Centers of projective vector fields of spatial quasi-homogeneous systems with weight (m, m, n) and degree 2 on the sphere

Similar documents
Qualitative Classification of Singular Points

2.10 Saddles, Nodes, Foci and Centers

THE NUMBER OF POLYNOMIAL SOLUTIONS OF POLYNOMIAL RICCATI EQUATIONS

PLANAR ANALYTIC NILPOTENT GERMS VIA ANALYTIC FIRST INTEGRALS

arxiv: v1 [math.ds] 11 Aug 2017

7 Planar systems of linear ODE

Lotka Volterra Predator-Prey Model with a Predating Scavenger

POLYNOMIAL FIRST INTEGRALS FOR WEIGHT-HOMOGENEOUS PLANAR POLYNOMIAL DIFFERENTIAL SYSTEMS OF WEIGHT DEGREE 4

Sample Solutions of Assignment 10 for MAT3270B

154 Chapter 9 Hints, Answers, and Solutions The particular trajectories are highlighted in the phase portraits below.

11 Chaos in Continuous Dynamical Systems.

STABILITY. Phase portraits and local stability

ẋ = f(x, y), ẏ = g(x, y), (x, y) D, can only have periodic solutions if (f,g) changes sign in D or if (f,g)=0in D.

Chapter 3 Second Order Linear Equations

HASSE-MINKOWSKI THEOREM

A BENDIXSON DULAC THEOREM FOR SOME PIECEWISE SYSTEMS

The Liapunov Method for Determining Stability (DRAFT)

1. < 0: the eigenvalues are real and have opposite signs; the fixed point is a saddle point

CUBIC SYSTEMS WITH INVARIANT AFFINE STRAIGHT LINES OF TOTAL PARALLEL MULTIPLICITY SEVEN

Chapter 4: First-order differential equations. Similarity and Transport Phenomena in Fluid Dynamics Christophe Ancey

Practice Problems for Final Exam

Solutions to Dynamical Systems 2010 exam. Each question is worth 25 marks.

Problem List MATH 5173 Spring, 2014

ZERO-HOPF BIFURCATION FOR A CLASS OF LORENZ-TYPE SYSTEMS

Uniqueness of Limit Cycles For Liénard Differential Equations of Degree Four

5 Quiver Representations

GEOMETRIC CLASSIFICATION OF CONFIGURATIONS OF SINGULARITIES WITH TOTAL FINITE MULTIPLICITY FOUR FOR A CLASS OF QUADRATIC SYSTEMS

Problem set 7 Math 207A, Fall 2011 Solutions

Construction of Lyapunov functions by validated computation

ON THE DEFINITION OF RELATIVE PRESSURE FOR FACTOR MAPS ON SHIFTS OF FINITE TYPE. 1. Introduction

GLOBAL TOPOLOGICAL CONFIGURATIONS OF SINGULARITIES FOR THE WHOLE FAMILY OF QUADRATIC DIFFERENTIAL SYSTEMS

Topological properties

Laplace s Equation. Chapter Mean Value Formulas

On the flat geometry of the cuspidal edge

ENGI Duffing s Equation Page 4.65

Contents. MATH 32B-2 (18W) (L) G. Liu / (TA) A. Zhou Calculus of Several Variables. 1 Multiple Integrals 3. 2 Vector Fields 9

Entrance Exam, Differential Equations April, (Solve exactly 6 out of the 8 problems) y + 2y + y cos(x 2 y) = 0, y(0) = 2, y (0) = 4.

Solution of Additional Exercises for Chapter 4

Chapter 6 Nonlinear Systems and Phenomena. Friday, November 2, 12

Practice Problems for Exam 3 (Solutions) 1. Let F(x, y) = xyi+(y 3x)j, and let C be the curve r(t) = ti+(3t t 2 )j for 0 t 2. Compute F dr.

THEODORE VORONOV DIFFERENTIABLE MANIFOLDS. Fall Last updated: November 26, (Under construction.)

ARCS IN FINITE PROJECTIVE SPACES. Basic objects and definitions

Stability of Feedback Solutions for Infinite Horizon Noncooperative Differential Games

arxiv: v1 [math.ds] 27 Jul 2017

A PROOF OF PERKO S CONJECTURES FOR THE BOGDANOV-TAKENS SYSTEM

Differential equations, comprehensive exam topics and sample questions

UNIVERSIDADE DE SÃO PAULO

Problem Set Number 5, j/2.036j MIT (Fall 2014)

Constructions with ruler and compass.

PERIODIC SOLUTIONS WITH NONCONSTANT SIGN IN ABEL EQUATIONS OF THE SECOND KIND

HIGHER ORDER CRITERION FOR THE NONEXISTENCE OF FORMAL FIRST INTEGRAL FOR NONLINEAR SYSTEMS. 1. Introduction

TEST CODE: MMA (Objective type) 2015 SYLLABUS

B5.6 Nonlinear Systems

ALGEBRAIC TOPOLOGY N. P. STRICKLAND

CALCULUS JIA-MING (FRANK) LIOU

(1) u (t) = f(t, u(t)), 0 t a.

GEOMETRIC CONFIGURATIONS OF SINGULARITIES FOR QUADRATIC DIFFERENTIAL SYSTEMS WITH TOTAL FINITE MULTIPLICITY m f = 2

10. Smooth Varieties. 82 Andreas Gathmann

The Kernel Trick. Robert M. Haralick. Computer Science, Graduate Center City University of New York

Mathematical Olympiad Training Polynomials

ON THE MULTIPLICITY OF EIGENVALUES OF A VECTORIAL STURM-LIOUVILLE DIFFERENTIAL EQUATION AND SOME RELATED SPECTRAL PROBLEMS

SOLUTIONS TO THE FINAL EXAM. December 14, 2010, 9:00am-12:00 (3 hours)

1. Introduction and statement of main results. dy dt. = p(x, y),

Math 312 Lecture Notes Linearization

An introduction to Birkhoff normal form

Half of Final Exam Name: Practice Problems October 28, 2014

Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling

TWELVE LIMIT CYCLES IN A CUBIC ORDER PLANAR SYSTEM WITH Z 2 -SYMMETRY. P. Yu 1,2 and M. Han 1

LIMIT CYCLES FOR A CLASS OF ELEVENTH Z 12 -EQUIVARIANT SYSTEMS WITHOUT INFINITE CRITICAL POINTS

7.3 Singular points and the method of Frobenius

GEOMETRIC CONFIGURATIONS OF SINGULARITIES FOR QUADRATIC DIFFERENTIAL SYSTEMS WITH TOTAL FINITE MULTIPLICITY m f = 2

CHAPTER 3. Gauss map. In this chapter we will study the Gauss map of surfaces in R 3.

3 Applications of partial differentiation

CLASSIFICATION AND PRINCIPLE OF SUPERPOSITION FOR SECOND ORDER LINEAR PDE

Research Article Oscillation Criteria of Certain Third-Order Differential Equation with Piecewise Constant Argument

Part II. Dynamical Systems. Year

Existence and multiple solutions for a second-order difference boundary value problem via critical point theory

1 Assignment 1: Nonlinear dynamics (due September

g(t) = f(x 1 (t),..., x n (t)).

Lecture 2 Linear Codes

LINEAR ALGEBRA W W L CHEN

Pattern generation, topology, and non-holonomic systems

Calculus 2502A - Advanced Calculus I Fall : Local minima and maxima

PHASE PORTRAITS OF PLANAR SEMI HOMOGENEOUS VECTOR FIELDS (III)

Kinematics of fluid motion

Exercise The solution of

A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION

Math 308 Final Exam Practice Problems

Quadratic forms. Here. Thus symmetric matrices are diagonalizable, and the diagonalization can be performed by means of an orthogonal matrix.

Formal Modules. Elliptic Modules Learning Seminar. Andrew O Desky. October 6, 2017

Lorenz like flows. Maria José Pacifico. IM-UFRJ Rio de Janeiro - Brasil. Lorenz like flows p. 1

Math 234 Final Exam (with answers) Spring 2017

Introduction of Partial Differential Equations and Boundary Value Problems

Normal forms for germs of vector fields with quadratic leading part. The remaining cases

Prof. Krstic Nonlinear Systems MAE281A Homework set 1 Linearization & phase portrait

EE222 - Spring 16 - Lecture 2 Notes 1

Math 868 Final Exam. Part 1. Complete 5 of the following 7 sentences to make a precise definition (5 points each). Y (φ t ) Y lim

arxiv: v2 [math.ag] 24 Jun 2015

ENGI 9420 Lecture Notes 4 - Stability Analysis Page Stability Analysis for Non-linear Ordinary Differential Equations

Transcription:

Electronic Journal of Qualitative Theory of Differential Equations 2016 No. 103 1 26; doi: 10.14232/ejqtde.2016.1.103 http://www.math.u-szeged.hu/ejqtde/ Centers of projective vector fields of spatial quasi-homogeneous systems with weight (m m n) and degree 2 on the sphere Haihua Liang 1 and Joan Torregrosa 2 1 Guangdong Polytechnic Normal University Zhongshan Highway Guangzhou 510665 P. R. China 2 Universitat Autònoma de Barcelona Bellaterra Barcelona 08193 Spain Received 4 November 2015 appeared 17 November 2016 Communicated by Gabriele Villari Abstract. In this paper we study the centers of projective vector fields Q T of threedimensional quasi-homogeneous differential system dx/ = Q(x) with the weight (m m n) and degree 2 on the unit sphere S 2. We seek the sufficient and necessary conditions under which Q T has at least one center on S 2. Moreover we provide the exact number and the positions of the centers of Q T. First we give the complete classification of systems dx/ = Q(x) and then using the induced systems of Q T on the local charts of S 2 we determine the conditions for the existence of centers. The results of this paper provide a convenient criterion to find out all the centers of Q T on S 2 with Q being the quasi-homogeneous polynomial vector field of weight (m m n) and degree 2. Keywords: projective vector field quasi-homogeneous system sufficient and necessary conditions for centers. 2010 Mathematics Subject Classification: Primary 37C10. Secondary 37C27. 1 Introduction We consider the polynomial differential systems in R 3 dx = Q(x) (1.1) where x = (x 1 x 2 x 3 ) and Q(x) = (Q 1 (x) Q 2 (x) Q 3 (x)). System (1.1) is called a quasihomogeneous polynomial differential system with weight (α 1 α 2 α 3 ) and degree d if Q(x) is a quasi-homogeneous polynomial vector field with weight (α 1 α 2 α 3 ) and degree d i.e. Q i (λ α 1 x 1 λ α 2 x 2 λ α 3 x 3 ) = λ α i 1+d Q i (x 1 x 2 x 3 ) i = 1 2 3 (1.2) where λ R and d α 1 α 2 α 3 Z +. In particular if (α 1 α 2 α 3 ) = (1 1 1) then system (1.1) is a homogeneous polynomial system of degree d. Corresponding author. Email: torre@mat.uab.cat

