ) ,

with parameters ( α , β , a , b ) R 4 .

Proposition 21

The origin of family (35) is monodromic if and only if the parameters lie in

(36) Λ = { ( α , β , a , b ) R 4 : 2 α 2 + β < 0 } .

Moreover, the origin is a center only when additionally

(37) a α b = 0 .

Proof

The origin is a monodromic nilpotent singularity with Andreev number 2 (see [6,14] for a definition) when the parameters lie in (36). Also it can be shown using the algorithm proposed in [14] that (37) is a necessary center condition. We are going to see using our theory that indeed it is also a sufficient center condition.

The vector field (35) has a decomposition in ( p , q ) -quasihomogeneous vector fields (16) with weights ( p , q ) = ( 1 , 2 ) with leading part X 1 = ( y + α x 2 , 2 α x y + β x 3 ) . Since the origin is monodromic for X 1 under condition (36), it follows that the family (35) lies in Mo ( 1 , 2 ) .

Thus, we perform the weighted polar blow-up (17) with these weights whose Jacobian is J ( φ , ρ ) = ρ 2 ( 3 + cos ( 2 φ ) ) / 2 and divide the resulting vector field by ρ to obtain a system (18) of the form ρ ˙ = F ˆ 1 ( φ ) ρ , φ ˙ = G ˆ 1 ( φ ) with G ˆ 1 ( φ ) 0 on all S 1 .

Some computations give

ξ ˆ = 0 2 π ˆ 1 ( φ ) d φ = 1 4 0 2 π B ( φ ) B ( φ ) d φ = 0 ,

where B ( φ ) = 8 3 β 4 ( 2 + β ) cos ( 2 φ ) β cos ( 4 φ ) + 8 α sin ( φ ) + 8 α sin ( 3 φ ) 0 in all S 1 and the prime indicates derivative with respect to φ .

The system exhibits the inverse integrating factor

v ( x , y ) = ( β x 4 2 y ( 2 α x 2 + y ) ) 5 / 4 ,

so that the function V ( φ , ρ ) = v ( ρ cos φ , ρ 2 sin φ ) / ( ρ J ( φ , ρ ) φ ˙ ) is an inverse integrating factor of the equation for the phase orbits (19) having the form V ( φ , ρ ) = A ( φ ) ρ 2 , hence of Puiseux type with multiplicity m = 2 n and arbitrary index n 1 . We continue taking the minimum values ( m , n ) = ( 2 , 1 ) .

Expanding ˆ ( φ , ρ ) / V ( φ , ρ ) at ρ = 0 , we obtain the Puiseux series ρ 1 i 0 g ˆ i ( φ ) ρ i , and we compute the quantities

ξ ˆ 0 = 0 2 π g ˆ 0 ( φ ) d φ = 0 , ξ ˆ 1 = 0 2 π g ˆ 1 ( φ ) d φ = 0 2 π ( 3 cos ( 2 φ ) ) ( a + a cos ( 2 φ ) + 2 b sin ( φ ) ) ( B ( φ ) ) 5 / 4 d φ .

We are unable to do the quadrature ξ ˆ 1 ( α , β , a , b ) , but we can compute its value when the parameters satisfy (37):

ξ ˆ 1 ( α , β , α b , b ) = 2 3 / 4 b cos ( φ ) ( B ( φ ) ) 1 / 4 0 2 π = 0 .

Therefore, (37) is not only a necessary center condition but also a sufficient one by the generalization of Theorem 7(ii) using Remark 9.

Recall that in this example, S 12 \ { ρ = 0 } = by construction since we are under monodromic conditions and φ ˙ is a function only of φ and does not depend on ρ . Therefore, Λ = Λ 12 = .□

7.2 Example 2 (not in Mo ( p , q ) )

We consider the family of vector fields

(38) x ˙ = λ 1 ( x 6 + 3 y 2 ) ( y + μ x ) + λ 2 ( x 2 + y 2 ) ( y + A x 3 ) , y ˙ = λ 1 ( x 6 + 3 y 2 ) ( x + μ y ) + λ 2 ( x 2 + y 2 ) ( x 5 + 3 A x 2 y ) ,

with parameter space ( λ 1 , λ 2 , μ , A ) R 4 . The origin is a degenerate singularity with characteristic directions Ω = { 0 , π } , and there are members of the family with both the origin monodromic and not monodromic. The family has the inverse integrating factor v ( x , y ) = ( x 2 + y 2 ) ( x 6 + 3 y 2 ) .

