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

Topological properties on isochronous centers of polynomial Hamiltonian differential systems

Guangfeng Dong Department of Mathematics, Jinan University, Guangzhou 510632, China donggf@jnu.edu.cn
Abstract.

In this paper, we study the topological properties of complex polynomial Hamiltonian differential systems of degree nn having an isochronous center. Firstly, we prove that if the critical level curve possessing an isochronous center contains only a single singular point, and the period 11-form does not have poles with zero residue at infinity on level curves sufficiently close to the critical curve, then the vanishing cycle associated to this center is trivial in the 1-dimensional homology group of the projective closure of a generic level curve. Our result provides a positive answer to a question asked by L. Gavrilov under relatively simple conditions and can be applied to achieve an equivalent description of the Jacobian conjecture on 2\mathbb{C}^{2}. Secondly, we obtain a very simple but useful necessary condition for isochronicity of Hamiltonian systems, which is that the (n+1)(n+1)-degree part of the Hamiltonian function must have a factor with multiplicity no less than (n+1)/2(n+1)/2. Thirdly, we show a relation between Gavrilov’s question and the conjecture proposed by X. Jarque and J. Villadelprat on the non-isochronicity of real Hamiltonian systems of even degree nn.

Key words and phrases:
Hamiltonian differential systems; isochronous center; vanishing cycle; Jacobian conjecture
2010 Mathematics Subject Classification:
Primary: 34M35, 34C05; Secondary: 34C08;

1. Introduction and main results

Consider the following complex polynomial Hamiltonian differential systems of degree nn

(1.5) (dxdtdydt)=(HyHx),(x,y)2,t,\displaystyle\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\\ \end{array}\right)=\left(\begin{array}[]{r}-\frac{\partial H}{\partial y}\\ \frac{\partial H}{\partial x}\\ \end{array}\right),\ \ (x,y)\in\mathbb{C}^{2},\ t\in\mathbb{C},

where the Hamiltonian function H(x,y)H(x,y) is a polynomial of degree n+1n+1 in [x,y]\mathbb{C}[x,y]. Assuming the origin OO is a center of Morse type, without loss of generality, H(x,y)H(x,y) can be written as H(x,y)=(x2+y2)/2+h.o.t.H(x,y)=(x^{2}+y^{2})/2+h.o.t.. For a generic level curve LhL_{h} defined by the algebraic equation H(x,y)=hH(x,y)=h where hh\in\mathbb{C} is sufficiently close to 0, one can associate a vanishing cycle γh\gamma_{h} to the critical value h=0h=0, which is a 1-dimensional cycle vanishing at h=0h=0 in the 1-dimensional homology group 1(Lh,)\mathcal{H}_{1}(L_{h},\mathbb{Z}) and can be characterized by the following purely topological property: modulo orientation and the free homotopy deformation on LhL_{h}, as h0h\rightarrow 0, the cycle γh\gamma_{h} can be represented by a continuous family of loops on LhL_{h} of length that tends to zero. This description explains the terminology(see, e.g., [8]). Respectively T(h)=γh𝑑tT(h)=\oint_{\gamma_{h}}dt is called a period function of system (1.5). If T(h)T(h) is a nonzero constant independent of hh for h0h\neq 0, then the origin is called an isochronous center. This definition coincides with the classical isochronous center when (x,y)2(x,y)\in\mathbb{R}^{2} and tt\in\mathbb{R}.

One of the most important problems on isochronous centers is to describe the role of the vanishing cycle γh\gamma_{h} in the 1-dimensional homology group of the compact Riemann surface of LhL_{h}. It is still an open problem until now. In [7], L. Gavrilov has asked the following question for systems (1.5) with only isolated singularities:

Question 1.1 (Gavrilov’s question).

Is it true that if a Morse singular point is isochronous, then the associated vanishing cycle represents a zero homology cycle on the Riemann surface of the level curve LhL_{h}?

In general cases, the above question has a negative answer. Example 3.23 in reference [4] provides a system with

H(x,y)=x2(x2+2)(x2+4)+2x2(x2+1)(x2+2)(x2+3)y+(x2+1)4(x2+2)y2,H(x,y)=x^{2}(x^{2}+2)(x^{2}+4)+2x^{2}(x^{2}+1)(x^{2}+2)(x^{2}+3)y+(x^{2}+1)^{4}(x^{2}+2)y^{2},

which has an isochronous center at the origin, but the corresponding vanishing cycle is not homologous to zero on the Riemann surface of LhL_{h}. In this counterexample, it is not difficult to see that the critical level curve L0L_{0} contains at least three different singularities on 2\mathbb{C}^{2}.

What conditions can give a positive answer to Gavrilov’s question? This is also an important and meaningful question, especially it is closely related with the famous Jacobian conjecture on 2\mathbb{C}^{2}, which asserts that the following polynomial map with a constant Jacobian determinant

(1.8) ΦP:22(x,y)(f(x,y),g(x,y))\displaystyle\begin{array}[]{lrll}\Phi_{P}:&\mathbb{C}^{2}&\longrightarrow&\mathbb{C}^{2}\\ &(x,y)&\longmapsto&(f(x,y),g(x,y))\end{array}

is a global homeomorphism, where f=x+h.o.t.f=x+h.o.t. and g=y+h.o.t.g=y+h.o.t. are polynomials in [x,y]\mathbb{C}[x,y]. At present it has been proved only when the degrees of ff and gg are not too large. Obviously the map ΦP\Phi_{P} induces a Hamiltonian system

(1.15) (dxdtdydt)=(ffyggyffx+ggx)=(HyHx)\displaystyle\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\\ \end{array}\right)=\left(\begin{array}[2]{r}-f\frac{\partial f}{\partial y}-g\frac{\partial g}{\partial y}\\ f\frac{\partial f}{\partial x}+g\frac{\partial g}{\partial x}\end{array}\right)=\left(\begin{array}[2]{r}-\frac{\partial H}{\partial y}\\ \frac{\partial H}{\partial x}\end{array}\right)

having an isochronous center of Morse type at the origin with the Hamiltonian function H(x,y)=(f2+g2)/2H(x,y)=(f^{2}+g^{2})/2.

Also in [7], Proposition 6.1 says that if the vanishing cycle associated to the origin for system (1.15) represents a zero homology cycle on the Riemann surface of a generic level curve, then the map ΦP\Phi_{P} is injective, which suffices to guarantee the Jacobian conjecture is true. In addition, he has also proved that(Theorem 4.1 of [7]) Question 1.1 has a positive answer under the conditions that the critical level curve L0L_{0} contains only a single singular point which is isochronous and H(x,y)H(x,y) is a ‘good’ polynomial having only isolated and simple singularities, where the definition of a good polynomial depends on the Milnor numbers of the complex projective closure Lh¯\overline{L_{h}} of LhL_{h} at infinity.

This paper is devoted to look for other conditions to give a positive answer to Question 1.1. Denote by

ω=dt=dxHy,Hy=Hy,\omega=dt=-\frac{dx}{H_{y}},\ \ H_{y}=\frac{\partial H}{\partial y},

the period 11-form of system (1.5). We have the following main theorem.

Theorem 1.2.

For system (1.5), if the critical level curve L0L_{0} contains a single singularity which is an isochronous center of Morse type, and the period 11-form ω\omega does not have poles with zero residue at infinity for any hh sufficiently close to 0, then the associated vanishing cycle γh\gamma_{h} is trivial in 1(Lh¯,)\mathcal{H}_{1}(\overline{L_{h}},\mathbb{Z}).

Applying the above theorem to system (1.15), one can achieve an equivalent description of the Jacobian conjecture.

Corollary 1.3.

The polynomial map ΦP\Phi_{P} with constant Jacobian determinant is a global homeomorphism, if and only if two algebraic curves f=0f=0 and g=0g=0 intersect only at a single point on 2\mathbb{C}^{2}.

To prove Theorem 1.2, we will carefully study some real systems induced by complex system (1.5) and the corresponding transformation linearizing an isochronous center. Such systems possess many good properties, such as commutativity, transversality, and so on. Besides, their topological structures near the points at infinity on LhL_{h} can also provide for us a lot of information for the isochronicity of system (1.5). Letting Hn+1(x,y)H_{n+1}(x,y) be the highest degree part of H(x,y)H(x,y), we have the following necessary condition for isochronicity:

Theorem 1.4.

For system (1.5), if the origin is an isochronous center, then Hn+1H_{n+1} must have a factor with multiplicity no less than (n+1)/2(n+1)/2.

In this paper, we will also show an interesting relation between Gavrilov’s question and the following conjecture, which was claimed by X. Jarque and J. Villadelprat in [9], on real systems (1.5), i.e., (x,y)2,(x,y)\in\mathbb{R}^{2}, and tt\in\mathbb{R}.

Conjecture 1.5 (Jarque-Villadelprat conjecture).

If nn is even, then the real system (1.5) has no isochronous centers.

At present, this conjecture is still open and a recent development can be found in [5]. The following theorem indicates that if the Jarque-Villadelprat conjecture is not true, then the Gavrilov’s question must have a negative answer for such real systems.