2 H. Liang and J. Torregrosa The three-dimensional polynomial differential systems occur as models or at least as simplifications of models in many domains in science. For example the population models in biology. In recent years the qualitative theory of three-dimensional polynomial differential systems has been and still is receiving intensive attention [1 2 5 7 9 10 13 15]. Just as the author of [16] point out that the study of three-dimensional polynomial differential system is much more difficult than that of planar polynomial system. For example it is an arduous task to determine the global topological structure of the Lorenz system ẋ 1 = σ(x 2 x 1 ) ẋ 2 = ρx 1 y xz ẋ 3 = βz + xy (σ β ρ 0) although this system has a simply form see [14]. An efficient method for studying the qualitative behavior of orbits of system (1.1) is to project the system to the unit sphere S 2. In what follows we will adopt the notations used in [6] to introduce some basic theory of the projective system on S 2. By taking the transformation of coordinates x = (x 1 x 2 x 3 ) = (r α 1 y 1 r α 2 y 2 r α 3 y 3 ) y = (y 1 y 2 y 3 ) S 2 r R + we get from system (1.1) that dr dτ = r y Q(y) =: r R(y) (1.3) dy dτ = ȳ y Q(y) y Q(y) ȳ =: Q T(y) (1.4) where ȳ = (α 1 y 1 α 2 y 2 α 3 y 3 ) and dτ = (r d 1 / ȳ y ). System (1.4) plays an important role in the analysis of the topology of system (1.1). Indeed if we write Γ g and C for trajectory singularity and closed orbit of system (1.4) on the S 2 respectively and let y = y(τ y 0 ) be the expression of Γ (resp. g C) with initial value y 0 = y(τ 0 y 0 ) then y(τ y 0 ) is defined on R and r(τ τ 0 ) = r 0 exp τ τ 0 R(y(s y 0 )) ds is the solution of (1.3). Hence we obtain the corresponding trajectory of system (1.1) W Γ (resp. W g W C ) = {(r α 1 (τ τ 0 )y 1 (τ y 0 ) r α 2 (τ τ 0 )y 2 (τ y 0 ) r α 3 (τ τ 0 )y 3 (τ y 0 )) τ R}. For any y S 2 we define a curve as S(y) = {(r α 1y 1 r α 2y 2 r α 3y 3 ) r > 0}. The orbit Γ of system (1.4) on S 2 can be regarded as the projection of W Γ along the family of curves {S(y) y S 2 }. In this sense we call Q T (y) the projective vector field of Q(y) on S 2 and call (1.4) the projective system of (1.1). To study the behavior of orbits of system (1.4) we use the local charts of S 2. Denoted by and H + i = {x R 3 : x i > 0} H i = {x R 3 : x i < 0} Π + i = { x R 3 : x i = 1} Π i = { x R 3 : x i = 1}. Define respectively the coordinate transformations φ+ i : H + i S 2 Π + i and φ i : H i S 2 Π i ( ) x = φ i +(y) = y 1 y α 1/α i i y 2 y α 2/α i i y 3 y α 3/α i i

Centers of projective quasi-homogeneous vector fields 3 and x = φ i (y) = ( y1 y 2 y i α 1/α i y i α 2/α i y 3 y i α 3/α i ) for i = 1 2 3. It is easy to see that {(H ± i S 2 φ i ±) : i = 1 2 3} is the set of local charts of S 2. System (1.4) in these local charts is topologically equivalent to ( d x d τ = Wi +( x) = Q 1 ( x) α 1 x 1 Q i ( x) Q 2 ( x) α 2 x 2 Q i ( x) Q 3 ( x) α ) 3 x 3 Q i ( x) α i α i α ( i d x d τ = Wi ( x) = Q 1 ( x) + α 1 x 1 Q i ( x) Q 2 ( x) + α 2 x 2 Q i ( x) Q 3 ( x) + α ) 3 x 3 Q i ( x) α i α i α i where i = 1 2 3 and d τ = ȳ y y i (d 1)/α idτ. In the literature many authors study the projective vector field of system (1.1) with degree two (d = 2). Most of them consider the homogeneous case i.e. α 1 = α 2 = α 3 = 1. For instance Camacho in [1] investigates the projective vector fields of homogeneous polynomial system of degree two. The classification of projective vector fields without periodic orbits on S 2 is given. Wu in [15] corrects some mistakes of [1] and provide several properties of homogeneous vector fields of degree two. Llibre and Pessoa in [10] study the homogeneous polynomial vector fields of degree two it was shown that if the vector field on S 2 has finitely many invariant circles then every invariant circle is a great circle. [11] deals with the phase portraits for quadratic homogeneous polynomial vector fields on S 2 they verify that if the vector field has at least a non-hyperbolic singularity then it has no limit cycles. They also give necessary and sufficient conditions for determining whether a singularity of (1.4) on S 2 is a center. Pereira and Pessoa in [12] classify all the centers of a certain class of quadratic reversible polynomial vector fields on S 2. Under the homogeneity assumption we know that whenever x(t) is a solution of system (1.1) then so is x = λx(λ d 1 t). This conclusion can be extended to quasi-homogeneous systems. Indeed from the quasi-homogeneity x(t) = diag(λ α 1 λ α 2 λ α 3)x(λ d 1 t) is a solution of (1.1) when x(t) is a solution of (1.1). Recently the authors of [6] study the projective vector field of a three-dimensional quasi-homogeneous system with weight (1 1 α) with α > 1 and degree d = 2. Some interesting qualitative behaviours are determined according to the parameters of the systems. Another meaningful work about the spatial quasi-homogeneous systems is [7]. In that paper the authors generalize the results of [2 13] by studying the limit set of trajectories of three-dimensional quasi-homogeneous systems. They also point out by a counterexample the mistake of [2]. However to the best of our knowledge there is no paper dealing with the center of the projective vector field of spatial quasi-homogeneous systems. Motivated by this fact in the present paper we study the sufficient and necessary conditions for the projective vector field Q T of the system (1.1) with the weight (m m n) and degree 2 to have at least one center on S 2. We would like to emphasize that in the above mentioned papers dealing with homogeneous systems many authors concern on the periodic orbits of system (1.4) see [1 6 11]. This is because the periodic behavior of system (1.4) provide a threshold to investigate the periodic and spirally behaviors of the spatial system. Our work provides a criterion for the projective vector field associated to system (1.1) to have a family of periodic orbits. This paper is organized as follows. In Section 2 we prove some properties and establish the canonical forms of quasi-homogeneous polynomial system (1.1) with weight (m m n) and degree 2. In Sections 3 4 and 5 we are going to seek the sufficient and necessary condi-

4 H. Liang and J. Torregrosa tions under which the projective system (1.4) has at least one center on S 2 where Section 3 (resp. Section 4 and Section 5) deals with the case that n = 1 (resp. m > 1 n > 1 and m = 1). 2 Properties and canonical forms for quasi-homogeneous systems with weight (m m n) and degree d = 2 The first goal of this section is to derive some properties of the three-dimensional quasihomogeneous polynomial vector field with weight (m m n) and degree d = 2. The results obtained will be used in the next sections. Define a homomorphism ψ : R 3 R 3 as ψ(x 1 x 2 x 3 ) = (( 1) m x 1 ( 1) m x 2 ( 1) n x 3 ). Proposition 2.1. Assume that Q is a quasi-homogeneous polynomial vector field with weight (m m n) and degree d = 2. Then Q(ψ(x)) = Dψ Q(x) Q T (ψ(y)) = Dψ Q T (y). Proof. Firstly it follows from the quasi-homogeneity property of Q that Q(ψ(x)) = Q(( 1) m x 1 ( 1) m x 2 ( 1) n x 3 ) = (( 1) m+1 Q 1 (x) ( 1) m+1 Q 2 (x) ( 1) n+1 Q 3 (x)) = Dψ Q(x). Secondly by the expression of Q T (y) we have Q T (ψ(y)) = ψ(y) ψ(y) Q(ψ(y)) ψ(y) Q(ψ(y)) ψ(y). Since ψ(y) = Dψ ȳ it follows that ψ(y) ψ(y) = ȳ y. Hence The proof is finished. Q T (ψ(y)) = ȳ y Dψ Q(y) + ψ(y) Dψ Q(y) ψ(y) = ȳ y Dψ Q(y) + y Q(y) ψ(y) = Dψ ( ȳ y Q(y) y Q(y) ȳ) = Dψ Q T (y). Proposition 2.2. Assume that Q is a quasi-homogeneous polynomial vector field with weight (m m n). Let L = {(λ cos α 0 λ sin α 0 1) λ R} be a straight line on Π + 3. If S S2 is a great circle which contains the points (0 0 ±1) and (cos α 0 sin α 0 0) then (φ 3 +) 1 (L) = H + 3 S. Proof. Since S is a great circle containing the points (0 0 ±1) and (cos α 0 sin α 0 0) we find H + 3 S = {(± cos α 0 sin θ ± sin α 0 sin θ cos θ) θ [0 π/2)}. Hence { ( ) φ+(h 3 + 3 S) = y1 (y 3 ) m/n y 2 (y 3 ) m/n 1 (y 1 y 2 y 3 ) = (± cos α 0 sin θ ± sin α 0 sin θ cos θ) } θ [0 π/2) { } = (λ cos α 0 λ sin α 0 1) λ = ± sin θ/(cos θ) m/n θ [0 π/2) = {(λ cos α 0 λ sin α 0 1) λ ( )}. The proof finishes because the above expression is equivalent to (φ 3 +) 1 (L) = H + 3 S.