The Newton diagram of (38) consists of two exterior vertices at ( 0 , 4 ) and ( 8 , 0 ) provided λ 2 3 λ 1 0 and λ 1 λ 2 0 , respectively, and an inner vertex at ( 2 , 2 ) if λ 1 2 + λ 2 2 0 . Therefore, the allowed weights ( p , q ) in (16) are either ( p , q ) = ( 1 , 1 ) or ( p , q ) = ( 1 , 3 ) whose leadings parts are X 2 = x + λ 1 3 y 2 ( x + y μ ) y and X 4 = λ 2 x 2 ( A x 3 + y ) x + y , respectively. Consequently, (38) is not in Mo ( p , q ) since the leading parts have invariant lines through the origin.

The solution of the center-focus problem for the full family (38) in Λ \ Λ is presented below.

Proposition 22

The origin of family (38) is monodromic if and only if the parameters lie in

(39) Λ = { ( λ 1 , λ 2 , μ , A ) R 4 : 3 λ 1 λ 2 > 0 , λ 1 λ 2 > 0 } .

We additionally restrict to Λ \ Λ , where Λ = Λ 11 Λ 13 and is explicitly given by

(40) Λ = { α 11 0 , α 13 0 , A λ 2 + α 11 0 , μ λ 1 + α 13 0 } ,

where α 11 3 λ 1 2 + 4 λ 1 λ 2 + ( 1 + A 2 ) λ 2 2 and α 13 ( 3 + μ 2 ) λ 1 2 + 4 λ 1 λ 2 λ 2 2 . Under these parameter constrains the following holds:

  1. The Poincaré map is linear: Π ( r 0 ) = η 1 r 0 ;

  2. The origin is a center if and only if

    (41) 3 λ 1 μ + 3 A λ 2 = 0 .

Proof

The necessary monodromic conditions in Remark 10 for the origin of (38) are given by 3 λ 1 λ 2 > 0 and 3 λ 1 2 λ 2 > 0 . We also notice that (38) does not belong to the class S k ω , see Remark 11. Indeed, we can perform the monodromy algorithm developed in [4], and we can check that the necessary and sufficient monodromic conditions for the origin of (38) are summarized in (39).

Since W ( N ( X ) ) = { ( 1 , 1 ) , ( 1 , 3 ) } , we do an analysis in each edge of N ( X ) .□

7.2.1 Analysis with weights ( p , q ) = ( 1 , 1 ) in Λ \ Λ 11

According to (11), the inverse integrating factor v ( x , y ) of (38) is transformed into

V ( θ , r ) = v ( r cos θ , r sin θ ) r 3 Θ ( θ , r ) = r ( 3 sin 2 θ + r 4 cos 6 θ ) ( 3 λ 1 λ 2 ) sin 2 θ + 2 A λ 2 r 2 cos 3 θ sin θ + ( λ 1 λ 2 ) r 4 cos 6 θ ,

which is a Puiseux inverse integrating factor with m = n = 1 .

Since V 1 ( 0 ) \ { r = 0 } = , Theorem 15 can be applied. Therefore, the function G ( r ) defined in (24), that is,

(42) G ( r ) = 0 2 π ( θ , r ) V ( θ , r ) d θ

is well defined in a sufficiently small positive half-neighborhood of the origin, and moreover, the origin is a center of X if and only if G ( r ) 0 .

On the other hand, in the parameter space Λ \ Λ 11 , the fundamental equation (28) holds, that is, V ( 0 , Π ( r ) ) = Π ( r ) V ( 0 , r ) . In our case, this equation is Π ( r ) r Π ( r ) = 0 , whose general solution is Π ( r ) = η 1 r proving statement (i). This result is in agreement with formula (25), which in our examples is

G ( r ) = r Π ( r ) d r V ( 0 , r ) = ( λ 1 λ 2 ) r η 1 r d r r = ( λ 1 λ 2 ) log η 1 .

Now we will compute G ( r ) from its definition (42). To this end, we rewrite

( θ , r ) V ( θ , r ) = λ 1 μ + A λ 2 r 2 I ( θ , r ) + J ( θ , r ) ,

where 0 2 π J ( θ , r ) d θ 0 because the function J possesses the symmetry J ( θ , r ) + J ( 2 π θ , r ) 0 . Here, I ( θ , r ) has the expression

I ( θ , r ) = cos 2 θ ( 2 cos ( 2 θ ) ) r 4 cos 6 θ + 3 sin 2 θ .

Defining I 11 ( r ) = 0 2 π I ( θ , r ) d θ we obtain G ( r ) = 2 π λ 1 μ + A λ 2 r 2 I 11 ( r ) . Doing the change of variables θ z with z = tan θ , we can compete I 11 ( r ) whose value is I 11 ( r ) = 2 3 π / ( 3 r 2 ) , and hence,

G ( r ) = 2 π λ 1 μ + 2 3 π A λ 2 / 3 .