Theorem 1.6.

For any isochronous center of a real system (1.5) with even nn, the corresponding vanishing cycle can not be homologous to zero on the projective closure of the complexification of a generic real level curve.

The paper is organized as follows. We shall first introduce some properties on the commuting real differential systems(or real vector fields) induced by system (1.5) and provide a powerful technique to extend the transformation linearizing an isochronous center. Then we give the detailed proof of the main results and some applications.

2. Commuting real systems

Note that if the origin is an isochronous center of Morse type for system (1.5), then there exists an analytic area-preserving transformation(see, e.g, [1, 11, 12])

Φ:(x,y)(u(x,y),v(x,y))\Phi:\ (x,y)\mapsto(u(x,y),v(x,y))

changing system (1.5) to a linear system

(2.5) (dudtdvdt)=(vu),\displaystyle\left(\begin{array}[]{c}\frac{du}{dt}\\ \frac{dv}{dt}\\ \end{array}\right)=\left(\begin{array}[]{r}-v\\ u\\ \end{array}\right),

here we say Φ\Phi is area-preserving is equivalent to say its Jacobian determinant det(J(Φ))1\det(J(\Phi))\equiv 1, where

J(Φ)=(uxuyvxvy).J(\Phi)=\left(\begin{array}[]{cc}\frac{\partial u}{\partial x}&\frac{\partial u}{\partial y}\\ \frac{\partial v}{\partial x}&\frac{\partial v}{\partial y}\end{array}\right).

Generally speaking, Φ\Phi is only well defined in a small neighborhood of the origin 𝒩(O)2\mathcal{N}(O)\subseteq\mathbb{C}^{2}.

By taking advantage of constant Jacobian determinant, one can construct another complex system in 𝒩(O)\mathcal{N}(O) as follows

(2.10) (dxdtdydt)=(JTJ)1(HxHy),\displaystyle\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\\ \end{array}\right)=(J^{T}J)^{-1}\left(\begin{array}[]{c}\frac{\partial H}{\partial x}\\ \frac{\partial H}{\partial y}\\ \end{array}\right),

which can be also linearized to a linear system

(2.15) (dudtdvdt)=(uv)\displaystyle\left(\begin{array}[]{c}\frac{du}{dt}\\ \frac{dv}{dt}\\ \end{array}\right)=\left(\begin{array}[]{r}u\\ v\\ \end{array}\right)

by the same transformation Φ\Phi, for the reasons that

J1=(vyuyvxux)\displaystyle J^{-1}=\left(\begin{array}[]{rr}\frac{\partial v}{\partial y}&-\frac{\partial u}{\partial y}\\ -\frac{\partial v}{\partial x}&\frac{\partial u}{\partial x}\end{array}\right)

and

(J1)T(HxHy)=(vyvxuyux)(HxHy)=(uv).\displaystyle(J^{-1})^{T}\left(\begin{array}[]{c}\frac{\partial H}{\partial x}\\ \frac{\partial H}{\partial y}\\ \end{array}\right)=\left(\begin{array}[]{rr}\frac{\partial v}{\partial y}&-\frac{\partial v}{\partial x}\\ -\frac{\partial u}{\partial y}&\frac{\partial u}{\partial x}\end{array}\right)\left(\begin{array}[]{c}\frac{\partial H}{\partial x}\\ \frac{\partial H}{\partial y}\\ \end{array}\right)=\left(\begin{array}[]{r}u\\ v\\ \end{array}\right).

Consequently, systems (1.5) and (2.15) induce the following four real differential systems(see, e.g. [3]) by taking (x,y)24(x,y)\in\mathbb{C}^{2}\cong\mathbb{R}^{4} but tt\in\mathbb{R}:

V:(dxdtdydt)=(HyHx),iV:(dxdtdydt)=(iHyiHx),\displaystyle V:\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\end{array}\right)=\left(\begin{array}[]{r}-\frac{\partial H}{\partial y}\\ \frac{\partial H}{\partial x}\end{array}\right),\ {\rm i}V:\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\end{array}\right)=\left(\begin{array}[]{r}-{\rm i}\frac{\partial H}{\partial y}\\ {\rm i}\frac{\partial H}{\partial x}\end{array}\right),

and

Vg:(dxdtdydt)=(JTJ)1(HxHy),iVg:(dxdtdydt)=(JTJ)1(iHxiHy),\displaystyle V_{g}:\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\\ \end{array}\right)=(J^{T}J)^{-1}\left(\begin{array}[]{c}\frac{\partial H}{\partial x}\\ \frac{\partial H}{\partial y}\\ \end{array}\right),\ {\rm i}V_{g}:\left(\begin{array}[]{c}\frac{dx}{dt}\\ \frac{dy}{dt}\\ \end{array}\right)=(J^{T}J)^{-1}\left(\begin{array}[]{c}{\rm i}\frac{\partial H}{\partial x}\\ {\rm i}\frac{\partial H}{\partial y}\\ \end{array}\right),

where i2=1{\rm i}^{2}=-1. They can be transformed to the following four real linear systems simultaneously by the same Φ\Phi respectively:

VV:(dudtdvdt)=(vu),iViV:(dudtdvdt)=(iviu),\displaystyle V\hookrightarrow V_{\ast}:\left(\begin{array}[2]{c}\frac{du}{dt}\\ \frac{dv}{dt}\end{array}\right)=\left(\begin{array}[2]{r}-v\\ u\end{array}\right),\ {\rm i}V\hookrightarrow{\rm i}V_{\ast}:\left(\begin{array}[2]{c}\frac{du}{dt}\\ \frac{dv}{dt}\end{array}\right)=\left(\begin{array}[2]{r}-{\rm i}v\\ {\rm i}u\end{array}\right),

and

VgVg:(dudtdvdt)=(uv),iVgiVg:(dudtdvdt)=(iuiv).\displaystyle V_{g}\hookrightarrow V_{g\ast}:\left(\begin{array}[2]{c}\frac{du}{dt}\\ \frac{dv}{dt}\end{array}\right)=\left(\begin{array}[2]{r}u\\ v\end{array}\right),\ {\rm i}V_{g}\hookrightarrow{\rm i}V_{g\ast}:\left(\begin{array}[2]{c}\frac{du}{dt}\\ \frac{dv}{dt}\end{array}\right)=\left(\begin{array}[2]{r}{\rm i}u\\ {\rm i}v\end{array}\right).

Letting u=u1+iu2u=u_{1}+\mathrm{i}u_{2} and v=v1+iv2v=v_{1}+\mathrm{i}v_{2} and regarding 24={(u1,u2,v1,v2)}\mathbb{C}^{2}\cong\mathbb{R}^{4}=\{(u_{1},u_{2},v_{1},v_{2})\}, the coefficient matrices of VV_{\ast}, iV{\rm i}V_{\ast}, VgV_{g\ast} and iVg{\rm i}V_{g\ast} are respectively

M1=(0I2I20),M2=(0E2E20),M3=(I200I2),M4=(E200E2),\displaystyle\begin{array}[]{lcclcc}M_{1}&=&\left(\begin{array}[]{rr}0&-I_{2}\\ I_{2}&0\\ \end{array}\right),&M_{2}&=&\left(\begin{array}[]{rr}0&-E_{2}\\ E_{2}&0\\ \end{array}\right),\\ \\ M_{3}&=&\left(\begin{array}[]{rr}I_{2}&0\\ 0&I_{2}\\ \end{array}\right),&M_{4}&=&\left(\begin{array}[]{rr}E_{2}&0\\ 0&E_{2}\\ \end{array}\right),\end{array}

where

I2=(1001),E2=(0110).\displaystyle I_{2}=\left(\begin{array}[]{cc}1&0\\ 0&1\\ \end{array}\right),\ E_{2}=\left(\begin{array}[]{cc}0&-1\\ 1&0\\ \end{array}\right).

Obviously we have

MiMj=MjMi,i,j=1,2,3,4.M_{i}M_{j}=M_{j}M_{i},\ \forall i,j=1,2,3,4.

Due to that Φ\Phi is a diffeomorphism, one can get the following important properties for vector fields VV, iV{\rm i}V, VgV_{g}, and iVg\ {\rm i}V_{g}:

  1. (1)

    they are commutative pairwise everywhere in 𝒩(O)\mathcal{N}(O), i.e., as real vector fields, the Lie bracket of any two of them vanishes. So for any two points p1,p2p_{1},p_{2} in 𝒩(O)\mathcal{N}(O) except OO, it takes the same time along any two continuous paths connecting p1p_{1} and p2p_{2} consisting of finitely many trajectories of those vector fields.