Centers of projective quasi-homogeneous vector fields 5 The second purpose of this section is to obtain the canonical form for the quasi-homogeneous polynomial vector fields with weight (m m n) and degree d = 2 where m and n are two different positive integers. Lemma 2.3. Every three-dimensional quasi-homogeneous polynomial differential system (1.1) with weight (m m n) and degree d = 2 can be written in one of the following forms: (1) If m = 1 then dx 1 dx 2 dx 3 = a ij0 x1 i xj 2 + (1 sgn(n 2))a 001x 3 i+j=2 = b ij0 x1 i xj 2 + (1 sgn(n 2))b 001x 3 (2.1) i+j=2 = a 101 x 1 x 3 + a 011 x 2 x 3 + c ij0 x1 i xj 2. i+j=n+1 (2) If n = 1 then dx 1 dx 2 dx 3 = a 101 x 1 x 3 + a 011 x 2 x 3 + a 00m+1 x m+1 3 = b 101 x 1 x 3 + b 011 x 2 x 3 + b 00m+1 x m+1 3 (2.2) = (1 sgn(m 2))(c 100 x 1 + c 010 x 2 ) + c 002 x 2 3. (3) If m 2 and n 2 then dx 1 = a 00 m+1 n x m+1 n 3 dx 2 = b 00 m+1 n x m+1 n 3 dx 3 = c ij0 x1 i xj 2. (2.3) i+j= n+1 m Here sgn( ) is the sign function and in (2.3) we define a 00 m+1 n and c ij0 = 0 if i + j = (n + 1)/m / N +. = b 00 m+1 n = 0 if (m + 1)/n / N + Proof. It follows from (1.2) that Q i (0 0 0) = (0 0 0). Thus we set where q ik (x 1 x 2 x 3 ) = Q i (x 1 x 2 x 3 ) = n i q ik (x 1 x 2 x 3 ) k=1 k a (i) k 1 k 2 k k 1 k 2 x k 1 1 xk 2 2 xk k 1 k 2 3 i = 1 2 3. k 1 +k 2 =0 Substituting the above expressions into (1.2) with (α 1 α 2 α 3 ) = (m m n) and d = 2 yields a (i) k 1 k 2 k k 1 k 2 (λ m(k 1+k 2 1)+n(k k 1 k 2 ) 1 1) = 0 k = 1 2... n i i = 1 2 (2.4) a (3) k 1 k 2 k k 1 k 2 (λ m(k 1+k 2 )+n(k k 1 k 2 1) 1 1) = 0 k = 1 2... n 3. (2.5) We will apply (2.4) and (2.5) to find out all the coefficients which vanish.

6 H. Liang and J. Torregrosa Case 1. m = 1 n 2. If a (i) k 1 k 2 k k 1 k 2 = 0 i = 1 2 then by (2.4) we have k 1 + k 2 + n(k k 1 k 2 ) 2 = 0. (2.6) Noting that 0 k 1 + k 2 k we deduce from (2.6) that k 1 + k 2 = 0 k = 1 n = 2 or k 1 + k 2 = k = 2. This prove the first and the second equation of (2.1). If a (3) k 1 k 2 k k 1 k 2 = 0 then by (2.5) k 1 + k 2 + n(k k 1 k 2 1) 1 = 0. (2.7) The equation (2.7) is satisfied if and only if k 1 + k 2 = 1 k = 2 or k 1 + k 2 = k = n + 1 3. This proves the third equation of (2.1). Case 2. n = 1 m 2. If a (i) k 1 k 2 k k 1 k 2 = 0 i = 1 2 then we get m(k 1 + k 2 1) + k k 1 k 2 1 = 0. (2.8) We deduce from (2.8) that k 1 + k 2 = 0 k = m + 1 3 or k 1 + k 2 = 1 k = 2. This proves the first and the second equation of (2.2). If a (3) k 1 k 2 k k 1 k 2 = 0 then m(k 1 + k 2 ) + k k 1 k 2 2 = 0. (2.9) The equation (2.9) is satisfied if and only if k 1 + k 2 = 0 k = 2 or k 1 + k 2 = k = 1 m = 2. This proves the third equation of (2.2). Case 3. n 2 m 2. If a (i) k 1 k 2 k k 1 k 2 = 0 i = 1 2 then we have m(k 1 + k 2 1) + n(k k 1 k 2 ) 1 = 0. (2.10) Since m(k 1 + k 2 1) = 1 n(k k 1 k 2 ) 1 it is enough to consider two cases: k 1 + k 2 = 0 and k 1 + k 2 = 1 k. Furthermore k 1 + k 2 = 1 k is impossible because n(k 1) = 1 for all n 2. If k 1 + k 2 = 0 then we get from (2.10) that kn = m + 1. This means that Q 1 and Q 2 are two nonzero functions if and only if n (m + 1). And hence the first and the second equation of (2.3) are obtained. If a (3) k 1 k 2 k k 1 k 2 = 0 then m(k 1 + k 2 ) + n(k k 1 k 2 1) 1 = 0. This equality holds if and only if k 1 + k 2 = k 1 km = n + 1. Thus we obtain the third equation of (2.3). From the above lemma we get next result. Theorem 2.4. Suppose that Q i (i = 1 2 3) of system (1.1) are nonzero functions. Then every quasihomogeneous polynomial vector field (1.1) with weight (m m n) and degree d = 2 can be changed under a suitable affine transformation to: (i) System with weight (1 1 n) where n 3 or dx 1 = a 1 x1 2 + a 2x 1 x 2 + a 3 x2 2 dx 2 = b 1 x1 2 + b 2x 1 x 2 + b 3 x2 2 n+1 dx 3 = c 1 x 1 x 3 + c 2 x 2 x 3 + i=0 d i x i 1 xn+1 i 2 (2.11)

Centers of projective quasi-homogeneous vector fields 7 (ii) system dx 1 dx 2 dx 3 = a 1 x 2 1 + a 2x 1 x 2 + a 3 x 2 2 = b 1 x 2 1 + b 2x 1 x 2 + b 3 x 2 2 + x 3 = c 1 x 1 x 3 + c 2 x 2 x 3 + 3 i=0 d i x i 1 x3 i 2 (2.12) with weight (1 1 2) or (iii) system with weight (2 2 1) where η = 0 1 or (iv) system dx 1 = a 1 x 1 x 3 + a 2 x 2 x 3 dx 2 = b 1 x 1 x 3 + b 2 x 2 x 3 + ηx3 3 dx 3 = c 1 x 1 + c 2 x 2 + c 0 x3 2 dx 1 = a 1 x 1 x 3 + a 2 x 2 x 3 dx 2 = b 1 x 1 x 3 + b 2 x 2 x 3 + ηx3 m+1 dx 3 = x3 2 with weight (m m 1) where η = 0 1 and m 3 or (2.13) (2.14) (v) system dx 1 = x 3 dx 2 = x 3 dx 3 = c 1 x 2 1 + c 2x 1 x 2 + c 3 x 2 2 (2.15) with weight (2 2 3) or (vi) system dx 1 = x 2 3 dx 2 = x 2 3 dx 3 = c 1 x 1 + c 2 x 2 (2.16) with weight (3 3 2). Proof. Firstly in the case that m = 1 the canonical forms (2.11) and (2.12) follow from [6] directly. Let us consider the case that n = 1 and m = 2. If a 003 = b 003 = 0 then system (2.2) becomes (2.13) with η = 0. If a 2 003 + b2 003 = 0 then by noting that system (2.2) has the same form under the change of variables (x 1 x 2 ) (x 2 x 1 ) we can assume without loss of generality that b 003 = 0. Taking the transformation z = (b 003 x 1 a 003 x 2 b003 1 x 2 x 3 ) and then using the symbol x instead of z we can change system (2.2) to (2.13) with η = 1. The case n = 1 and m 3 can be dealt with in a similar way. Finally consider the case that m 2 n 2. Since Q i (i = 1 2 3) is nonzero function it follows that (m + 1)/n (n + 1)/m N +. Thus m = n 1 or m = n + 1. If m = n 1 we