By identifying both expressions of G ( r ) , we obtain that log η 1 = 0 if and only if 3 λ 1 μ + 3 A λ 2 = 0 . The proof is finished.

7.2.2 Analysis with weights ( p , q ) = ( 1 , 3 ) in Λ \ Λ 13

We do not enter in the details, we only comment that the linear expression of Π is obtained easily, but we are not able to do the necessary quadratures to obtain the function G ( ρ ) .

Remark 23

It is interesting to note that if instead to restrict family (39) to Λ \ Λ as we did in Proposition 22, we now restrict (39) to the set Λ M Λ defined by Λ M = Λ { 9 λ 1 2 λ 2 6 λ 1 λ 2 > 0 } , then family (39) falls into the subclass of monodromic singularities defined by Mañosa in Corollary 12 of [27]. In this monodromic class, the quantity η 1 was computed in [27], and applying that formula to family (39) gives

log η 1 = 2 π 3 λ 1 λ 2 ( 3 μ λ 1 + 3 A λ 2 ) ,

in complete agreement with the expression of η 1 obtained in Proposition 22. Recall that, due to different parameterizations of the transversal section, where Π is defined, the explicit expression of η 1 can vary up to a positive multiplicative constant. This computation suggests that perhaps Proposition 22 is true in all Λ and not only in Λ \ Λ .

Remark 24

In the statement of Proposition 22, we need to impose a parameter restriction that guarantees (8) and (20) with ( p , q ) = ( 1 , 3 ) . Actually the existence of the critical set depends on the weights ( p , q ) as follows:

  1. With the weights ( p , q ) = ( 1 , 1 ) the equation Θ ( θ , r ) = 0 is biquadratic in r , hence easy to analyze the characterization of the critical set S : it satisfies (8) if and only if either

    α 11 3 λ 1 2 + 4 λ 1 λ 2 + ( 1 + A 2 ) λ 2 2 < 0

    or

    α 11 0 and A λ 2 + α 11 λ 2 λ 1 < 0 .

  2. Analogously, with the weights ( p , q ) = ( 1 , 3 ) , we obtain that (20) is characterized by either

    α 13 ( 3 + μ 2 ) λ 1 2 + 4 λ 1 λ 2 λ 2 2 < 0

    or

    α 13 0 and μ λ 1 + α 13 3 λ 1 λ 2 < 0 .

In particular, joining the above parts (i), (ii), and the monodromic conditions (39) of Proposition 22, it follows that the subset of the parameter space of family (38) that has the origin monodromic and such that S \ { r = 0 } and S 13 \ { r = 0 } holds simultaneously is not empty. Actually it is characterized by (39) and (40).

7.3 Example 3 (not in Mo ( p , q ) )

We consider the family of vector fields

(43) x ˙ = λ 1 y ( C x 6 6 A x 3 y 3 B y 2 ) + λ 2 ( x 2 + y 2 ) ( A x 3 + B y ) , y ˙ = λ 1 x ( C x 6 6 A x 3 y 3 B y 2 ) + λ 2 ( x 2 + y 2 ) ( C x 5 3 A x 2 y ) ,

with parameter space { ( λ 1 , λ 2 , A , B , C ) R 5 } . The origin is a degenerate singularity of the family with nonmonodromic and monodromic members whose characteristic directions are Ω = { 0 , π } .

The Newton diagram of (43) is made by two exterior vertices at ( 0 , 4 ) and ( 8 , 0 ) if B ( 3 λ 1 + λ 2 ) 0 and C ( λ 1 + λ 2 ) 0 , respectively, and an inner vertex at ( 2 , 2 ) if B ( λ 1 2 + λ 2 2 ) 0 . Therefore, the weights ( p , q ) of (43) are W ( N ( X ) ) = { ( 1 , 1 ) , ( 1 , 3 ) } . The leadings parts are X 2 = x 3 B λ 1 x y 2 y and X 4 = λ 2 x 2 ( A x 3 + B y ) x + y with weights ( 1 , 1 ) and ( 1 , 3 ) , respectively. The leading parts have invariant lines through the origin, that’s why family (43) lies off Mo ( p , q ) .

Family (43) possesses the polynomial inverse integrating factor v ( x , y ) = ( x 2 + y 2 ) ( C x 6 6 A x 3 y 3 B y 2 ) . Nevertheless, from the first integral of (43) computed through v ( x , y ) , we can find other inverse integrating factors like, for example,

(44) v ( x , y ) = ( C x 6 6 A x 3 y 3 B y 2 ) 1 λ 2 3 λ 1

when λ 1 0 . In the next proposition, we will use (44) to analyze the center-focus problem for family (43) in Λ Λ ˆ and ( Λ \ Λ ) Λ ˆ , where

Λ ˆ = { ( λ 1 , λ 2 , A , B , C ) R 5 : ( 3 λ 1 + 2 λ 2 , λ 1 ) Z × N } .