  2. (2)

    their trajectories are transversal pairwise everywhere on 𝒩(O)L0\mathcal{N}(O)-L_{0}; while on L0L_{0}, VV(resp. iV{\rm i}V) coincides with iVg{\rm i}V_{g}(resp. Vg-V_{g}) on one of two branches near OO and with iVg-{\rm i}V_{g}(resp. VgV_{g}) on the other one;

  3. (3)

    the domain in which VgV_{g} and iVg{\rm i}V_{g} can be well defined is the same to the domain of Φ\Phi, but VV and iV{\rm i}V are well defined on the whole complex plane 2\mathbb{C}^{2};

  4. (4)

    near the origin, all of the orbits of system VV are closed; on the contrary, system iV{\rm i}V does not have any closed orbits in 𝒩(O)\mathcal{N}(O);

  5. (5)

    the trajectories of systems VV and iV{\rm i}V are both tangent to LhL_{h} everywhere, so their restrictions, denoted by VhV_{h} and iVh{\rm i}V_{h}, are two real systems defined well on LhL_{h}.

Denote by φ(,t)\varphi(\cdot,t)(resp. iφ,φg,iφg{\rm i}\varphi,\ \varphi_{g},\ {\rm i}\varphi_{g}, φ,iφ,φg,\varphi_{\ast},\ {\rm i}\varphi_{\ast},\ \varphi_{g\ast}, and iφg{\rm i}\varphi_{g\ast}) the flow map induced by VV(resp. iV{\rm i}V, Vg,iVgV_{g},\ {\rm i}V_{g}, VV_{\ast}, iV{\rm i}V_{\ast}, Vg,V_{g\ast}, and iVg{\rm i}V_{g\ast}), i.e., for any given point pp, φ(p,t)\varphi(p,t) takes the value at time tt of the solution of equations VV with initial value pp at t=0t=0. The commutativity between those systems means that each one of the flow maps above preserves the orbits of any other system in the domain of Φ\Phi. We shall take advantage of this observation to extend the domain of Φ\Phi to a bigger one than 𝒩(O)\mathcal{N}(O). Without loss of generality, we assume 𝒩(O)\mathcal{N}(O) is a sufficiently small and homeomorphic to a open ball {(x,y)2:|x|2+|y|2<ϵ}\{(x,y)\in\mathbb{C}^{2}:\ \left|x\right|^{2}+\left|y\right|^{2}<\epsilon\} centered at OO with radius ϵ\epsilon. Denote by HH the following map

H:2(x,y)H(x,y).\displaystyle\begin{array}[]{lrll}H:&\mathbb{C}^{2}&\longrightarrow&\mathbb{C}\\ &(x,y)&\longmapsto&H(x,y).\end{array}

Continuation technique for Φ\Phi:

For a closed orbit σ\sigma of VV in 𝒩(O)\mathcal{N}(O) such that H(σ)0H(\sigma)\neq 0, and two sufficiently small number t1,t2t_{1},t_{2}, the space

Γσ0st1iφg(0tt2φg(σ,t),s)\Gamma_{\sigma}\triangleq\cup_{0\leq s\leq t_{1}}{\rm i}\varphi_{g}\left(\cup_{0\leq t\leq t_{2}}\varphi_{g}(\sigma,t),s\right)

is a real 3-dimension sub-manifold of 𝒩(O)\mathcal{N}(O) and transversal to iV{\rm i}V at every point. Then Φ(Γσ)\Phi(\Gamma_{\sigma}) is also a real 3-dimension sub-manifold of Φ(𝒩(O))\Phi(\mathcal{N}(O)) and transversal to iV{\rm i}V_{\ast} at every point.

Along the trajectories of iV{\rm i}V passing through Γσ\Gamma_{\sigma} in 𝒩(O)\mathcal{N}(O), the transformation Φ\Phi can be expressed by the flow map iφ{\rm i}\varphi as follows: for any point pΓσ,p\in\Gamma_{\sigma}, and any sufficiently small tt, we have

(2.27) Φ(iφ(p,t))=iφ(Φ(p),t).\Phi({\rm i}\varphi(p,t))={\rm i}\varphi_{\ast}(\Phi(p),t).

Clearly the vector fields iV{\rm i}V and iV{\rm i}V_{\ast} are well defined globally, the above equation can be extended to a larger interval II for time tt\in\mathbb{R} such that iφ(p,t)𝒩(O){\rm i}\varphi(p,t)\not\in\mathcal{N}(O), if iV{\rm i}V satisfies the following two conditions:

  1. C1.

    the trajectories of iV{\rm i}V could not return into the domain where Φ\Phi has already been defined well;

  2. C2.

    there is no point PP at infinity such that iφ(p,t){\rm i}\varphi(p,t) tends to PP as tt tends a finite moment t0t_{0} for some a point p0Γσp_{0}\in\Gamma_{\sigma}.

If the trajectories of iV{\rm i}V from a open subset of Γσ\Gamma_{\sigma} go to a point at infinity when tt tends \infty, then the interval II for those points can be [0,+)[0,+\infty) or (,0](-\infty,0]. While if C1 holds but C2 not, then II can only be [0,t0)[0,t_{0}) or (t0,0](t_{0},0] at such a point p0p_{0}(see Figure 1).

Refer to caption
Figure 1. The continuation of Φ\Phi

Noticing that the vector fields VV and VV_{\ast} are also well defined globally, we can also perform the above operation along the trajectories of VV if it satisfies the conditions C1 and C2.

In a word, one can extend the transformation Φ\Phi to an open domain 𝒟\mathcal{D} as big as possible according to the above operation along the trajectories of VV and iV{\rm i}V. Although 𝒟\mathcal{D} may be much bigger than 𝒩(O)\mathcal{N}(O), we have 𝒟Lh\mathcal{D}\cap L_{h} is still homeomorphic to 𝒩(O)Lh\mathcal{N}(O)\cap L_{h} for any hh sufficiently close to 0.

3. Points at infinity

To prove the main results, we still need to know some information about the points at infinity on LhL_{h}. It is better to deal with it in the projective space 2\mathbb{CP}^{2}. Assume the projective closure Lh¯\overline{L_{h}} are defined by the following homogeneous equations

k=2n+1zn+1kHk(x,y)hzn+1=0,[x:y:z]2,\sum_{k=2}^{n+1}z^{n+1-k}H_{k}(x,y)-hz^{n+1}=0,\ [x:y:z]\in\mathbb{CP}^{2},

where HkH_{k} represents the homogeneous part of degree kk of H(x,y)H(x,y). For a generic value hh, the set of singularities on Lh¯\overline{L_{h}}, denoted by Σh\Sigma_{h}, consists of only some points at infinity on LhL_{h}. The Riemann surface of LhL_{h} coincides with the resolution of Lh¯\overline{L_{h}} by a birational map. Generally speaking, The algebraic curve Lh¯\overline{L_{h}} may have more than one connected branches near a point PΣhP\in\Sigma_{h}. The number of such branches is equal to the number of essentially different Puiseux expressions associated to PP(see, e.g., [10]).

Rewriting the homogeneous part Hn+1(x,y)H_{n+1}(x,y) of degree n+1n+1 as follows:

(3.1) Hn+1(x,y)=i=1N(αixβiy)ni,ni1,i=1Nni=n+1,\displaystyle H_{n+1}(x,y)=\prod_{i=1}^{N}\left(\alpha_{i}x-\beta_{i}y\right)^{n_{i}},\ \ n_{i}\geq 1,\ \ \sum_{i=1}^{N}n_{i}=n+1,

where αi,βi\alpha_{i},\ \beta_{i}\in\mathbb{C} such that αi:βiαj:βj\alpha_{i}:\beta_{i}\neq\alpha_{j}:\beta_{j} if ij,i\neq j, the projective coordinate of a point PiΣhP^{i}\in\Sigma_{h} can be represented by [βi:αi:0][\beta_{i}:\alpha_{i}:0]. Up to a projective change of coordinates, we can always assume its projective coordinate is [1:0:0][1:0:0]. Then it is convenient to adopt a pair of new affine coordinates (X,Y)(X,Y), where

X=1x,Y=yx,X=\frac{1}{x},\ Y=\frac{y}{x},

and the Puiseux expressions near PiP^{i} are totally determined by the Puiseux expressions of equation

(3.2) Hh(X,Y)=Xn+1H(1X,YX)hXn+1=0\displaystyle H_{h}^{\ast}(X,Y)=X^{n+1}H\left(\frac{1}{X},\frac{Y}{X}\right)-hX^{n+1}=0

near the origin. According to the classical theory of Puiseux(see, e.g., ([6])), each branch of an algebraic curve near a singularity can be parameterized by a Puiseux series of the following form.

Lemma 3.1 (Puiseux).

If Hh(0,0)=0H_{h}^{\ast}(0,0)=0 and Hh(0,Y)0H_{h}^{\ast}(0,Y)\neq 0, then there exist numbers 𝗉,𝗊+\mathsf{p},\mathsf{q}\in\mathbb{Z}_{+}, a parameter ss\in\mathbb{C}, and a a holomorphic function ρ(s)=s𝗊(c0+i=1+cisi),c00,\rho(s)=s^{\mathsf{q}}(c_{0}+\sum_{i=1}^{+\infty}c_{i}s^{i}),\ c_{0}\neq 0, such that Hh(s𝗉,s𝗊ρ(s))=0H_{h}^{\ast}(s^{\mathsf{p}},s^{\mathsf{q}}\rho(s))=0 for all ss in a neighbourhood of 0.