8 H. Liang and J. Torregrosa get from n (m + 1) and m (n + 1) that n = 3 m = 2. If m = n + 1 then n = 2 m = 3. Consequently we obtain dx 1 = ax 3 dx 2 = bx 3 dx 3 = c 1 x1 2 + c 2x 1 x 2 + c 3 x2 2 ab = 0 with weight (2 2 3) and dx 1 = ax 2 3 dx 2 = bx 2 3 dx 3 = c 1 x 1 + c 2 x 2 ab = 0 with weight (3 3 2). By taking an affine transformation of variables we get systems (2.15) and (2.16). The systems (2.11) and (2.12) are considered in [6]. It is shown that the projective system of system (2.11) has no closed orbits on S 2. But the authors do not give the conditions for projective systems of system (2.12) to acquire at least one center. The purpose of the rest of this paper is to find the sufficient and necessary conditions for all the projective systems (1.4) of the systems in Theorem 2.4 to have at least one center. 3 Center of the quasi-homogeneous systems with weight (m m 1) In this section we deal with the canonical forms of (2.13) and (2.14). The main results of this section are the following two theorems. Theorem 3.1. Suppose that Q T is the projective vector field of system (2.13) then the following statements hold. (A) For η = 0 Q T has at least one center if and only if one of the following two conditions is satisfied: (1) a 1 + b 2 = 4c 0 (b 2 a 1 ) 2 + 4a 2 b 1 < 0; (2) J(c 1 c 2 ) = a 2 c 2 1 + (b 2 a 1 )c 1 c 2 b 1 c 2 2 = 0. In addition (i) If (1) is satisfied and c 2 1 + c2 2 = 0 then Q T has exactly three centers respectively at (0 0 1) ( ) (0 0 1) and E c 2 / c 2 1 + c2 2 c 1/ c 21 + c22 0 when J(c 1 c 2 ) > 0 or E when J(c 1 c 2 ) < 0. (ii) If (1) is satisfied and c 1 = c 2 = 0 then Q T has exactly two centers at (0 0 1) and (0 0 1) respectively. (iii) If (2) is satisfied but (1) is not satisfied then Q T has a unique center at E when J(c 1 c 2 ) > 0 or at E when J(c 1 c 2 ) < 0. (B) For η = 1 Q T has at least one center if and only if J(c 1 c 2 ) = 0 or one of following conditions is satisfied: (1) a 2 = 0 b 2 = 2ā 1 ā 1 c 1 b 1 c 2 = 0 ā 2 1 + 2c 2 < 0; (2) a 2 = 0 b 2 = ā 1 c 1 a 2 = ā 1 c 2 (ā 4 1 + 2ā2 1 a 2b 1 + a 2 2 b2 1 + 2a 2b 1 c 2 )(a 2 b 1 + ā 2 1 ) < 0; (3) a 2 = 0 2ā 3 1 3ā2 1 b 2 + 9ā 1 a 2 b 1 3ā 1 b2 2 36ā 1 c 2 + v9a 2 b 1 b2 + 2 b 3 2 + 54a 2c 1 + 18 b 2 c 2 = 0 (9ā 2 1 a 2c 1 + 27a 2 2 b 1c 1 9ā 1 a 2 b2 c 1 + 9a 2 b2 2 c 1 8ā 3 1 c 2 27ā 1 a 2 b 1 c 2 + 12ā 2 1 b 2 c 2 6ā 1 b2 2 c 2 + b 3 2 c 2) (3a 2 c 1 + b 2 c 2 2ā 1 c 2 ) < 0

Centers of projective quasi-homogeneous vector fields 9 with ā 1 = a 1 2c 0 b 2 = b 2 2c 0. Moreover if J(c 1 c 2 ) > 0 (resp. J(c 1 c 2 ) < 0) then on the equator Q T has a unique center at E (resp. E). If the condition (i) (i {1 2 3}) of (B) holds then Q T has a unique center at y i on S 2 H + 3 and it has a unique center at y i on S 2 H 3 which satisfies φ3 +(y i ) = (xi y i 1) where x1 ā 2 1 = + 2c 2 y1 = ā1b 1 + 2c 1 2ā 1 c 1 2b 1 c 2 2ā 1 c 1 2b 1 c 2 a 2 x2 = a 2 b 1 + ā 2 1 y2 = a 2 b 1 + ā 2 1 x3 a 2 (ā = 1 + b 2 ) 2(3a 2 c 1 + b 2 c 2 2ā 1 c 2 ) y 3 = ( b 2 2ā 1 )( b 2 + ā 1 ) 6(3a 2 c 1 + b 2 c 2 2ā 1 c 2 ). ā 1 (3.1) Theorem 3.2. The projective vector field Q T of system (2.14) with m > 2 has at least one center on S 2 if and only if a 1 + b 2 = 2m 1 = (b 2 a 1 ) 2 + 4a 2 b 1 < 0. (3.2) Furthermore if (3.2) is satisfied then Q T has exactly two centers at the points ( ηa2 λ0 m η(m a 1)λ m ) ( 0 ηa2 ( λ 0 ) ηλ 0 + m 1 η and η(m a 1)( λ 0 ) m ) ηλ 0 +η 1 1 1 1 1 where λ = λ 0 is the unique positive solution of equation (a 2 2 + (m a 1 ) 2 )λ 2m + 2 1 λ2 = 2 1. (3.3) 3.1 Quasi-homogeneous systems with weight (2 2 1) In this subsection we assume that Q T is the projective vector field of system (2.13). We will firstly study the centers on the H + 3 S2. The centers on H 3 S2 can be obtained by the symmetry (see Proposition 2.1). By straightforward calculations we find that the induced system of Q T on Π + 3 is d x 1 d τ = Q 1( x) 2 x 1 Q 3 ( x) = (a 1 2c 0 ) x 1 + a 2 x 2 2c 1 x 1 2 2c 2 x 1 x 2 d x 2 d τ = Q 2( x) 2 x 2 Q 3 ( x) = η + b 1 x 1 + (b 2 2c 0 ) x 2 2c 1 x 1 x 2 2c 2 x 2 2 η = 0 1. Proposition 3.3. Assume that p = (p 1 p 2 p 3 ) S 2 H + 3 with p 3 = 0 ±1 is a singularity of Q T. If η = 0 then p is not a center of Q T. Proof. Since Q T (p) = 0 it follows that where q = (2p 1 2p 2 p 3 ). Thus q p Q i (p) 2 p Q(p) p i = 0 i = 1 2 0 = p 1 (Q 2 + 2p 2 Q 3 ) p 2 (Q 1 + 2p 1 Q 3 ) = p 1 Q 2 p 2 Q 1 = p 3 [I(p 1 p 2 ) + ηp 2 3] (3.5) where I(x y) = b 1 x 2 + (b 2 a 1 )xy a 2 y 2. Let l + = {( p 1 λ p 2 λ 1) λ R} be a straight line on Π + 3 where p 1 = p 1 / 1 p 2 3 p 2 = p 2 / 1 p 2 3. By direct computation we obtain that on l + writing x = ( p 1 λ p 2 λ 1) we have f (λ) := p 1 (Q 2 ( x) 2 x 2 Q 3 ( x)) p 2 (Q 1 ( x) 2 x 1 Q 3 ( x)) = λi( p 1 p 2 ) + p 1 η η = 0 1. (3.4)

10 H. Liang and J. Torregrosa If η = 0 then it follows from (3.5) that I( p 1 p 2 ) = I(p 1 p 2 )/(1 p 2 3 ) = 0. This means that f (λ) 0. Therefore l + is an invariant straight line of W + 3. Let S be the great circle containing the points ( p 1 p 2 0) and (0 0 ±1). Clearly p S. By Proposition 2.2 the half-great circle S H + 3 is an integral curve of the vector field Q T. Thus p can no be a center of Q T. Next consider the critical point (0 0 ±1) of Q T with η = 0. We need the following result. Lemma 3.4 ([8]). The origin is a center of the following system dx = ax + by + a 20x 2 + a 11 xy + a 02 y 2 dy = cx ay + b 20x 2 + b 11 xy + b 02 y 2 with a 2 + bc < 0 if and only if one of the following conditions holds: (1) Aα Bβ = γ = 0 (2) α = β = 0 (3) 5A β = 5B α = δ = 0 where A = a 20 + a 02 B = b 20 + b 02 α = a 11 + 2b 02 β = b 11 + 2a 20 γ = b 20 A 3 (a 20 b 11 )A 2 B + (b 02 a 11 )AB 2 a 02 B 3 and δ = a 2 02 + b2 20 + a 02A + b 20 B. Proposition 3.5. Assume that η = 0. The points (0 0 ±1) are centers of Q T on S 2 if and only if a 1 + b 2 = 4c 0 (b 2 a 1 ) 2 + 4a 2 b 1 < 0. (3.6) Proof. To study the singularity (0 0 1) we will use the induced system on Π + 3 which is (3.4) with η = 0. Clearly φ+(0 3 0 1) = (0 0 1). The characteristic equation of the linear approximation system of (3.4) at the singularity (0 0) is λ 2 (a 1 + b 2 4c 0 )λ + (a 1 2c 0 )(b 2 2c 0 ) a 2 b 1 = 0. The singularity (0 0) is a center or focus of the system (3.4) if and only if a 1 + b 2 4c 0 = 0 (a 1 2c 0 )(b 2 2c 0 ) a 2 b 1 > 0 which is equivalent to (3.6). By straightforward calculations we find that Aα Bβ = γ = 0 where A = 2c 1 B = 2c 2 α = 6c 2 β = 6c 1. Therefore (0 0) is a center of system (3.4) if and only if the relation (3.6) is satisfied. By applying the Proposition 2.1 we can conclude that the relations (3.6) are also the sufficient and necessary conditions for (0 0 1) to be a center of Q T on S 2. Let us consider the case η = 1. Proposition 3.6. Assume that η = 1 then Q T has at least a center on S 2 H + 3 if and only if one of the conditions of Theorem 3.1 (B) is satisfied. Moreover if the condition (i) (i {1 2 3}) of Theorem 3.1 (B) holds then Q T has a unique center y i on S 2 H + 3 and has a unique center y i on S 2 H 3 which satisfy φ3 +(y i ) = (xi y i 1) where x i and yi are defined in (3.1).