Proposition 25

The origin of family (43) is monodromic if and only if the parameters lie in

(45) Λ = { ( λ 1 , λ 2 , A , B , C ) R 5 : C ( 3 λ 1 + λ 2 ) > 0 , C ( 3 λ 1 + 2 λ 2 ) > 0 , B ( 3 λ 1 + λ 2 ) < 0 , B ( 3 λ 1 + 2 λ 2 ) < 0 , 3 A 2 + B C < 0 } .

Moreover, if we restrict to the subset of the parameter space Λ \ Λ , then the origin is a center, where Λ is

(46) Λ = { α 11 0 , λ 1 + λ 2 > 0 , 3 λ 1 + λ 2 < 0 } ,

and α 11 B C ( λ 1 + λ 2 ) ( 3 λ 1 + λ 2 ) + A 2 ( 3 λ 1 + 2 λ 2 ) 2 .

Proof

The necessary monodromic conditions in Remark 10 for the origin of (43) are B ( 3 λ 1 + λ 2 ) < 0 and B ( 3 λ 1 + 2 λ 2 ) 0 . Also we can see that (43) is not in the class S k ω , see Remark 11. Actually, by using the monodromy algorithm presented in [4], we know that the origin of (43) is monodromic if and only if its parameters belong to the set (45).□

Wa analyze the example in each weight of W ( N ( X ) ) = { ( 1 , 1 ) , ( 1 , 3 ) } .

7.3.1 Analysis with weights ( p , q ) = ( 1 , 1 ) in Λ \ Λ 11

We transform the inverse integrating factor v ( x , y ) of (43) using (11), and we obtain

V ( θ , r ) = v ( r cos θ , r sin θ ) r 3 Θ ( θ , r ) = r ( 3 B sin 2 θ 6 A r 2 cos 3 θ sin θ + C r 4 cos 6 θ ) C r 4 λ 1 cos 6 θ + C r 4 λ 2 cos 6 θ 6 A r 2 λ 1 cos 3 θ sin θ 4 A r 2 λ 2 cos 3 θ sin θ 3 B λ 1 sin 2 θ B λ 2 sin 2 θ ,

which is a Puiseux inverse integrating factor with m = n = 1 .

Since V 1 ( 0 ) \ { r = 0 } = we are going to apply Theorem 15. The function G ( r ) defined in (24) for all r > 0 sufficiently small is

G ( r ) = 0 2 π cos θ ( A r 2 cos 3 θ + ( B + C r 4 cos 4 θ ) sin θ 3 A r 2 cos θ sin 2 θ ) C r 4 cos 6 θ 6 A r 2 cos 3 θ sin θ 3 B sin 2 θ d θ = λ 2 ( P ( 2 π ; r ) P ( 0 ; r ) ) 0 ,

where the primitive P is

P ( θ ; r ) = 1 6 log ( 48 B + 10 C r 4 + 3 ( 16 B + 5 C r 4 ) cos ( 2 θ ) + C r 4 ( 6 cos ( 4 θ ) + cos ( 6 θ ) ) 192 A r 2 cos 3 θ sin θ ) .

Therefore, the origin is a center finishing thee proof.

7.3.2 Analysis with weights ( p , q ) = ( 1 , 3 ) in Λ \ Λ 13

We do not display the explicit computations, but we obtain the same results as in the previous section.

Remark 26

The existence of the critical sets S p q for all ( p , q ) W ( N ( X ) ) is characterized below:

  1. With the weights ( p , q ) = ( 1 , 1 ) the equation Θ ( θ , r ) = 0 is biquadratic in r . The subset Λ 11 is characterized by the inequality α 11 B C ( λ 1 + λ 2 ) ( 3 λ 1 + λ 2 ) + A 2 ( 3 λ 1 + 2 λ 2 ) 2 0 .

  2. Similarly, with the weights ( p , q ) = ( 1 , 3 ) , we obtain that Λ 13 = { λ 1 + λ 2 > 0 , 3 λ 1 + λ 2 < 0 }

By Joining aforementioned (i) and (ii), we obtain (46).

Acknowledgements

The authors are grateful to the referee for his/her valuable comments and suggestions to improve this paper. The authors are partially supported by the Agencia Estatal de Investigación grant PID2020-113758GB-I00 and an AGAUR (Generalitat de Catalunya) grant number 2021SGR 01618.

  1. Funding information: The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: The authors reports financial support was provided by Spain Ministry of Science and Innovation. The authors reports financial support was provided by Government of Catalonia Agency for Administration of University and Research Grants.

  2. Conflict of interest: Authors state no conflict of interest.

Appendix

Here, we give some technical details that are implicit throughout this work.