In general the coefficients {ci}\{c_{i}\} may depend on hh on different level curve LhL_{h}, so sometimes we replace ρ(s)\rho(s) with ρ(s,h)\rho(s,h) to emphasize it. Taking the Puiseux parameterization x=s𝗉,y=s𝗊𝗉ρ(s,h)x=s^{-\mathsf{p}},\ y=s^{\mathsf{q}-\mathsf{p}}\rho(s,h) into system (1.5), we obtain a complex 1-dimension ordinary differential equation

(3.3) dsdt=s𝗉+1𝗉Hy(s𝗉,s𝗊𝗉ρ(s,h))=λs𝗄+o(s𝗄)\displaystyle\frac{ds}{dt}=\frac{s^{\mathsf{p}+1}}{\mathsf{p}}\frac{\partial H}{\partial y}(s^{-\mathsf{p}},s^{\mathsf{q}-\mathsf{p}}\rho(s,h))=\lambda s^{\mathsf{k}}+o(s^{\mathsf{k}})

on a branch of Lh¯\overline{L_{h}} near PiP^{i}, where λ0,𝗄.\lambda\not=0,\ \mathsf{k}\in\mathbb{Z}. Then the real systems VhV_{h} and iVh{\rm i}V_{h} are changed to the following forms respectively under this parameterization:

(3.4) Vh:dsdt=λs𝗄+o(s𝗄),t,\displaystyle V_{h}:\ \ \frac{ds}{dt}=\lambda s^{\mathsf{k}}+o(s^{\mathsf{k}}),\ t\in\mathbb{R},

and

(3.5) iVh:dsdt=iλs𝗄+o(s𝗄),t.\displaystyle{\mathrm{i}}V_{h}:\ \ \frac{ds}{dt}={\mathrm{i}}\lambda s^{\mathsf{k}}+o(s^{\mathsf{k}}),\ t\in\mathbb{R}.

Their topological structures can be classified into the following four classes according to the value of 𝗄\mathsf{k} near s=0s=0:

  • 𝗄>1\mathsf{k}>1. The orbits of real system (3.4)(or system (3.5)) form 2(𝗄1)2(\mathsf{k}-1) petals in a sufficiently small neighborhood of s=0s=0, any one of them is tangent to a separatrix of the petals at s=0s=0.

  • 𝗄=1\mathsf{k}=1. If λ\lambda is a pure imaginary number, the point s=0s=0 is of center-focus type; while for λ\lambda\in\mathbb{R}, it is a node, and for other numbers, it is a focus.

  • 𝗄=0\mathsf{k}=0. The point s=0s=0 is not a singularity for real systems (3.4) and (3.5).

  • 𝗄<0\mathsf{k}<0. The system (3.4)(or system (3.5)) has a saddle structure in a sufficiently small neighborhood of 0 except s=0s=0.

Remark 3.2.

It should be pointed out that there may exist an orbit of system (3.4) such that it can reach the origin s=0s=0 at a finite moment from a fixed point s0s\not=0. It is not difficult to see this phenomenon occurs only in the cases 𝗄=0\mathsf{k}=0 and 𝗄<0\mathsf{k}<0, and in the latter one such an orbit is just the separatrix of the saddle.

The Puiseux parameterizations can be determined completely by the so-called Newton polygon of the singularity. Given an irreducible polynomial F(X,Y)=k,lbklXkYlF(X,Y)=\sum_{k,l}b_{kl}X^{k}Y^{l} with F(0,0)=0F(0,0)=0, denote by Λ(F)\Lambda(F) the carrier of F(X,Y)F(X,Y), i.e. Λ(F)={(k,l)2|bkl0}.\Lambda(F)=\{(k,l)\in\mathbb{Z}^{2}\ |\ b_{kl}\neq 0\}. Assuming that Q1,Q22Q_{1},Q_{2}\in\mathbb{R}^{2}, let

[Q1,Q2]={σQ1+(1s)Q2| 0s1}[Q_{1},Q_{2}]=\{\sigma Q_{1}+(1-s)Q_{2}\ |\ 0\leq s\leq 1\}

be the straight line segment from Q1Q_{1} to Q2Q_{2}. Consider the convex subset AA on 2\mathbb{R}^{2} consisting of those (X,Y)2(X,Y)\in\mathbb{R}^{2} such that XX0X\geq X_{0} and YY0Y\geq Y_{0} for some (X0,Y0)[Q1,Q2](X_{0},Y_{0})\in[Q_{1},Q_{2}] where Q1,Q2Λ(F)Q_{1},Q_{2}\in\Lambda(F).

Definition 3.3 (Newton Polygon).

The boundary of set AA excluding the axes is called the Newton polygon of F(X,Y)F(X,Y) at the origin, which consists of only finitely many straight line segments.

4. Important lemmas

In this section, we first prove the following important lemmas. It is not difficult to see the vanishing cycle γh\gamma_{h} can be represented by a given closed orbit of system VhV_{h} near the origin(we still denote this orbit by γh\gamma_{h}).

Lemma 4.1.

If the origin is an isochronous center of system (1.5), then every orbit of iVh{\mathrm{i}}V_{h} passing through a point on γh\gamma_{h} is not closed for any hh sufficiently close to 0.

Proof.

Suppose otherwise, i.e., suppose there exists a point p0γhp_{0}\in\gamma_{h} such that the orbit of iVh{\mathrm{i}}V_{h} passing through p0p_{0} is closed. Then by the commutativity between VV and iV{\mathrm{i}}V, there exists a sufficiently small neighborhood 𝒩p0Γγh𝒩(O)\mathcal{N}_{p_{0}}\subset\Gamma_{\gamma_{h}}\subset\mathcal{N}(O) of p0p_{0} such that for any p𝒩p0p\in\mathcal{N}_{p_{0}}, the orbit of iV{\mathrm{i}}V passing through pp is also closed.

Consider the inverse transformation Φ~=Φ1\tilde{\Phi}=\Phi^{-1} that also has a constant Jacobian determinant 11 in the domain Φ(𝒩(O))\Phi(\mathcal{N}(O)) on the (u,v)(u,v)-plane. Since iV{\mathrm{i}}V_{\ast} satisfies the conditions C1 and C2, by using the same continuation technique introduced in Section 2, we can extend Φ~\tilde{\Phi} from Φ(𝒩(O))\Phi(\mathcal{N}(O)) to a bigger domain 𝒟~\tilde{\mathcal{D}} along the trajectories of iV{\mathrm{i}}V_{\ast} by the following equation

Φ~(iφ(q,t))iφ(Φ1(q),t),qΦ(𝒩p0),\tilde{\Phi}({\rm i}\varphi_{\ast}(q,t))\triangleq{\rm i}\varphi(\Phi^{-1}(q),t),\ \forall q\in\Phi({\mathcal{N}_{p_{0}}}),

such that Φ~(𝒟~)\tilde{\Phi}(\tilde{\mathcal{D}}) covers all closed orbits of iV{\rm i}V passing through 𝒩p0\mathcal{N}_{p_{0}}(see Figure 2).

Refer to caption
Figure 2. The continuation of Φ~\tilde{\Phi}

In the domain Φ~(𝒟~)\tilde{\Phi}(\tilde{\mathcal{D}}), the vector field VgV_{g} is well defined and commuting with iV{\mathrm{i}}V. This implies that the periods of those closed orbits of iV{\mathrm{i}}V are the same for any hH(𝒩p0)h\in H(\mathcal{N}_{p_{0}}). The above operation is valid for any hh sufficiently close to 0. So we get a series of closed orbits of iV{\mathrm{i}}V with the same period as h0h\rightarrow 0 along a trajectory of VgV_{g}, whose lengths tend to 0 since |iV|0|{\mathrm{i}}V|\rightarrow 0 when h0h\rightarrow 0. This means that such a closed orbit also represents the vanishing cycle of the isochronous center, which leads a contradiction, because the origin is of Morse type having only one vanishing cycle and the intersection number of two closed orbits of VV and iV\mathrm{i}V respectively is equal to 11 so that they can not represent the same one cycle in 1(Lh,)\mathcal{H}_{1}(L_{h},\mathbb{Z}). Thus the lemma holds. ∎

By this lemma, we have the following immdiately.

Lemma 4.2.

If the origin is an isochronous center of system (1.5), then there exists a subset γh1γh\gamma^{1}_{h}\subset\gamma_{h} consisting of at most finitely many points such that:

  1. (1)

    for any pγh1p\in\gamma^{1}_{h}, iφ(p,t){\rm i}\varphi(p,t) tends to a point PiP^{i} at infinity as tt tends to some a finite time t0t_{0}, and VhV_{h} has a form (3.4) with number 𝗄0\mathsf{k}\leq 0 on one of the branches of Lh¯\overline{L_{h}} near PiP^{i};

  2. (2)

    for any pγhγh1p\in\gamma_{h}-\gamma^{1}_{h}, iφ(p,t){\rm i}\varphi(p,t) tends to a point PiP^{i} at infinity as t±t\rightarrow\pm\infty, and VhV_{h} has a form (3.4) with number 𝗄1\mathsf{k}\geq 1 on one of the branches of Lh¯\overline{L_{h}} near PiP^{i}.