Centers of projective quasi-homogeneous vector fields 11 Proof. Suppose that p H + 3 S2 is a singularity of Q T and let (x 0 y 0 1) = φ + 3 (p). It is easy to see that (x 0 y 0 ) is a singularity of system (3.4). By taking the transformation u = x 1 x 0 v = x 2 y 0 we change system (3.4) to du d τ = (ā 1 4c 1 x 0 2c 2 y 0 )u + (a 2 2c 2 x 0 )v 2c 1 u 2 2c 2 uv dv d τ = (b 1 2c 1 y 0 )u + ( b 2 2c 1 x 0 4c 2 y 0 )v 2c 1 uv 2c 2 v 2. (3.7) In what follows we will consider the singularity (0 0) of system (3.7). One can check directly that system (3.7) satisfies the condition (1) of Lemma 3.4. Thus the point (0 0) is a center of system (3.7) if and only if the following two equalities hold ā 1 4c 1 x 0 2c 2 y 0 + b 2 2c 1 x 0 4c 2 y 0 = 0 (3.8) = (ā 1 4c 1 x 0 2c 2 y 0 ) 2 + (a 2 2c 2 x 0 )(b 1 2c 1 y 0 ) < 0 (3.9) where x 0 and y 0 are the isolated solutions of the following equations ā 1 x 0 + a 2 y 0 2c 1 x 2 0 2c 2 x 0 y 0 = 0 (3.10) 1 + b 1 x 0 + b 2 y 0 2c 1 x 0 y 0 2c 2 y 2 0 = 0. (3.11) By equations (3.8) and (3.10) we get 3a 2 y 0 = ( b 2 2ā 1 )x 0. We will now split our discussion into two cases. Case 1. a 2 = 0. By the inequality (3.9) we know that x 0 = 0. It follows from 3a 2 y 0 = ( b 2 2ā 1 )x 0 that b 2 = 2ā 1. Hence under the condition a 2 = 0 (3.8) (3.11) are equivalent to b 2 = 2ā 1 2c 1 x 0 + 2c 2 y 0 ā 1 = 0 b 1 x 0 + ā 1 y 0 + 1 = 0 = (2ā 1 c 1 2b 1 c 2 )x 0 < 0. (3.12) In view of 2ā 1 c 1 2b 1 c 2 equivalent to = 0 we can get the solution (x 0 y 0 ) and then find that (3.12) is b 2 = 2ā 1 ā 1 c 1 b 1 c 2 = 0 ā 2 1 + 2c 2 < 0 x 0 = ā 2 1 + 2c 2 2ā 1 c 1 2b 1 c 2 =: x 1 y 0 = ā1b 1 + 2c 1 2ā 1 c 1 2b 1 c 2 =: y 1. Consequently under the condition a 2 = 0 the origin of system (3.7) is a center if and only if b 2 = 2ā 1 ā 1 c 1 b 1 c 2 = 0 ā 2 1 + 2c 2 < 0. (3.13) And if the above conditions are satisfied then Q T has a unique center y 1 S 2 H + 3 such that φ+(y 3 1 ) = (x1 y 1 1). By the symmetry we know that Q T also has a unique center y 1 S 2 H 3 if the condition (3.13) is satisfied. Case 2. a 2 = 0. By 3a 2 y 0 = ( b 2 2ā 1 )x 0 and (3.8) we obtain a 2 (ā 1 + b 2 ) = (6a 2 c 1 + 2 b 2 c 2 4ā 1 c 2 )x 0. If 6a 2 c 1 + 2 b 2 c 2 4ā 1 c 2 = 0 then ā 1 = b 2 and thus (3.8) (3.11) are equivalent to c 1 x 0 + c 2 y 0 = 0 ā 1 x 0 + a 2 y 0 = 0 1 + b 1 x 0 ā 1 y 0 = 0 = ā 2 1 + a 2b 1 2b 1 c 2 x 0 < 0. (3.14)

12 H. Liang and J. Torregrosa Further by using (3.14) (3.8) (3.11) are equivalent to ā 1 = b 2 c 1 a 2 = ā 1 c 2 a 2 b 1 + ā 2 1 = 0 = ā4 1 + 2ā2 1 a 2b 1 + a 2 2 b2 1 + 2a 2b 1 c 2 a 2 b 1 + ā 2 1 a 2 x 0 = a 2 b 1 + ā 2 =: x2 y 0 = 1 a 2 b 1 + ā 2 =: y2. 1 ā 1 < 0 It turns out that under the condition a 2 = 0 and 6a 2 c 1 + 2 b 2 c 2 4ā 1 c 2 = 0 the origin of system (3.7) is a center if and only if ā 1 = b 2 c 1 a 2 = ā 1 c 2 a 2 b 1 + ā 2 1 = 0 ā 4 1 + 2ā2 1 a 2b 1 + a 2 2 b2 1 + 2a 2b 1 c 2 a 2 b 1 + ā 2 1 < 0. (3.15) If (3.15) is true then Q T has a unique center y 2 S 2 H + 3 such that φ3 +(y 2 ) = (x2 y 2 1). By the symmetry we know that Q T also has a unique center y 2 S 2 H 3 if the condition (3.15) is satisfied. Next assume that 6a 2 c 1 + 2 b 2 c 2 4ā 1 c 2 = 0. The equations (3.8) and (3.10) have a unique solution x 0 = Substituting into (3.11) yields a 2 (ā 1 + b 2 ) 6a 2 c 1 + 2 b 2 c 2 4ā 1 c 2 =: x 3 y 0 = ( b 2 2ā 1 )( b 2 + ā 1 ) 6(3a 2 c 1 + b 2 c 2 2ā 1 c 2 ) =: y 3. 2ā 3 1 3ā2 1 b 2 + 9ā 1 a 2 b 1 3ā 1 b2 2 36ā 1 c 2 + 9a 2 b 1 b2 + 2 b 3 2 + 54a 2 c 1 + 18 b 2 c 2 = 0. Moreover it follows that = 1 9(3a 2 c 1 + b 2 c 2 2ā 1 c 2 ) (9ā2 1 a 2c 1 + 27a 2 2 b 1c 1 9ā 1 a 2 b2 c 1 + 9a 2 b2 2 c 1 8ā 3 1 c 2 This complete the proof. 27ā 1 a 2 b 1 c 2 + 12ā 2 1 b 2 c 2 6ā 1 b2 2 c 2 + b 3 2 c 2) < 0. Next we are going to study the singularities of Q T on the equator. Proposition 3.7. (i) If c 1 = c 2 = 0 then the equator is a singular circle of Q T. In other words each point of the equator is a singularity of Q T. (ii) If c 2 1 + c2 2 = 0 then on the equator Q T has two singularities at E 1 = E and E 2 = E respectively where E c 2 / c 2 1 + c2 2 c 1/ ( ) c 21 + c22 0. The direction of Q T on the equator is shown in Figure 3.1. Proof. The conclusion follows directly from Q T (±p1 ±p 2 0) = (0 0 ±2c 1 p 1 ± 2c 2 p 2 ). Proposition 3.8. Q T has centers on the equator if and only if J(c 1 c 2 ) = 0. Moreover if J(c 1 c 2 ) > 0 (resp. J(c 1 c 2 ) < 0) then Q T has a unique center at E 1 (resp. E 2 ).

Centers of projective quasi-homogeneous vector fields 13 E E Figure 3.1: The direction of Q T on the equator. Proof. By Proposition 3.7 the equator contains centers only if c 2 1 + c2 2 = 0. In what follows we will split our discussion into three cases. Case 1. c 2 > 0. Obviously E 1 H + 1 E 2 H 1. By taking the transformation φ1 + : H + 1 S2 Π + 1 we obtain the induced system of Q T on Π + 1 : d x 2 d τ = x 3(b 1 + (b 2 a 1 ) x 2 a 2 x 2 2 + ηx3) 2 =: p + 1 ( x 2 x 3 ) ( d x 3 d τ = c 1 + c 2 x 2 + c 0 1 ) 2 a 1 x 3 2 1 2 a 2 x 2 x 3 2 =: q + 1 ( x 2 x 3 ) (3.16) where η = 0 1 and d τ = (2y 2 1 + 2y2 2 + y2 3 ) y 1 dτ. In particular φ 1 +(E 2 ) = (1 c 1 /c 2 0). The characteristic equation of the linear approximation system of (3.16) at the critical point ( c 1 /c 2 0) is λ 2 + J(c 1 c 2 )/c 2 = 0. Therefore if J(c 1 c 2 ) < 0 then ( c 1 /c 2 0) is a saddle of system (3.16). If J(c 1 c 2 ) > 0 then ( c 1 /c 2 0) is a center or a focus. Noting that (p + 1 ( x 2 x 3 ) q + 1 ( x 2 x 3 )) = ( p + 1 ( x 2 x 3 ) q + 1 ( x 2 x 3 )) we conclude that the critical point ( c 1 /c 2 0) is a center. If J(c 1 c 2 ) = 0 then the equation (3.16) becomes ( d x 2 d τ = x 3 x 2 + c ) ( 1 a 2 x 2 b ) 1c 2 + η x 3 3 c 2 c 1 d x 3 d τ = 1 [ ( 2c 2 2 x 2 + c ) ( 1 + (a 2 c 1 a 1 c 2 + 2c 0 c 2 ) x 23 a 2 c 2 x 2 + c ) ] (3.17) 1 x 3 2. c 2 c 2 c 2 In the case η = 0 the straight line x 2 = c 1 /c 2 is an invariant line of system (3.16) when a 2 c 1 a 1 c 2 + 2c 0 c 2 = 0 and it is a singular line when a 2 c 1 a 1 c 2 + 2c 0 c 2. In any case the critical point ( c 1 /c 2 0) cannot be a center. When η = 1 on the straight line x 2 = c 1 /c 2 we have d x 2 /d τ = x 3 3 meaning that the orbits of system (3.17) pass through the straight line x 2 = c 1 /c 2 from the left to the right on the upper half-plane and from the right to the left on the lower half-plane. On the other hand noting also that the direction of vector field (3.17) at the x 2 axis is upward on the right-hand side of the critical point ( c 1 /c 2 0) and is downward on the left-hand side of the critical point ( c 1 /c 2 0). See Figure 3.2. We conclude that there is no closed orbit around the singularity ( c 1 /c 2 0). So the singularity ( c 1 /c 2 0) is not a center.