A.1 Inverse integrating factors in polar coordinates

Let V ˆ ( θ , r ) be an inverse integrating factor in C \ { r = 0 } of the vector field X ˆ = ( θ , r ) r + Θ ( θ , r ) θ associated to system (5), that is, V ˆ is a not locally null real C 1 ( C \ { r = 0 } ) function that is a solution of the linear partial differential equation X ˆ ( V ˆ ) = V ˆ div ( X ˆ ) . Then

(A1) V ( θ , r ) = V ˆ ( θ , r ) Θ ( θ , r )

is an inverse integrating factor of (9) in C \ { S { r = 0 } } .

Clearly, the zero-set V 1 ( 0 ) is an invariant set for (9). Moreover, if ( θ 0 , r 0 ) V 1 ( 0 ) , then all the orbit of (9) through the point ( θ 0 , r 0 ) lies in V 1 ( 0 ) . If the origin is not monodromic, it may happen that V 1 ( 0 ) contains some separatrices of singularities ( θ , r ) = ( θ , 0 ) with θ Ω . The aforementioned situation is clearly forbidden in the monodromic scenario when the parameters lie in Λ .

We consider the differential 1-form ω = d r ( r , θ ) d θ associated to (9) restricted to Λ . Then from (A1),

ω V = d r ( r , θ ) d θ V ( θ , r ) = Θ ( θ , r ) d r ( θ , r ) d θ V ˆ ( θ , r )

is a differential 1-form on C \ { V 1 ( 0 ) } , where

V 1 ( 0 ) = { ( θ , r ) C : V ( θ , r ) = 0 } = V ˆ 1 ( 0 ) .

holds in the parameter space Λ .

Remark 27

It may happen that the origin is monodromic and the invariant set { r = 0 } is a nonisolated component of V 1 ( 0 ) because there is a sequence of components of V 1 ( 0 ) accumulating on { r = 0 } . The example of this kind we are going to explain is presented in [8]. System

x ˙ = y + x ( x 2 + y 2 ) , y ˙ = x + y ( x 2 + y 2 ) ,

has a nondegenerate focus at the origin, hence without characteristic directions. The corresponding equation (9) is d r / d θ = r 3 , and possesses the inverse integrating factor V ( θ , r ) = r 3 sin ( 2 θ + r 2 ) given in (13) that is C 1 ( C \ { r = 0 } ) . Moreover, V 1 ( 0 ) = { ( θ , r ) : 2 θ + r 2 = k π , k Z } { r = 0 } .

A.2 Puiseux inverse integrating factors

We say that an inverse integrating factor V ˆ ( r , θ ) in C \ { r = 0 } of the polar system (5) is a Puiseux inverse integrating factor if it can be expanded in convergent Puiseux series about r = 0 of the form

(A2) V ˆ ( r , θ ) = i m v ˆ i ( θ ) r i / n

for some positive integer n called the index, m Z , and C 1 coefficient functions v ˆ i : S 1 R with leading coefficient v ˆ m ( θ ) 0 . Clearly, if V ˆ is a Puiseux inverse integrating factor in C \ { r = 0 } of (5), then the function V given by (A1) is also a Puiseux inverse integrating factor of (9) in C \ { S { r = 0 } } with the same index n , that is,

V ( r , θ ) = i m v i ( θ ) r i / n

whose coefficients are C 1 functions v i : S 1 \ Ω R and the leading coefficient is v m ( θ ) = v ˆ m ( θ ) / F k ( θ ) .

References

[1] A. Algaba, N. Fuentes, and C. García, Centers of quasi-homogeneous polynomial planar systems, Nonlinear Anal. Real World Appl. 13 (2012), 419–431. 10.1016/j.nonrwa.2011.07.056Search in Google Scholar

[2] A. Algaba, C. García, and J. Giné, Analytic integrability for some degenerate planar vector fields, J. Differential Equations 257 (2014), 549–565. 10.1016/j.jde.2014.04.010Search in Google Scholar

[3] A. Algaba, C. García, and J. Giné, Geometric criterium in the center problem, Mediterr. J. Math. 13 (2016), 2593–2611. 10.1007/s00009-015-0641-0Search in Google Scholar

[4] A. Algaba, C. García, and M. Reyes, Characterization of a monodromic singular point of a planar vector field, Nonlinear Anal. 74 (2011), 5402–5414. 10.1016/j.na.2011.05.023Search in Google Scholar

[5] A. Algaba, C. García, and M. Reyes, A new algorithm for determining the monodromy of a planar differential system, Appl. Math. Comput. 237 (2014), 419–429. 10.1016/j.amc.2014.03.129Search in Google Scholar