Proof.

By Lemma 4.1, if the orbit δp\delta_{p} of iVh{\mathrm{i}}V_{h} passing through a point pγhp\in\gamma_{h} can not tend any point at infinity, then there remains two possible cases:

  • δp\delta_{p} tends a closed orbit δ0\delta_{0} of iVh{\mathrm{i}}V_{h}. If such a δ0\delta_{0} exists, then it is isolated or semi-isolated. However, by the commutativity between VV and iV{\mathrm{i}}V, there exist annuli such that δ0\delta_{0} is not the boundary.

  • δp\delta_{p} is ergodic on a subset of LhL_{h}. If so, we can also extend the transformation Φ~\tilde{\Phi} to a domain 𝒟~\tilde{\mathcal{D}} such that Φ~(𝒟~)\tilde{\Phi}(\tilde{\mathcal{D}}) covers δh\delta_{h} as shown in the above lemma along δp\delta_{p}(in fact, we only need to do this on LhL_{h}). One can choose a trajectory of lhl_{h} for VV such that δplh\delta_{p}\cap l_{h} is dense in lhl_{h}. Noticing that iV{\rm i}V_{\ast} on the curve ChC_{h} defined by u2+v2=hu^{2}+v^{2}=h is integrability, there is a non trivial analytic first integral G~\tilde{G} defined on ChC_{h} such that G~(Φ~1(δh))\tilde{G}(\tilde{\Phi}^{-1}(\delta_{h})) is a constant. Defining a function G(p)G~(Φ~1(p))G(p)\triangleq\tilde{G}(\tilde{\Phi}^{-1}(p)) for pΦ~(𝒟~)p\in\tilde{\Phi}(\tilde{\mathcal{D}}), it is a non trivial analytic first integral for iV{\mathrm{i}}V such that G(lh)G(l_{h}) is not a constant. However, GG is a constant on a dense subset δplh\delta_{p}\cap l_{h} of lhl_{h}, which implies G(lh)G(l_{h}) should be also a constant. This is a contradiction.

Finally, every orbit of iVh{\mathrm{i}}V_{h} passing through a point pγhp\in\gamma_{h} can only tend to a point PiP^{i} at infinity on one of the branches of Lh¯\overline{L_{h}} near PiP^{i}. According to the arguments in Remark 3.2, if the number 𝗄0\mathsf{k}\leq 0 for PiP^{i}, then iφ(p,t){\rm i}\varphi(p,t) will reache at PiP^{i} at some a finite moment t0t_{0}. In addition, due to that the numbers of points at infinity and separatrices of the saddles are both finite, the number of such points pp are also finite. The lemma is proved. ∎

Below we shall show that, under the assumption of Theorem 1.2, in the second case of the above lemma, the number 𝗄\mathsf{k} must be equal to 11.

Lemma 4.3.

Under the assumption of Theorem 1.2, if a point PiP^{i} at infinity on LhL_{h} is the limit of iφ(p,t){\mathrm{i}}\varphi(p,t) as t+t\rightarrow+\infty(or -\infty) for some a point pγhp\in\gamma_{h} on one of the branches near PiP^{i}, then the number 𝗄=1\mathsf{k}=1 for corresponding system (3.3), and the orbits of VhV_{h} are closed encircling PiP^{i}.

Proof.

Suppose system (1.5) has a Hamiltonian H=(x2+y2)/2+h.o.tH=(x^{2}+y^{2})/2+h.o.t and PiP^{i} has a projective coordinate [βi:αi:0][\beta_{i}:\alpha_{i}:0], by a linear change of coordinates (x,y)(x1,y1)=((βi¯x+αi¯y)/ri,(αixβiy)/ri)(x,y)\mapsto(x_{1},y_{1})=(-\left(\overline{\beta_{i}}x+\overline{\alpha_{i}}y\right)/r_{i},\left(\alpha_{i}x-\beta_{i}y\right)/r_{i}), where ri=|αi|2+|βi|2r_{i}=\sqrt{|\alpha_{i}|^{2}+|\beta_{i}|^{2}}, its coordinate can be changed to [1:0:0][1:0:0]. If the linearization transformation Φ\Phi maps (x,y)(x,y) to (u,v)(u,v), by taking a linear change of coordinates

(u1v1)=12ri(αi¯iβiiαi¯+βiiαiβi¯αi+iβi¯)(uv),\left(\begin{matrix}u_{1}\\ v_{1}\end{matrix}\right)=\frac{1}{\sqrt{2}r_{i}}\left(\begin{matrix}\overline{\alpha_{i}}-{\rm i}\beta_{i}&-{\rm i}\overline{\alpha_{i}}+\beta_{i}\\ -{\rm i}\alpha_{i}-\overline{\beta_{i}}&\alpha_{i}+{\rm i}\overline{\beta_{i}}\end{matrix}\right)\left(\begin{matrix}u\\ v\end{matrix}\right),

then one of the points at infinity also has a coordinate [1:0:0][1:0:0] on (u1,v1)(u_{1},v_{1})-plane, i.e., the Hamiltonian function has the form u1v1u_{1}v_{1}.

Under the conditions of the lemma, the equation (2.27) holds for I=[0,+)I=[0,+\infty) and a sufficiently small neighborhood of pp in Γγh\Gamma_{\gamma_{h}}, i.e., the domain 𝒟\mathcal{D} where Φ\Phi is well defined can be sufficiently close to PiP^{i} along the orbits of iV{\mathrm{i}}V.

We take the coordinates of Puiseux parameters (s,h)(s,h) near PiP^{i} in 𝒟\mathcal{D}, and the coordinates (s~,h)(\tilde{s},h) near [1:0:0][1:0:0] in Φ(𝒟)\Phi(\mathcal{D}). Then Φ\Phi induces a map Ψ\Psi from an open set 𝒫\mathcal{P} in the (s,h)(s,h) plane to (s~,h)(\tilde{s},h) plane 𝒫~\tilde{\mathcal{P}}, so that the following diagram is commutative.

(4.1) 𝒟{\mathcal{D}}𝒫{\mathcal{P}}Φ(𝒟){\Phi(\mathcal{D})}𝒫~{\tilde{\mathcal{P}}}Φ\scriptstyle{\Phi}R\scriptstyle{\mathrm{R}}Ψ\scriptstyle{\Psi}R~\scriptstyle{\tilde{\mathrm{R}}}                  (x1,y1){(x_{1},y_{1})}(s,h){(s,h)}(u1,v1){(u_{1},v_{1})}(s~,h){(\tilde{s},h)}Φ\scriptstyle{\Phi}R\scriptstyle{\mathrm{R}}Ψ\scriptstyle{\Psi}R~\scriptstyle{\tilde{\mathrm{R}}}

where

(4.4) R:(s,h)(s𝗉,s𝗊𝗉ρ(s,h))R~:(s~,h)(s~1,hs~).\displaystyle\begin{array}[]{llcl}\mathrm{R}:&(s,h)&\mapsto&(s^{-\mathsf{p}},\ s^{\mathsf{q}-\mathsf{p}}\rho(s,h))\\ \tilde{\mathrm{R}}:&(\tilde{s},h)&\mapsto&(\tilde{s}^{-1},h\tilde{s})\end{array}.

are Puiseux parameterizations respectively, and

(4.7) Φ:(x1,y1)(u1,v1)Ψ:(s,h)(s~,h).\displaystyle\begin{array}[]{llcl}\Phi:&(x_{1},y_{1})&\mapsto&(u_{1},v_{1})\\ \Psi:&(s,h)&\mapsto&(\tilde{s},h)\end{array}.

Denoting by Ψ(s,h)=(ψ(s,h),h)\Psi(s,h)=(\psi(s,h),h), we have

(4.8) ψsdsdt=ds~dt=s~=ψ(s,h),\frac{\partial\psi}{\partial s}\frac{ds}{dt}=\frac{d\tilde{s}}{dt}=\tilde{s}=\psi(s,h),

that is,

lnψs=1ds/dt.\frac{\partial\ln\psi}{\partial s}=\frac{1}{ds/dt}.