14 H. Liang and J. Torregrosa x 3 x 2 x 2 = c 1 /c 2 Figure 3.2: The direction of system (3.17). Next we study the singular point E 2 of Q T. With the transformation φ 1 : H 1 S2 Π 1 we obtain the induced system of Q T on Π 1 : d x 2 d τ = x 3( b 1 + (b 2 a 1 ) x 2 + a 2 x 2 2 + η x 2 3) =: p 1 ( x 2 x 3 ) d x 3 d τ = c 1 + c 2 x 2 + (c 0 1 2 a 1) x 2 3 + 1 2 a 2 x 2 x 2 3 =: q 1 ( x 2 x 3 ) (3.18) where η = 0 1 and d τ = (2y + 1 2y2 2 + y2 3 ) y 1 dτ. Moreover φ 1 (E 2 ) = (1 c 1 /c 2 0). The characteristic equation of the linear approximation system of (3.18) at the critical point (c 1 /c 2 0) is λ 2 1 c 2 J(c 1 c 2 ) = 0 where J(x y) = a 2 x 2 + (b 2 a 1 )xy b 1 y 2. (3.19) In the same way we can easily verify that the critical point (c 1 /c 2 0) is a center if and only if J(c 1 c 2 ) < 0. Case 2. c 2 < 0. Obviously E 1 H 1 and E 2 H + 1. We make the transformation φ1 : H 1 S2 Π 1 and then we obtain the induced system of Q T on Π 1 which is system (3.18). By direct calculation we get φ (E 1 1 ) = (1 c 1 /c 2 0). The characteristic equation of the linear approximation system of (3.18) at the critical point (c 1 /c 2 0) is the equation (3.19). Thus the critical point (c 1 /c 2 0) is a center if and only if c 2 J(c 1 c 2 ) < 0. Using analogous arguments we conclude that E 2 is a center of Q T on S 2 if and only if c 2 J(c 1 c 2 ) > 0. Case 3. c 2 = 0 c 1 = 0. That is E 1 = (0 sgn(c 1 ) 0) E 2 = (0 sgn(c 1 ) 0). Suppose firstly that c 1 > 0. By taking the transformation φ 2 : H 2 S2 Π 2 we obtain the induced system of Q T on Π 2 : d x 1 d τ = x 3( a 2 + (a 1 b 2 ) x 1 + b 1 x 1 2 + η x 1 x 3) 2 d x 3 d τ = c 1 x 1 + (c 0 b 2 2 ) x2 3 + b 1 2 x 1 x 3 2 + η 2 x4 3. (3.20) Obviously φ 2 (E 1 ) = (0 0). The characteristic equation of the linear approximation system of (3.20) at the critical point (0 0) is λ 2 + a 2 c 1 = 0. By similar arguments as above we find that the critical point (0 0) is a center of system (3.20) if and only if a 2 c 1 > 0 i.e. c 1 J(c 1 0) > 0. Analogously the singularity E 2 = (0 1 0) is a center of Q T on S 2 if and only if c 1 J(c 1 0) < 0. The case that c 1 < 0 can be studied in a similar way we omit the discussion for the sake of brevity.

Centers of projective quasi-homogeneous vector fields 15 In conclusion the vector field Q T has a center on the equator if and only if J(c 1 c 2 ) = 0. Remark 3.9. If η = 0 and a 1 + b 2 = 4c 0 (b 2 a 1 ) 2 + 4a 2 b 1 < 0 then J(c 1 c 2 ) = a 2 c 2 1 + (b 2 a 1 )c 1 c 2 b 1 c 2 2 and it is different from zero except for c 1 = c 2 = 0. It follows from Proposition 3.5 that Q T has exactly three centers at (0 0 1) (0 0 1) and E or E. Proof of Theorem 3.1. The proof follows from Propositions 3.3 3.5 3.6 3.8 and Remark 3.9. 3.2 Quasi-homogeneous systems with weight (m m 1) Throughout this subsection we suppose that Q T is the projective vector field of system (2.14). Proof of Theorem 3.2. It is easy to see that the equator of S 2 is a singular great circle of Q T. Thus the center (if exists) is located on S 2 H ± 3. A direct computation shows that the induced systems of Q T on Π + 3 and Π 3 are respectively d x 1 d τ = (a d x 2 1 m) x 1 + a 2 x 2 d τ = η + b 1 x 1 + (b 2 m) x 2 (3.21) and d x 1 d τ = (m a 1) x 1 a 2 x 2 d x 2 d τ = ( 1)m+1 η b 1 x 1 + (m b 2 ) x 2 (3.22) where η = 0 1 and d τ = (my 2 1 + my2 2 + y2 3 ) y 3 dτ. System (3.21) (resp. (3.22)) has a center if and only if a 1 + b 2 = 2m (a 1 m)(b 2 m) a 2 b 1 > 0 which is equivalent to (3.2). If (3.2) is true then the center is ( ηa 2 / 1 η(m a 1 )/ 1 ) (resp. (( 1) m ηa 2 / 1 ( 1) m η(m a 1 )/ 1 )). Let y + = (y + 1 y+ 2 y+ 3 ) S2 such that Then ( ηa2 η(m a ) ( 1) y 1 = φ 3 1 +(y + + 1 ) = 1 (y + y + ) 2 3 )m (y + 1 y + 3 )m 3 > 0. ( (y + 1 y+ 2 y+ 3 ) = a2 λ0 m (m a 1)λ m ) 0 λ 0 η = 1 1 1 or (y + 1 y+ 2 y+ 3 ) = (0 0 1) η = 0. Similarly let y = (y 1 y 2 y 3 ) S2 such that Then ( ( 1) m ) ηa 2 ( 1)m η(m a 1 ) 1 ) = φ 1 (y 3 ) = 1 ( y 1 (y y 2 3 )m (y 3 ) 1 )m ( ( 1) (y m 1 y 2 y 3 ) = a 2 λ0 m ( 1)m (m a 1 )λ m ) 0 λ 0 η = 1 1 y 3 < 0. or (y 1 y 2 y 3 ) = (0 0 1) η = 0 where λ 0 is the unique positive solution of (3.3). 4 Center of the quasi-homogeneous systems with weight (2 2 3) and (3 3 2) This section is devoted to derive the sufficient and necessary conditions for the projective vector field Q T of systems (2.15) and (2.16) to possess at least one center. The main results of this section are the following two theorems.

16 H. Liang and J. Torregrosa Theorem 4.1. Suppose that Q T is the projective vector field of system (2.15) then Q T has at least one center on S 2 if and only if one of the following conditions holds: (1) c 1 = 0 2 = c 2 2 4c 1c 3 > 0 2c 1 + c 2 + 2 = 0 (2) c 1 = 0 c 2 < 0 or c 1 = 0 c 2 > 0 c 2 + c 3 > 0. Moreover (i) If (1) is satisfied and 2c 1 + c 2 + 2 > 0 (resp. < 0) then Q T has exactly two centers at ±G 1 (resp. ±G 2 ) where G i = c 2 + ( 1) i 2 4c 21 + (( 1)i c 2 2c 1 2 ) 4c 2 21 + (( 1)i c 2 0 i = 1 2. (4.1) 2 ) 2 (ii) Suppose that (2) is satisfied. If c 2 < 0 c 2 + c 3 > 0 then Q T has exactly four centers at ±(0 0 1) and ±D; If c 2 < 0 c 2 + c 3 0 then Q T has exactly two centers at ±(0 0 1); if c 2 + c 3 > 0 c 2 > 0 then Q T has exactly two centers at ±D. Here D = c 3 c 2 2 + c2 3 c 2 c 2 2 + c2 3 0. Theorem 4.2. The projective vector field Q T of system (2.16) has no centers on S 2. To prove our results we need the following lemma which is a part of the Nilpotent Singular Points Theorem. The readers are referred to [3] for the complete result. Lemma 4.3. Let (0 0) be the isolate singularity of system dx = y + A(x y) dy = B(x y) (4.2) where A and B are analytic in a neighborhood of (0 0) and also j 1 A(0 0) = j 1 B(0 0) = 0. Let y = f (x) be the solution y + A(x y) = 0 in a neighborhood of the point (0 0). And let F(x) = B(x f (x)) G(x) = ( A/ x + B/ y)(x f (x)). If F(x) = ax m + 0(x m ) and G(x) = bx n + 0(x n ) with m n N m 2 n 1 and ab = 0 then we have (i) if m is even and m < 2n + 1 then the origin of system (4.2) is a cusp. (ii) if m is even and m > 2n + 1 then the origin of system (4.2) is a saddle-node. Let us consider firstly system (2.15). Proof of Theorem 4.1. The induced systems of Q T on Π + 3 is d x 1 d τ = 1 2 3 x 1(c 1 x 2 1 + c 2 x 1 x 2 + c 3 x 2 2) d x 2 d τ = 1 2 3 x 2(c 1 x 2 1 + c 2 x 1 x 2 + c 3 x 2 2). (4.3) System (4.3) has an isolated singularity if and only if c 1 + c 2 + c 3 = 0. Assume that c 1 + c 2 + c 3 = 0 then system (4.3) has a unique isolated singularity at the point ( E 3 1/3 (2c 1 + 2c 2 + 2c 3 ) 1/3 3 1/3 (2c 1 + 2c 2 + 2c 3 ) 1/3).