[6] A. Andreev, Investigation on the behavior of the integral curves of a system of two differential equationsin the neighborhood of a singular point, Transl. Amer. Math. Soc. 8 (1958), 187–207. 10.1090/trans2/008/07Search in Google Scholar

[7] V. I. Arnold, Chapitres Supplémentaires de la Théorie des Équations Différentielles Ordinaires, Editions Mir, Moscow, 1980. Search in Google Scholar

[8] L. R. Berrone and H. Giacomini, On the vanishing set of inverse integrating factors, Qual. Theory Dyn. Syst. 1 (2000), 211–230. 10.1007/BF02969478Search in Google Scholar

[9] E. Brieskorn and H. Knörrer, Plane Algebraic Curves, Translated from the German original by John Stillwell. [2012] reprint of the 1986 edition. Modern Birkhäuser Classics. Birkhäuser/Springer Basel AG, Basel, 1986. Search in Google Scholar

[10] F. Dumortier, Singularities of vector fields on the plane, J. Differential Equations 23 (1977), 53–106. 10.1016/0022-0396(77)90136-XSearch in Google Scholar

[11] J. Écalle, Introduction auxfonctions analysables et preuve constructive de la conjecture de Dulac. Actualités Mathématiques. Hermann, Paris, 1992. Search in Google Scholar

[12] W. W. Farr, C. Li, I. S. Labouriau, and W. F. Langford, Degenerate Hopf-bifurcation formulas and Hilbertas 16th problem, SIAM J. Math. Anal. 20 (1989), 13–29. 10.1137/0520002Search in Google Scholar

[13] W. Fulton, Algebraic topology. A first course. Graduate Texts in Mathematics, vol. 153, Springer-Verlag, NewYork, 1995. 10.1007/978-1-4612-4180-5_11Search in Google Scholar

[14] I. A. García, Cyclicity of nilpotent centers with minimum Andreev number, Eur. J. Math. 5 (2019), 1293–1330. 10.1007/s40879-018-0304-3Search in Google Scholar

[15] I. A. García, H. Giacomini, and M. Grau, The inverse integrating factor and the Poincaré map, Trans. Amer. Math. Soc. 362 (2010), 3591–3612. 10.1090/S0002-9947-10-05014-2Search in Google Scholar

[16] I. A. García and J. Giné, Center problem with characteristic directions and inverse integrating factors, Commun. Nonlinear Sci. Numer. Simul. 108 (2022), 14 pp. 10.1016/j.cnsns.2022.106276Search in Google Scholar

[17] I. A. García, J. Giné, and M. Grau, A necessary condition in the monodromy problem for analyticdifferential qquations on the plane, J. Symbolic Comput. 41 (2006)943–958. 10.1016/j.jsc.2006.04.007Search in Google Scholar

[18] I. A. García and S. Maza, A new approach to center conditions for simple analytic monodromic singularities, J. Differential Equations 248 (2010), 363–380. 10.1016/j.jde.2009.09.002Search in Google Scholar

[19] A. Gasull, J. Llibre, V. Mañosa, and F. Mañosas, The focus-centre problem for a type of degenerate system, Nonlinearity 13 (2000), 699–729. 10.1088/0951-7715/13/3/311Search in Google Scholar

[20] A. Gasull, V. Mañosa, and F. Mañosas, Monodromy and stability of a class of degenerate planar critical points, J. Differential Equations 217 (2005), 363–376. 10.1016/j.jde.2005.06.022Search in Google Scholar

[21] H. Giacomini, J. Giné, and J. Llibre, The problem of distinguishing between a center and a focus for nilpotent and degenerate analytic systems, J. Differential Equations 227 (2006), no. 2, 406–426; Corrigendum, ibid. 232 (2007) 702. Search in Google Scholar

[22] J. Giné, Sufficient conditions for a center at a completely degenerate critical point, Internat. J. Bifur. Chaos Appl. Sci. Eng. 12 (2002), no. 7, 1659–1666. 10.1142/S0218127402005315Search in Google Scholar

[23] J. Giné, On the centers of planar analytic differential systems, Int. J. Bifur. Chaos Appl. Sci. Eng. 17 (2007), no. 9, 3061–3070. 10.1142/S0218127407018865Search in Google Scholar

[24] J. Giné, On the degenerate center problem, Int. J. Bifur. Chaos Appl. Sci. Eng. 21 (2011), no. 5, 1383–1392. 10.1142/S0218127411029082Search in Google Scholar

[25] J. Giné and S. Maza, The reversibility and the center problem, Nonlinear Anal. 74 (2011), no. 2, 695–704. 10.1016/j.na.2010.09.028Search in Google Scholar