Recall the system (3.3) is the following

dsdt=λs𝗄+o(s)τ(s),\frac{ds}{dt}=\lambda s^{\mathsf{k}}+o(s)\triangleq\tau(s),

and the period 11-form dt=ds/τ(s)dt=ds/\tau(s) can not have a pole at PiP^{i} with zero residue, i.e. in the Laurent series of 1/τ(s)1/\tau(s), the coefficient of 1/s1/s is a nonzero number c0c_{0}. Thus, ψ(s,h)\psi(s,h) can be expressed in ss as follows:

ψ(s,h)=sec0+τ1(s)+τ2(s),\psi(s,h)=se^{c_{0}+\tau_{1}(s)+\tau_{2}(s)},

where τ1(s)=j1c1jsj\tau_{1}(s)=\sum_{j\geq 1}c_{1j}s^{j}, τ2(s)=𝗄+1j1c2jsj\tau_{2}(s)=\sum_{-\mathsf{k}+1\leq j\leq-1}c_{2j}s^{j}, c2(𝗄+1)=1/λc_{2(-\mathsf{k}+1)}=1/\lambda, and equation (4.8) becomes

ec0+τ1(s)+τ2(s)(1+τ1+τ2)(λs𝗄+o(s))=sec0+τ1(s)+τ2(s),e^{c_{0}+\tau_{1}(s)+\tau_{2}(s)}(1+\tau_{1}^{\prime}+\tau_{2}^{\prime})(\lambda s^{\mathsf{k}}+o(s))=se^{c_{0}+\tau_{1}(s)+\tau_{2}(s)},

which implies that τ2=0\tau_{2}=0, 𝗄=1\mathsf{k}=1. Furthermore, ψ(s,h)\psi(s,h) can be analytically extended to a sufficiently small disc encircling (0,h)(0,h). Clearly the orbits of VV_{\ast} are all closed, so are the orbits of vector fields on (s~,h)(\tilde{s},h) and (s,h)(s,h) planes induced by VV and VV_{\ast} respectively. Besides, due to that the Puiseux parameterization R\mathrm{R} is a finitely many cover mapping near PiP^{i}, the orbits of VhV_{h} are also closed. ∎

Remark 4.4.

In the proof of the above lemma, the conclusion, that the function ψ(s,h)\psi(s,h) can be analytically extended to the origin of (s,h)(s,h) plane, does not mean the transformation Φ\Phi can be also analytically extended to PiP^{i}, one of the reasons is the inverse function of Puiseux parameterization R\mathrm{R} is usually multi-valued near PiP^{i}.

This lemma tells us the period 11-form ω\omega of system (1.5) has at least one pole at a point PiP^{i} at infinity on a branch of Lh¯\overline{L_{h}} near PiP^{i}. The following lemma will show that the multiplicity of such a point PiP^{i} can not be too low, i.e., we have

Lemma 4.5.

If ω\omega has a pole at a point PiP^{i} at infinity on a branch of Lh¯\overline{L_{h}} near PiP^{i}, then the multiplicity nin_{i} of PiP^{i} satisfies ni(n+1)/2n_{i}\geq(n+1)/2.

Proof.

We still assume the projective coordinate of PiP^{i} is [1:0:0][1:0:0]. Let {(ki,li),i=0,,r}\{(k_{i},l_{i}),\ i=0,\cdots,r\} be the vertex set of the Newton polygon of Hh(X,Y)H_{h}^{\ast}(X,Y) near the origin, where l0l1lr=0l_{0}\geq l_{1}\geq\cdots\geq l_{r}=0, 0=k0k1kr0=k_{0}\leq k_{1}\leq\cdots\leq k_{r}. Denoting by Ni=min{𝗉ki+𝗊li,i=0,,r},N_{i}=\min\{\mathsf{p}k_{i}+\mathsf{q}l_{i},\ i=0,\cdots,r\}, the minimum of 𝗉ki+𝗊li\mathsf{p}k_{i}+\mathsf{q}l_{i}, there exists a straight line on (k,l)(k,l)-plane

:𝗉k+𝗊l=Ni\mathcal{L}:\ \ \mathsf{p}k+\mathsf{q}l=N_{i}

passing all the points contained in {(ki,li):𝗉ki+𝗊li=Ni}\{(k_{i},l_{i}):\ \mathsf{p}k_{i}+\mathsf{q}l_{i}=N_{i}\}. We define the Newton principal polynomial gNi(X,Y)g_{N_{i}}(X,Y) by the following

gNi(X,Y)=(k,l)bk,lXkYl,g_{N_{i}}(X,Y)=\sum_{(k,l)\in\mathcal{L}}b_{k,l}X^{k}Y^{l},

where bk,lb_{k,l} is the coefficient of term XkYlX^{k}Y^{l} of Hh(X,Y)H_{h}^{\ast}(X,Y).

Taking the Puiseux parameterization x=s𝗉,y=s𝗊𝗉ρ(s)x=s^{-\mathsf{p}},\ y=s^{\mathsf{q}-\mathsf{p}}\rho(s) into ω\omega, we get that ω\omega has a pole at PiP^{i} if and only if

(4.9) Ord(s𝗉1Hy(s𝗉,s𝗊𝗉ρ(s)))1,\displaystyle Ord\left(\frac{s^{-\mathsf{p}-1}}{\frac{\partial H}{\partial y}(s^{-\mathsf{p}},s^{\mathsf{q}-\mathsf{p}}\rho(s))}\right)\leq-1,

where Ord()Ord(\cdot) represents the lowest degree of a Laurent series.

By comparing the coefficients of terms {si}\{s^{i}\} in both sides of equation

(4.10) Hh(s𝗉,s𝗊ρ(s,h))=0.\displaystyle H_{h}^{\ast}(s^{\mathsf{p}},s^{\mathsf{q}}\rho(s,h))=0.

one can easily get c0c_{0} is a root of gNi(1,Y)=0g_{N_{i}}(1,Y)=0, so we can assume gNi(1,Y)=(Yc0)kNig(Y),g_{N_{i}}(1,Y)=(Y-c_{0})^{k_{N_{i}}}g(Y), where g(c0)0g(c_{0})\neq 0. Letting cm(h)c_{m}(h) be the first coefficient depending on hh in ρ(s,h)=s𝗊(c0+i=1ci(h)si)\rho(s,h)=s^{\mathsf{q}}(c_{0}+\sum_{i=1}^{\infty}c_{i}(h)s^{i}) and taking the derivative on hh in both sides of equation (4.10), we have

(4.11) s𝗊(imci(h)si)(sn𝗉Hy(s𝗉,s𝗊𝗉ρ(s)))=s(n+1)𝗉,\displaystyle s^{\mathsf{q}}\left(\sum_{i\geq m}c^{\prime}_{i}(h)s^{i}\right)\left(s^{n\mathsf{p}}\frac{\partial H}{\partial y}(s^{-\mathsf{p}},s^{\mathsf{q}-\mathsf{p}}\rho(s))\right)=s^{(n+1)\mathsf{p}},

Comparing the coefficients of terms {si}\{s^{i}\} on both sides of the above equation, we can get the following estimations:

  • 0<m2𝗉𝗊0<m\leq 2\mathsf{p}-\mathsf{q}, i.e., 𝗉𝗊/2\mathsf{p}\geq\mathsf{q}/2, by inequality (4.9);

  • Ni+mkNi(n+1)𝗉N_{i}+mk_{N_{i}}\geq(n+1)\mathsf{p}, this is because the lowest degree of the left side of the above equation is not more than 𝗊+m+(Ni𝗊)+(kNi1)m=Ni+mkNi\mathsf{q}+m+(N_{i}-\mathsf{q})+(k_{N_{i}}-1)m=N_{i}+mk_{N_{i}}.

In addition, from the convexity of the Newton polygon(see Figure 3 below), inequalities kNinik_{N_{i}}\leq n_{i} and Nini𝗊N_{i}\leq n_{i}\mathsf{q} are both obvious. Finally, combining these inequalities we have ni(n+1)/2.n_{i}\geq(n+1)/2.

Refer to caption
Figure 3. Newton polygon of Hh(X,Y)H_{h}^{\ast}(X,Y)

In general, there may vanishing cycles associated to a singularity in Σh\Sigma_{h} on Lh¯\overline{L_{h}} when h0h\rightarrow 0, which may be non-trivial cycles in 1(L¯h,)\mathcal{H}_{1}(\overline{L}_{h},\mathbb{Z}). However, for Hamiltonian systems with an isochronous center, we have the following lemma.

Lemma 4.6.

For system (1.5), if the origin is an isochronous center, then there does not exist a vanishing cycle γh\gamma^{\prime}_{h} associated to a singular point PiP^{i} at infinity such that limh0γhω0\lim_{h\rightarrow 0}\oint_{\gamma^{\prime}_{h}}\omega\neq 0.

Proof.

Suppose otherwise, i.e., suppose that such a vanishing cycle γh\gamma^{\prime}_{h} exists and is associated to a point PiP^{i} at infinity with the projective coordinate [1:0:0][1:0:0].

Denote by (L¯h)\sharp(\overline{L}_{h}) the number(counting multiplicity) of branches determined by the parts of the segments with slope 1/2\leq-1/2 in the Newton polygons of HhH^{\ast}_{h} at PiP^{i}. On one hand, when γh0\gamma^{\prime}_{h}\rightarrow 0 as h0h\rightarrow 0, it yields at least one more branch on L0¯\overline{L_{0}} than Lh¯\overline{L_{h}}, where the period 11-form ω\omega has a pole at PiP^{i} with a nonzero residue. So this branch is determined by a segment with slope 1/2\leq-1/2 in the Newton polygons of H0H^{\ast}_{0}, by Lemma 4.5 and its proof. This implies (L¯0)>(L¯h)\sharp(\overline{L}_{0})>\sharp(\overline{L}_{h}).