Centers of projective quasi-homogeneous vector fields 17 The characteristic equation of the linear approximation system of (4.3) at the singularity E is 3 1/3 λ 2 + 2 7/3 (c 1 + c 2 + c 3 ) 1/3 λ + 6 2/3 (c 1 + c 2 + c 3 ) 2/3 = 0. Hence it is easy to see that E is not a center of system (4.3). This mean that Q T has no centers on S 2 H + 3. By the symmetry of system (2.15) we know that Q T has also no centers on S 2 H 3. By direct computation we have Q T (y1 y 2 0) = 2(0 0 c 1 y 2 1 + c 2y 1 y 2 + c 3 y 2 2) with y 2 1 + y2 2 = 1. Therefore (i) if c 2 2 4c 1c 3 < 0 then Q T has no singularities on the ( equator; (ii) if c 2 2 ) 4c 1c 3 = 0 and c 2 1 + c2 2 = 0 then Q T has exactly two singularities at G c2 0 2c 1 0 and G 4c 2 1 +c 2 2 4c 2 0 1 +c 2 2 respectively; If c 1 = c 2 = 0 then Q T has exactly two singularities at (1 0 0) and ( 1 0 0) respectively; (iii) if c 2 2 4c 1c 3 > 0 then Q T has exactly four singularities ±G 1 and ±G 2 on the equator where ±G i (i = 1 2) are defined in (4.1). Suppose that c 1 > 0. The induced system of Q T giving by φ+ 2 : H + 2 S2 Π + 2 is Let d x 1 d τ = x 3 x 1 x 3 ( x 0 1 0) = d x 3 d τ = c 3 + c 2 x 1 + c 1 x 2 1 3 2 x2 3. (4.4) { φ 2 + (G i ) i = 1 2 if c 2 2 4c 1c 3 > 0 φ 2 +(G 0 ) i = 1 2 if c 2 2 4c 1c 3 = 0. Taking the transformation u = x 1 x 0 v = x 3 system (4.4) changes to du d τ = (1 x 0)v uv dv d τ = (c 2 + 2c 1 x 0 )u + c 1 u 2 3 2 v2. (4.5) The singularity G i (i = 0 1 2) is a center of Q T if and only if the origin is a center of system (4.5). If c 2 2 4c 1c 3 = 0 then c 2 + 2c 1 x 0 = 0. Obviously in this case the origin of system (4.5) is not a center if 1 x 0 = 0. Thus we assume that 1 x 0 = 0 which means that by the time rescaling system (4.5) can be transformed to the same form as (4.2). By the result of Lemma 4.3 we know that the origin of system (4.5) is a cusp. If c 2 2 4c 1c 3 > 0 then c 2 + 2c 1 x 0 = 0. By Lemma 3.4 the origin is a center of system (4.5) if and only if (c 2 + 2c 1 x 0 )(1 x 0 ) < 0. Therefore using the explicit expression of x 0 G i (and hence G i ) is center of Q T if and only if ( ) ( 1) i c 2 2 4c 1c 3 2c 1 + c 2 ( 1) i c 2 2 4c 1c 3 2c 1 < 0 i = 1 2. In other words ±G 1 are centers of Q T if and only if ) c 2 2 4c 1 c 3 > 0 c 1 (2c 1 + c 2 + c 22 4c 1c 3 > 0 and ±G 2 are centers of Q T if and only if ) c 2 2 4c 1 c 3 > 0 c 1 (2c 1 + c 2 c 22 4c 1c 3 < 0.

18 H. Liang and J. Torregrosa Similarly if c 1 < 0 then by the induced system of Q T on Π 2 we conclude that ±G 1 are centers of Q T if and only if and ±G 2 are centers of Q T if and only if ) c 2 2 4c 1 c 3 > 0 c 1 (2c 1 + c 2 + c 22 4c 1c 3 < 0 ) c 2 2 4c 1 c 3 > 0 c 1 (2c 1 + c 2 c 22 4c 1c 3 > 0. Finally consider the case that c 1 = 0. If c 2 = 0 then Q T has two pairs of singularities on the equator: ±(1 0 0) and ±D. Using the procedure as above we conclude that ±D are centers of Q T if and only if c 2 + c 3 > 0. To study the singularities ±(1 0 0) we use the induced system of Q T giving by φ+ 1 : H + 1 S2 Π + 1 which is d x 2 d τ = x 3 x 2 x 3 d x 3 d τ = c 2 x 2 + c 3 x 2 2 3 2 x2 3. (4.6) And φ 1 +(1 0 0) = (0 0 0). If c 2 = 0 then by the result of Lemma 3.4 we know that the origin is a center of system (4.6) if and only if c 2 < 0. If c 2 = 0 then ±(1 0 0) is the unique pair of singularities of Q T on the equator. By Lemma 4.3 the origin is not a center of system (4.6) meaning that Q T has no centers in this case. We are now in the position to prove the result for system (2.16). Proof of Theorem 4.2. The induced system of Q T on Π + 3 is d x 1 d τ = 1 3 2 x 1(c 1 x 1 + c 2 x 2 ) d x 2 d τ = 1 3 2 x 2(c 1 x 1 + c 2 x 2 ). (4.7) System (4.7) has an isolated singularity if and only if c 1 + c 2 > 0. And if c 1 + c 2 > 0 then system (4.3) has a unique isolated singularity at ( ) 2 2. 3c 1 + 3c 2 3c 1 + 3c 2 By direct computation we obtain that the eigenvalues of the linear approximation of system of (4.7) at that singularity are λ 1 = λ 2 = 3c1 + 3c 2 2 > 0. Hence it is not a center of system (4.3) meaning that Q T has no centers on S 2 H + 3. By the symmetry of (4.7) we know that Q T also has no centers on S 2 H 3. At the equator we have Q T (y1 y 2 0) = 3(0 0 c 1 y 1 + c 2 y 2 ) with y 2 1 + y2 2 = 1. ( ) Therefore Q T has two singularities at E c 2 / c 2 1 + c2 2 c 1/ c 21 + c22 0 and E respectively.

Centers of projective quasi-homogeneous vector fields 19 Assume firstly that c 1 > 0. Then E H + 2 S2. The induced system of Q T giving by φ 2 + : H + 2 S2 Π + 2 is d x1 d τ = x2 3 x 1 x 2 3 d x 3 d τ = c 2 + c 1 x 1 3 2 x2 3. (4.8) And φ 2 +(E) = ( c 2 /c 1 0 0). Taking the transformation u = x 3 v = c 1 x 1 + c 2 system (4.8) changes to du d τ = v 2 3 u2 =: u + A(u v) dv d τ = (c 1 + c 2 )u 2 u 2 v =: B(u v). (4.9) The singularity E is a center of Q T if and only if the origin is a center of system (4.9). Let f (u) = 2 3 u2 and F(u) = B(u f (u)) = (c 1 + c 2 )u 2 2 3 u5 G(u) = ( A/ u + B/ v)(u f (u)) = 3u 3. By the result of Lemma 4.3 we know that the origin of system (4.9) is a cusp. Therefore E (and hence E) is not a center of Q T. 5 Center of the quasi-homogeneous systems with weight (1 1 n) In this section we will study the centers of projective vector field Q T of system (1.1) with weight (1 1 n) and degree d = 2. By Theorem 1.1 of [6] the corresponding Q T of system (2.11) has no centers on S 2. Thus in the rest of this section we assume that Q T is the projective vector field of system (2.12). The main result of this section is the following theorem. Theorem 5.1. The Q T of system (2.12) has at least one center on S 2 if and only if one of the following conditions is satisfied: (1) a 3 = 0 2B 3 + C 2 = 0 G(x 1 ) = 0 F 1(x 1 ) = 0 F 2(x 1 ) = 0 (2B 3 C 2 ) 2 + 8d 0 < 0 with x 1 = (B 2 + C 1 )/(2B 3 + C 2 ) (2) a 3 = 2B 3 + C 2 = B 2 + C 1 = 0 G(x 2 ) = 0 F 1(x 2 ) = 0 2B2 3 + d 0 < 0 with x 2 = (3B 2 B 3 + d 1 )/(6B 2 3 + 3d 0) (3) a 3 = 0 2B 3 + C 2 = 0 G(x 3 ) = 0 F 1(x 3 ) < 0 F 2(x 3 ) = 0 with x 3 = (B 2 + C 1 )/(2B 3 + C 2 ) (4) a 3 = 2B 3 + C 2 = B 2 + C 1 = 0 and there exists x 4 such that G(x 4 ) = 0 F 1(x 4 ) < 0 (5) a 3 = 0 and there exists x 5 such that D(x 5 ) = F 4(x 5 ) = F 3(x 5 ) = G(x 5 ) = 0 F 1(x 5 ) < 0 where B 2 = b 2 a 1 B 3 = b 3 a 2 C 1 = c 1 2a 1 C 2 = c 2 2a 2 and G(x) = d 3 b 1 c 1 + (d 2 b 1 C 2 B 2 C 1 )x + (d 1 + 2a 3 b 1 B 3 C 1 B 2 C 2 )x 2 + (d 0 + 2a 3 B 2 + a 3 C 1 B 3 C 2 )x 3 + a 3 (2B 3 + C 2 )x 4 2a 2 3 x5 D(x) = B 2 + C 1 + (2B 3 + C 2 )x 5a 3 x 2 F 1 (x) = B 2 2 b 1 C 2 + d 2 + (4B 2 B 3 B 2 C 2 + 4a 3 b 1 + 2d 1 )x + (4B 2 3 B 3 C 2 2a 3 B 2 + 3d 0 )x 2 a 3 (8B 3 C 2 )x 3 + 5a 2 3 x4 F 2 (x) = B 2 B 3 + B 2 C 2 d 1 + (2B 3 C 2 2B 2 3 3d 0 )x F 3 (x) = 5a 3 F 1 (x) (B 3 + C 2 7a 3 x)f 2 (x) F 4 (x) = 13B 3 C 2 25a 3 B 2 12B 2 3 3C 2 2 25d 0 15a 3 (2B 3 + C 2 )x + 75a 2 3 x2. Moreover if the condition (i) (i {1 2 3 4 5}) holds then (±y 1 ±y 2 y 3 ) S2 are two centers of Q T if and only if y 2 = y 1 x i and y 3 = (y 1 )2 f (x i ) where f (x) = a 3x 3 B 3 x 2 B 2 x b 1.