[26] Yu. S Il’yashenko, Finiteness theorems for limit cycles, Translated Russian by H. H. McFaden, Translations of Mathematical Monographs, vol. 94, American Mathematical Society, Providence, RI, 1991. 10.1090/mmono/094Search in Google Scholar

[27] V. Mañosa, On the center problem for degenerate singular points of planar vector fields, Int. J. Bifur. Chaos Appl. Sci. Eng. 12 (2002), no. 4, 687–707. 10.1142/S0218127402004693Search in Google Scholar

[28] N. B. Medvedeva, The principal term of the asymptotic expansion of the monodromy transformation: Calculation in blowing-up geometry, Siberian Math. J. 38 (1997), 114–126. 10.1007/BF02674907Search in Google Scholar

[29] N. B. Medvedeva and E. Batcheva, The second term of the asymptotics of the monodromy map in case of two even edges of Newton diagram, Electron. J. Qual. Th. Diff. Eqs. 19 (2000), 1–15. Search in Google Scholar

[30] N. B. Medvedeva, On the analytic solvability of the problem of distinguishing between center and focus, Proc. Steklov Inst. Math. 254 (2006), 7–93. 10.1134/S0081543806030023Search in Google Scholar

[31] R. Moussu, Symétrie et forme normale des centres et foyers dégénérés, Ergodic Theory Dynam. Systems 2 (1982), 241–251. 10.1017/S0143385700001553Search in Google Scholar

[32] V. G. Romanovski and D. S. Shafer, The Center and Cyclicity Problems: A Computational Algebra Approach, Birkhäuser Boston, Inc., Boston, MA, 2009. Search in Google Scholar

Received: 2021-09-15
Revised: 2022-11-22
Accepted: 2023-03-13
Published Online: 2023-05-25