On the other hand, the isochronous center is of Morse type, and the part of H(x,y)H(x,y) with degree 22 has the form a20x2+a11xy+a02y2a_{20}x^{2}+a_{11}xy+a_{02}y^{2} such that a20a_{20} and a11a_{11} can not be zero simultaneously. So we have

(4.14) Hh(X,Y)=hXn+1+a20Xn1+a11Xn1Y+a02Xn1Y2+k=0n2l=0ka(n+1kl)lXkYl.\displaystyle\begin{array}[]{rcl}H^{\ast}_{h}(X,Y)&=&-hX^{n+1}+a_{20}X^{n-1}+a_{11}X^{n-1}Y+a_{02}X^{n-1}Y^{2}\\ &&+\sum_{k=0}^{n-2}\sum_{l=0}^{k}a_{(n+1-k-l)l}X^{k}Y^{l}.\end{array}

Then there are the following three possible cases, and each of them yields a contradiction.

  • If the Newton polygon for h0h\neq 0 does not contain two points (k1,l1)=(n1,1)(k_{1},l_{1})=(n-1,1) and (k2,l2)=(n+1,0)(k_{2},l_{2})=(n+1,0) simultaneously, then H0H^{\ast}_{0} and HhH^{\ast}_{h} have the same parts of the segments with slope 1/2\leq-1/2, which give the same number (L¯0)=(L¯h)\sharp(\overline{L}_{0})=\sharp(\overline{L}_{h}).

  • If the Newton polygon for h0h\neq 0 contains two points (k1,l1)=(n1,1)(k_{1},l_{1})=(n-1,1) and (k2,l2)=(n+1,0)(k_{2},l_{2})=(n+1,0) simultaneously, and the line segment 𝗅𝗌𝟣\mathsf{ls_{1}} with slope 1/21/2 contains only two points (n1,1)(n-1,1) and (n+1,0)(n+1,0), then then H0H^{\ast}_{0} and HhH^{\ast}_{h} have the same shapes of the part of the segments with slope <1/2<-1/2, and the branch determined by 𝗅𝗌𝟣\mathsf{ls_{1}} on Lh¯\overline{L_{h}} has a Puiseux parameterization

    X=s,Y=s2(h+i>1di(h)si),X=s,\ Y=s^{2}\left(h+\sum_{i>1}d_{i}(h)s^{i}\right),

    which tends to the branch Y=0Y=0 on L0¯\overline{L_{0}}. Consequently, we still have (L¯0)=(L¯h)\sharp(\overline{L}_{0})=\sharp(\overline{L}_{h}).

  • If the Newton polygon for h0h\neq 0 contains two points (k1,l1)=(n1,1)(k_{1},l_{1})=(n-1,1) and (k2,l2)=(n+1,0)(k_{2},l_{2})=(n+1,0) simultaneously, but the line segment 𝗅𝗌𝟤\mathsf{ls_{2}} with slope 1/2-1/2 contains not only these two points, then we will show that in this case the system (1.5) can not be linearizable at the point (x,y)=(0,0)(x,y)=(0,0).

    Let k0k_{0} be the maximum value such that (k0,l)𝗅𝗌𝟤(k_{0},l)\in\mathsf{ls_{2}} and k0<n1k_{0}<n-1. Then in this case H(x,y)H(x,y) has the form

    (4.17) H(x,y)=a11xy+a22y2+ai0i0xi0yi0+i<j,i+j2i0aijxiyj+ij,i+j>2i0aijxiyj,\displaystyle\begin{array}[]{rcl}H(x,y)&=&a_{11}xy+a_{22}y^{2}+a_{i_{0}i_{0}}x^{i_{0}}y^{i_{0}}+\sum_{i<j,i+j\leq 2i_{0}}a_{ij}x^{i}y^{j}\\ &&+\sum_{i\leq j,i+j>2i_{0}}a_{ij}x^{i}y^{j},\\ \end{array}

    where ai0i00,i0=(n+1k0)/2a_{i_{0}i_{0}}\neq 0,\ i_{0}=(n+1-k_{0})/2. In fact, i0ai0i0i_{0}a_{i_{0}i_{0}} is nothing other than the first nonzero linearization constant. This is because, from the results in [2], the (2i02)(2i_{0}-2)-jet of system (1.5) has admissible nonlinearities and can be linearized by a transformation of the form

    (4.20) U=x+ijuijxiyjV=y+i<j1vijxiyj.\displaystyle\begin{array}[]{rcl}U&=&x+\sum_{i\leq j}u_{ij}x^{i}y^{j}\\ V&=&y+\sum_{i<j-1}v_{ij}x^{i}y^{j}\end{array}.

    However, this transformation can not change the resonant terms i0ai0i0xi0yi01i_{0}a_{i_{0}i_{0}}x^{i_{0}}y^{i_{0}-1} in dx/dtdx/dt and i0ai0i0xi01yi0-i_{0}a_{i_{0}i_{0}}x^{i_{0}-1}y^{i_{0}} in dy/dtdy/dt.

5. The proof of main theorems

Now we can prove our main theorems.

Proof of Theorem 1.2.

By Lemma 4.2, without loss of generality, we assume γh\gamma_{h} has been divided into kk parts {γhj,j=1,2,,k}\{\gamma^{j}_{h},\ j=1,2,...,k\}, such that for any point pγhjp\in\gamma_{h}^{j}, iφ(p,t){\rm i}\varphi(p,t) goes to the same point PjP^{j} at infinity on the same branch when t+t\rightarrow+\infty. Here γhj\gamma^{j}_{h} may not be a continuous arc of γh\gamma_{h} but a union of finitely many continuous arcs. By Lemma 4.3, we can choose a closed orbit δhj\delta_{h}^{j} of VhV_{h} encircling and sufficiently close to PjP^{j} on this branch.

We shall prove that, for each 1jk1\leq j\leq k, there exists a moment tjt_{j} such that iφ(γhj,tj)¯=δhj\overline{{\rm i}\varphi(\gamma_{h}^{j},t_{j})}=\delta_{h}^{j}. Then γh\gamma_{h} is homologous to the summation of those cycles j=1kδhj\sum_{j=1}^{k}\delta_{h}^{j}, and the theorem holds since that each δhj\delta_{h}^{j} represents a zero homology cycle in 1(Lh¯,)\mathcal{H}_{1}(\overline{L_{h}},\mathbb{Z}).

Suppose otherwise, i.e. suppose that there exists a number jj and a branch of LhL_{h} near a point PjP^{j} such that the loop δhj\delta_{h}^{j} contains at least two continuous arcs Arh1Ar_{h1} and Arh2Ar_{h2} satisfying that Ar1iφ(γhj,tj)Ar_{1}\subseteq{\rm i}\varphi(\gamma_{h}^{j},t_{j}) but Arh2iφ(γhj,tj)=Ar_{h2}\cap{\rm i}\varphi(\gamma_{h}^{j},t_{j})=\emptyset.

By the continuation technique, we can extend the transformation Φ\Phi to a domain 𝒟\mathcal{D} containing Arh1,Arh2Ar_{h1},\ Ar_{h2} and iφ(Arh2,t){\rm i}\varphi(Ar_{h2},t^{\prime}) for any sufficiently small |h|\left|h\right|, where t(t0,0]t^{\prime}\in(-t_{0},0] and t0t_{0} is a real number such that t0>tjt_{0}>t_{j}. To avoid that iφ(Arh2,t){\rm i}\varphi(Ar_{h2},t^{\prime}) meets a point at infinity, Arh2Ar_{h2} can be shortened properly. Consequently, for any point q′′Arh2q^{\prime\prime}\in Ar_{h2}, letting q=iφ(q′′,tj)q^{\prime}={\mathrm{i}}\varphi(q^{\prime\prime},-t_{j}), there exists a point qγhq\in\gamma_{h} such that Φ(q)=Φ(q)\Phi(q^{\prime})=\Phi(q) but qγhq^{\prime}\not\in\gamma_{h} (see Figure 4).

Refer to caption
Figure 4. Proof of Theorem 1.2

Note that iφ(Arh2,tj){\mathrm{i}}\varphi(Ar_{h2},-t_{j}) must be a part of a closed orbit γh\gamma^{\prime}_{h} of vector field VV. If not so, then for almost every tt sufficiently close to tj-t_{j}, the orbit of VV containing iφ(Arh2,t){\mathrm{i}}\varphi(Ar_{h2},t) tends to a point PP at infinity, so the orbits of iV{\rm i}V near PP either also tend to PP or are closed near PP, which implies the number 𝗄1\mathsf{k}\geq 1 for this branch of PP. However, Φ(iφ(Arh2,t))\Phi({\mathrm{i}}\varphi(Ar_{h2},t)) is a part of a closed orbit of VV_{\ast}, then by Lemma 4.3 and its proof, the orbit of VV must also be closed near PP, this is a contradiction.

Below we shall show that γh\gamma^{\prime}_{h} must be a vanishing cycle associated to the origin or a singularity at infinity.