20 H. Liang and J. Torregrosa We obtain from Theorem 5.1 the following corollary. Corollary 5.2. The set of number of centers of the projective vector field Q T of system (2.12) on S 2 for all possible Q with weight (1 1 2) and degree 2 is {0 2 4}. Condition (5) of Theorem 5.1 can be replaced by the following criterion which is more convenient to be manipulated with a computer. Corollary 5.3. The Q T of system (2.12) with a 3 = 0 has at least one center on S 2 if and only if one of the following conditions is satisfied: 1. β 1 = 0 D(x 51 ) = 0 F 1(x 51 ) < 0 F 3(x 51 ) = 0 F 4(x 51 ) = 0 with x 51 = α 1/β 1 2. β 1 = α 1 = 0 3 0 F 1 (x 52 ) < 0 F 3(x 52 ) = 0 F 4(x 52 ) = 0 with x 52 = (2B 3 + C 2 + 3 )/(10a 3 ) or x 52 = (2B 3 + C 2 3 )/(10a 3 ) where 3 = (2B 3 + C 2 ) 2 + 20a 3 (B 2 + C 1 ) α 1 = 250a 2 3 b 1B 2 + 110a 3 B2 2 B 3 + 24B 2 B3 3 375a 2 3 b 1C 1 + 45a 3 B 2 B 3 C 1 + 24B3 3 C 1 65a 3 B 3 C1 2 70a 3B2 2 C 2 14B 2 B3 2 C 2 40a 3 B 2 C 1 C 2 14B3 2 C 1C 2 + 30a 3 C1 2 C 2 7B 2 B 3 C2 2 7B 3 C 1 C2 2 + 3B 2 C2 3 + 3C 1 C2 3 + 50B 2 B 3 d 0 + 50B 3 C 1 d 0 + 25B 2 C 2 d 0 + 25C 1 C 2 d 0 + 125a 3 B 2 d 1 + 125a 3 C 1 d 1 + 625a 2 3 d 3 and β 1 = (200a 2 3 B2 2 + 500a 2 3 b 1B 3 + 280a 3 B 2 B 2 3 + 48B 4 3 350a 2 3 B 2C 1 70a 3 B 2 3 C 1 + 75a 2 3 C2 1 375a2 3 b 1C 2 95a 3 B 2 B 3 C 2 4B 3 3 C 2 70a 3 B 3 C 1 C 2 55a 3 B 2 C 2 2 28B 2 3 C2 2 + 45a 3 C 1 C 2 2 B 3 C 3 2 + 3C 4 2 + 125a 3 B 2 d 0 + 100B 2 3 d 0 + 125a 3 C 1 d 0 + 100B 3 C 2 d 0 + 25C 2 2 d 0 + 250a 3 B 3 d 1 + 125a 3 C 2 d 1 + 625a 2 3 d 2). Before proving the above results we give some necessary information about the projective vector field Q T. Firstly we have Q T y1 =0 = (a 2 (1 y 4 3) y 3 y 2 p 4 (y 2 y 3 ) q 4 (y 2 y 3 )) with y 2 2 + y 2 3 = 1 where p 4 (y 2 y 3 ) and q 4 (y 2 y 3 ) are two polynomials of degree not more than 4. Therefore Q T has no singularities at S 1 := {(y 1 y 2 y 3 ) S 2 y 1 = 0} if a 2 = 0. Suppose that a 2 = 0 then the first component of Q T is identically zero. This means that S 1 is invariant under the vector field Q T. In any case Q T has no centers at S 1. Next let us study the singularity of Q T on S 2 \ S 1. By the symmetry (see Proposition 2.1) it is enough to study the singularity on S 2 H + 1. The induced system of Q T on Π + 1 is d x 2 d τ = b 1 + B 2 x 2 + x 3 + B 3 x 2 2 a 3 x 3 2 d x 3 d τ = d 3 + d 2 x 2 + C 1 x 3 + d 1 x 2 2 + C 2 x 2 x 3 + d 0 x 3 2 2a 3 x 2 2 x 3. Suppose that (x 0 y 0 ) is a singularity of system (5.1) i.e. y 0 = f (x 0 ) = a 3 x 3 0 b 1 B 2 x 0 B 3 x 2 0 G(x 0 ) = d 3 b 1 c 1 + (d 2 b 1 C 2 B 2 C 1 )x 0 + (d 1 + 2a 3 b 1 B 3 C 1 B 2 C 2 )x 2 0 + (d 0 + 2a 3 B 2 + a 3 C 1 B 3 C 2 )x 3 0 + a 3 (2B 3 + C 2 )x 4 0 2a 2 3 x5 0 = 0. (5.1) (5.2)

Centers of projective quasi-homogeneous vector fields 21 Taking the transformation u = x 2 x 0 v = x 3 y 0 we change system (5.1) to where du d τ = au + v + bu2 a 3 u 3 dv d τ = cu + āv + du2 + euv + d 0 u 3 2a 3 u 2 v (5.3) a = B 2 +2B 3 x 0 3a 3 x 2 0 ā = C 1+C 2 x 0 2a 3 x 2 0 b = B 3 3a 3 x 0 c = d 2 b 1 C 2 +(4a 3 b 1 B 2 C 2 +2d 1 )x 0 +(4a 3 B 2 B 3 C 2 +3d 0 )x 2 0+a 3 (4B 3 +C 2 )x 3 0 4a 2 3 x4 0 d = d 1 +2a 3 b 1 +(2a 3 B 2 +3d 0 )x 0 +2a 3 B 3 x 2 0 2a 2 3 x3 0 e = C 2 4a 3 x 0. (5.4) Thus the point (x 0 y 0 ) is a center of system (5.1) if and only if system (5.3) has a center at the origin. By regarding x 0 and y 0 as the parameters of system (5.3) we have the following result. Lemma 5.4. The origin is a center of system (5.3) if and only if (x 0 y 0 ) satisfies one of the following conditions: (1) a 3 = 0 D(x 0 ) = 0 and α 1 β 1 x 0 = 0; (2) a 3 = 0 2B 3 + C 2 = 0 x 0 = (B 2 + C 1 )/(2B 3 + C 2 ) and G(x 0 ) = 0; (3) a 3 = 2B 3 + C 2 = B 2 + C 1 = 0 G 0 (x 0 ) = 0 where G 0 = G a3 =2B 3 +C 2 =B 2 +C 1 =0 i.e. G 0 (x) = b 1 B 2 + d 3 + (2b 1 B 2 + B 2 2 + d 2 )x + (3B 2 B 3 + d 1 )x 2 + (2B 2 B 3 + d 0 )x 3. Proof. It is easy to see that system (5.3) has a center at the origin only if (x 0 y 0 ) satisfies the relations (5.2) and a + ā = B 2 + C 1 + (2B 3 + C 2 )x 0 5a 3 x 2 0 = D(x 0 ) = 0. (5.5) If we regard the equality (5.5) as an equation in the variable x 0 then (5.5) has solution if and only if one of the following three conditions is satisfied: (i 1 ) a 3 = 0 3 = (2B 3 + C 2 ) 2 + 20a 3 (B 2 + C 1 ) 0 (i 2 ) a 3 = 0 2B 3 + C 2 = 0 (i 3 ) a 3 = 2B 3 + C 2 = B 2 + C 1 = 0. If a 3 = 0 then we use (5.5) to reduce the power of x 0 in (5.2) and we obtain that α 1 β 1 x 0 = 0. If a 3 = 0 2B 3 + C 2 = 0 then we get from (5.5) that x 0 = (B 2 + C 1 )/(2B 3 + C 2 ). Then substituting into (5.5) yields that G( (B 2 + C 1 )/(2B 3 + C 2 )) = 0. Finally if a 3 = 2B 3 + C 2 = B 2 + C 1 = 0 then a + ā = 0 holds automatically and the function G reduces to G 0. Remark 5.5. We would like to point out here that if a 3 = 2B 3 + C 2 = B 2 + C 1 = 0 then the divergence of system (5.3) is identically zero.

22 H. Liang and J. Torregrosa Next to obtain the sufficient and necessary conditions under which the origin is a center of system (5.3) we will apply a result of [4] which provides a criterion to decide whether the origin is a center of the following Liénard differential equation ẋ = y (a 1 x + a 2 x 2 + a 3 x 3 + a 4 x 4 + a 5 x 5 ) ẏ = (b 1 x + b 2 x 2 + b 3 x 3 + b 4 x 4 + b 5 x 5 ). (5.6) Lemma 5.6 ([4]). A system of type (5.6) has either a center or a focus at the origin if and only if one of the following three conditions holds: (i) a 2 1 4b 1 < 0 (ii) a 1 = b 1 = b 2 = 0 and 2a 2 2 4b 3 < 0 (iii) a 1 = a 2 = b 1 = b 2 = b 3 = b 4 = 0 and 3a 2 3 4b 5 < 0. The next lemma gives the sufficient and necessary conditions under which system (5.6) has a center at the origin. By using the result of the above lemma and for the sake of simplicity we only consider the cases that b 1 = 1 or b 1 = b 2 = 0 b 3 = 1 or b 1 = b 2 = b 3 = b 4 = 0 b 5 = 1. Lemma 5.7 ([4]). A system of type (5.6) has a center at the origin if and only if one of the following conditions holds: (i) b 1 = 1 and (a) a 3 = a 5 = b 2 = b 4 = 0 (b) a 2 = a 3 = a 4 = a 5 = 0 (c) a 4 = a 5 = 0 3a 3 = 2a 2 b 2 3b 5 = 2b 2 2 b 3 and 3b 4 = 5b 2 b 3 (d) b 5 = 0 3a 3 = 2a 2 b 2 2a 4 = a 2 b 3 and 5a 5 = 2a 2 b 4. (ii) b 1 = b 2 = 0 b 3 = 1 a 2 2 < 2 and (a) a 3 = a 5 = b 4 = 0 (b) a 2 = a 3 = a 4 = a 5 = 0 (c) a 2 = a 3 = b 5 = 0 5a 5 = 4a 4 b 4 (d) a 4 = a 5 = 0 5a 3 = 2a 2 b 4 and 25b 5 = 6b 2 4. (iii) b 1 = b 2 = b 3 = b 4 = 0 b 5 = 1 and a 1 = a 2 = a 3 = a 5 = 0. Proof of Theorem 5.1. From now on we assume in (5.3) that one of the conditions of Lemma 5.4 is satisfied and hence ā = a. By taking the transformation x = u y = v (āu + 1 2 eu2 2 3 a 3u 3 ) system (5.3) is changed to the Liénard differential system where dx d τ = y (ā 2x 2 + ā 3 x 3 ) dy d τ = ( b 1 x + b 2 x 2 + b 3 x 3 + b 4 x 4 + b 5 x 5 ) (5.7) ā 2 = b e 2 ā 3 = 5 3 a 3 b1 = a 2 c b2 = a(e b) d b 3 = be a 3 a d 0 b4 = a 3 (2b + e) = 2a 3 ā 2 b5 = 2a 2 3. We apply Lemma 5.7 to prove our results. Noting firstly that ā 3 = 0 implies b 4 = b 5 = 0 we find that the condition (iii) of Lemma 5.7 is not true for system (5.7). Consider the case