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. On sufficient “local” conditions for existence results to generalized p(.)-Laplace equations involving critical growth
  3. On the critical Choquard-Kirchhoff problem on the Heisenberg group
  4. On the local behavior of local weak solutions to some singular anisotropic elliptic equations
  5. Viscosity method for random homogenization of fully nonlinear elliptic equations with highly oscillating obstacles
  6. Double-phase parabolic equations with variable growth and nonlinear sources
  7. Logistic damping effect in chemotaxis models with density-suppressed motility
  8. Bifurcation diagrams of one-dimensional Kirchhoff-type equations
  9. Standing wave solution for the generalized Jackiw-Pi model
  10. Weak-strong uniqueness and energy-variational solutions for a class of viscoelastoplastic fluid models
  11. Eigenvalue inequalities for the buckling problem of the drifting Laplacian of arbitrary order
  12. Homoclinic solutions for a differential inclusion system involving the p(t)-Laplacian
  13. Equivalence between a time-fractional and an integer-order gradient flow: The memory effect reflected in the energy
  14. Bautin bifurcation with additive noise
  15. Small solitons and multisolitons in the generalized Davey-Stewartson system
  16. Nonstationary Poiseuille flow of a non-Newtonian fluid with the shear rate-dependent viscosity
  17. A global compactness result with applications to a Hardy-Sobolev critical elliptic system involving coupled perturbation terms
  18. On a strongly damped semilinear wave equation with time-varying source and singular dissipation
  19. Singularity for a nonlinear degenerate hyperbolic-parabolic coupled system arising from nematic liquid crystals
  20. Stability of stationary solutions to the three-dimensional Navier-Stokes equations with surface tension
  21. Uniform decay estimates for the semi-linear wave equation with locally distributed mixed-type damping via arbitrary local viscoelastic versus frictional dissipative effects
  22. Symmetry and nonsymmetry of minimal action sign-changing solutions for the Choquard system
  23. Periodic and quasi-periodic solutions of a four-dimensional singular differential system describing the motion of vortices
  24. Multiple nontrivial solutions of superlinear fractional Laplace equations without (AR) condition
  25. Existence and blow-up of solutions in Hénon-type heat equation with exponential nonlinearity
  26. Existence and concentration behavior of positive solutions to Schrödinger-Poisson-Slater equations
  27. On the fractional Korn inequality in bounded domains: Counterexamples to the case ps < 1
  28. Gradient estimates for nonlinear elliptic equations involving the Witten Laplacian on smooth metric measure spaces and implications
  29. Dynamical analysis of a reaction–diffusion mosquito-borne model in a spatially heterogeneous environment
  30. Existence and multiplicity of solutions for a quasilinear system with locally superlinear condition
  31. Nontrivial solution for Klein-Gordon equation coupled with Born-Infeld theory with critical growth
  32. Modeling Wolbachia infection frequency in mosquito populations via a continuous periodic switching model
  33. Infinitely many localized semiclassical states for nonlinear Kirchhoff-type equation
  34. Fujita-type theorems for a quasilinear parabolic differential inequality with weighted nonlocal source term
  35. Approximations of center manifolds for delay stochastic differential equations with additive noise
  36. Periodic solutions to a class of distributed delay differential equations via variational methods
  37. Groundstate for the Schrödinger-Poisson-Slater equation involving the Coulomb-Sobolev critical exponent
  38. Existence of nontrivial solutions for the Klein-Gordon-Maxwell system with Berestycki-Lions conditions
  39. Global Sobolev regular solution for Boussinesq system
  40. Normalized solutions for the p-Laplacian equation with a trapping potential
  41. Nonlinear elliptic–parabolic problem involving p-Dirichlet-to-Neumann operator with critical exponent
  42. Blow-up for compressible Euler system with space-dependent damping in 1-D
  43. High energy solutions of general Kirchhoff type equations without the Ambrosetti-Rabinowitz type condition
  44. On the dynamics of grounded shallow ice sheets: Modeling and analysis
  45. A survey on some vanishing viscosity limit results
  46. Blow-up for logarithmic viscoelastic equations with delay and acoustic boundary conditions
  47. Generalized Liouville theorem for viscosity solutions to a singular Monge-Ampère equation
  48. Front propagation in a double degenerate equation with delay
  49. Positive solutions for a class of singular (pq)-equations
  50. Higher integrability for anisotropic parabolic systems of p-Laplace type
  51. The Poincaré map of degenerate monodromic singularities with Puiseux inverse integrating factor
  52. On a system of multi-component Ginzburg-Landau vortices
  53. Existence and concentration of solutions to Kirchhoff-type equations in ℝ2 with steep potential well vanishing at infinity and exponential critical nonlinearities
  54. Multiplicity results for fractional Schrödinger-Kirchhoff systems involving critical nonlinearities
  55. On double phase Kirchhoff problems with singular nonlinearity
  56. Estimates for eigenvalues of the Neumann and Steklov problems
  57. Nodal solutions with a prescribed number of nodes for the Kirchhoff-type problem with an asymptotically cubic term
  58. Dirichlet problems involving the Hardy-Leray operators with multiple polars
  59. Incompressible limit for compressible viscoelastic flows with large velocity
  60. Characterizations for the genuine Calderón-Zygmund operators and commutators on generalized Orlicz-Morrey spaces
  61. Non-local gradients in bounded domains motivated by continuum mechanics: Fundamental theorem of calculus and embeddings
  62. Noncoercive parabolic obstacle problems
  63. Touchdown solutions in general MEMS models
  64. Existence and nonexistence of nontrivial solutions for the Schrödinger-Poisson system with zero mass potential
  65. Short-time existence of a quasi-stationary fluid–structure interaction problem for plaque growth
  66. Fine bounds for best constants of fractional subcritical Sobolev embeddings and applications to nonlocal PDEs
  67. Symmetries of Ricci flows
  68. Asymptotic behavior of solutions of a second-order nonlinear discrete equation of Emden-Fowler type
  69. On the topological gradient method for an inverse problem resolution
  70. Supersolutions to nonautonomous Choquard equations in general domains
  71. Uniform complex time heat Kernel estimates without Gaussian bounds
  72. Global existence for time-dependent damped wave equations with nonlinear memory
  73. Normalized solutions for a critical fractional Choquard equation with a nonlocal perturbation
  74. Reversed Hardy-Littlewood-Sobolev inequalities with weights on the Heisenberg group
  75. Lamé system with weak damping and nonlinear time-varying delay
  76. Global well posedness for the semilinear edge-degenerate parabolic equations on singular manifolds
  77. Blowup in L1(Ω)-norm and global existence for time-fractional diffusion equations with polynomial semilinear terms
  78. Boundary regularity results for minimisers of convex functionals with (p, q)-growth
  79. Parametric singular double phase Dirichlet problems
  80. Special Issue on Nonlinear analysis: Perspectives and synergies
  81. Editorial to Special issue “Nonlinear analysis: Perspectives and synergies”
  82. Identification of discontinuous parameters in double phase obstacle problems
  83. Global boundedness to a 3D chemotaxis-Stokes system with porous medium cell diffusion and general sensitivity
  84. On regular solutions to compressible radiation hydrodynamic equations with far field vacuum
  85. On Cauchy problem for fractional parabolic-elliptic Keller-Segel model
  86. The evolution of immersed locally convex plane curves driven by anisotropic curvature flow
  87. Stability of combination of rarefaction waves with viscous contact wave for compressible Navier-Stokes equations with temperature-dependent transport coefficients and large data
  88. On concave perturbations of a periodic elliptic problem in R2 involving critical exponential growth
  89. Existence of solutions for a double-phase variable exponent equation without the Ambrosetti-Rabinowitz condition
Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2022-0314/html
Scroll to top button