Under the assumption of the theorem, L0L_{0} has the same structures at PjP^{j} to LhL_{h}, by Lemma 4.6 and its proof. So the above 𝒟\mathcal{D} can also extended to L0L_{0}: given two closed orbits γ0\gamma_{0} and δ0\delta_{0} of V0V_{0} sufficiently close to OO and PjP^{j} on the branch of L0L_{0} that is the limit of branch containing Arh2Ar_{h2} on LhL_{h} when h0h\rightarrow 0, then iφ(γ0,t)¯δ0\overline{{\rm i}\varphi(\gamma_{0},t)}\not=\delta_{0} for any tt. Thus 𝒟\mathcal{D} can contain a trajectory Ar02δ0Ar_{02}\subset\delta_{0} but iφ(γ0,t)¯\not\subset\overline{{\rm i}\varphi(\gamma_{0},t)} and iφ(Ar02,t){\rm i}\varphi(Ar_{02},t^{\prime}) for some a t(,0]t^{\prime}\in(-\infty,0].

Noticing that L0L_{0} has only a single finite singular point, iφ(Ar02,t){\rm i}\varphi(Ar_{02},t^{\prime}) can be well defined for tt^{\prime}\rightarrow-\infty and its limit is either the origin or a point at infinity, by Lemma 4.2 and its proof. Besides, vector field VgV_{g} coincides with iV-{\rm i}V on L0L_{0}, so iφ(Arh2,tj){\mathrm{i}}\varphi(Ar_{h2},-t_{j}) will tend to limtiφ(Ar02,t)\lim_{t^{\prime}\rightarrow-\infty}{\rm i}\varphi(Ar_{02},t^{\prime}) along the trajectory of VgV_{g}. This means γh\gamma^{\prime}_{h} is a vanishing cycle of a singularity at infinity or the origin.

However, by Lemma 4.6 and the assumption on the period 11-form, the former case is impossible. As for the latter case, recalling that Φ(q)=Φ(q)\Phi(q^{\prime})=\Phi(q) but iφ(Arh2,tj)γh={\rm i}\varphi(Ar_{h2},-t_{j})\bigcap\gamma_{h}=\emptyset and Φ\Phi is a homeomorphism near the origin, it is also impossible.

Proof of Corollary 1.3.

If the linearization change Φ\Phi is well defined on the whole plane 2\mathbb{C}^{2}, for instance, polynomial map ΦP\Phi_{P} appearing in the Jacobian conjecture, then it maps a small disc punctured by a pole of the period 11-form ω\omega to a small disc(topologically) punctured by a pole of 11-form du/v-du/v on LhL_{h} for any hh. This means that ω\omega dose not have poles with zero residue at infinity.

Besides, for polynomial map ΦP\Phi_{P}, due to

det(fxfygxgy)0,\det\left(\begin{array}[]{cc}\frac{\partial f}{\partial x}&\frac{\partial f}{\partial y}\\ \frac{\partial g}{\partial x}&\frac{\partial g}{\partial y}\end{array}\right)\not=0,

the singularities on critical level curve H(x,y)=(f2+g2)/2=0H(x,y)=(f^{2}+g^{2})/2=0 are just intersections of two algebraic curves f=0f=0 and g=0g=0. Thus, by Theorem 1.2 and Proposition 6.1 in [7], the corollary holds. ∎

Proof of Theorem 1.4.

Noticing that ω\omega has a pole at infinity is equivalent to say 𝗄1\mathsf{k}\geq 1 for system (3.3), this theorem is a direct conclusion of Lemma 4.2 and Lemma 4.5. ∎

6. Non-isochronicity of real Hamiltonian systems of even degree nn

In the last section we focus on the relation between the Gavrilov’s question and Jarque-Villadelprat conjecture. It is worthy mentioned that the latter is not true in the complex setting, some counterexamples can be found in Gavrilov’s paper [7]. Firstly we shall prove Theorem 1.6.

Proof of Theorem 1.6.

If H(x,y)H(x,y) is a real polynomial of odd degree n+1n+1, then the real algebraic curve LhL_{h} has at least two connected components on 2\mathbb{R}^{2}, one of them is just the closed orbit γh\gamma_{h} near the center which can represent the corresponding vanishing cycle, and another one, denoted by γh\gamma^{\prime}_{h}, tends to a point at infinity. The real systems can be embedded in 2={(x,y)=(x1+ix2,y1+iy2)}{(x1,x2,y1,y2)}=4\mathbb{C}^{2}=\{(x,y)=(x_{1}+{\rm i}x_{2},y_{1}+{\rm i}y_{2})\}\cong\{(x_{1},x_{2},y_{1},y_{2})\}=\mathbb{R}^{4}. Then the real plane 2\mathbb{R}^{2} is a subset defined by x2=y2=0x_{2}=y_{2}=0, and the closed orbit γh\gamma_{h} on 2\mathbb{R}^{2} can be represented by H(x1,y1)=hH(x_{1},y_{1})=h.

If

Φ:22,(x1,y1)(u1,v1)=(ϕ1(x1,y1),ϕ2(x1,y1))\Phi_{\mathbb{R}}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2},\ \ (x_{1},y_{1})\mapsto(u_{1},v_{1})=(\phi_{1}(x_{1},y_{1}),\phi_{2}(x_{1},y_{1}))

is the transformation linearizing real isochronous center, then the following map

Φ:22,(x,y)(u,v)=(ϕ1(x,y),ϕ2(x,y))\Phi:\mathbb{C}^{2}\rightarrow\mathbb{C}^{2},\ \ (x,y)\mapsto(u,v)=(\phi_{1}(x,y),\phi_{2}(x,y))

can linearized the complex isochronous center of system (1.5). Denoting by

Π:22,(x,y)(x¯,y¯)\Pi:\mathbb{C}^{2}\rightarrow\mathbb{C}^{2},\ \ (x,y)\mapsto(\overline{x},\overline{y})

the conjugate operation on 2\mathbb{C}^{2}, we have ΦΠ=ΠΦ\Phi\circ\Pi=\Pi\circ\Phi, since

(u¯,v¯)=(ϕ1(x,y)¯,ϕ2(x,y)¯)=(ϕ1(x¯,y¯),ϕ2(x¯,y¯)).(\overline{u},\overline{v})=(\overline{\phi_{1}(x,y)},\overline{\phi_{2}(x,y)})=(\phi_{1}(\overline{x},\overline{y}),\phi_{2}(\overline{x},\overline{y})).

If γh\gamma_{h} is a trivial cycle on the closure Lh¯\overline{L_{h}} of the generic complex curve LhL_{h} defined by H(x,y)=hH(x,y)=h, then LhL_{h} is divided into two path-connected open components A1A_{1} and A2A_{2} such that A1A2=A_{1}\cap A_{2}=\emptyset and their common boundary is γh\gamma_{h}. Without loss of generality, we assume γhA1\gamma^{\prime}_{h}\cap A_{1}\neq\emptyset and can construct a smooth curve lhLhl_{h}\subset L_{h} connecting two points pγhp\in\gamma_{h} and pγhA1p^{\prime}\in\gamma^{\prime}_{h}\cap A_{1}, such that lhl_{h} intersects γh\gamma_{h} transversally at only one point pp.

In the domain 𝒟\mathcal{D} where Φ\Phi is well defined, we have Π(A1𝒟)A2\Pi(A_{1}\cap\mathcal{D})\subset A_{2}, because Φ\Phi is a homeomorphism so that Φ(A1𝒟)Φ(A2𝒟)=\Phi(A_{1}\cap\mathcal{D})\cap\Phi(A_{2}\cap\mathcal{D})=\emptyset and Π(Φ(A1𝒟))Φ(A2𝒟)\Pi(\Phi(A_{1}\cap\mathcal{D}))\subset\Phi(A_{2}\cap\mathcal{D}). Therefore the complex conjugate lh¯\overline{l_{h}} of lhl_{h} belongs to A2A_{2} in 𝒟\mathcal{D}. Finally lhlh¯{p,p}l_{h}\cup\overline{l_{h}}\cup\{p,p^{\prime}\} forms a closed curve intersecting γh\gamma_{h} at only one point pp with intersection number 11 on LhL_{h}(see Figure 5). This means that γh\gamma_{h} can not be trivial on Lh¯\overline{L_{h}}, which leads a contradiction.

Refer to caption
Figure 5. lhl_{h} and its complex conjugate

By Theorem 1.6, we observe an interesting relation between Gavrilov’s question and Jarque-Villadelprat conjecture, that is, if the latter conjecture is not true, then such real systems possessing isochronous centers provide a negative answer to Gavrilov’s question.

In the end, as applications of Theorem 1.2, we present a conclusion which verifies the Jarque-Villadelprat conjecture for a large class of real systems. Note that in the real setting, an isochronous center must be a non-degenerated singularity, i.e., it must be of Morse type, so we have

Corollary 6.1.

For a real polynomial Hamiltonian system (1.5) of even degree, if each (complex) critical level curve having a center contains only a single singularity, and the period 11-form has no pole at infinity with zero residue on any level curve, then it does not admit any isochronous center at all.

Acknowledgements

This work is supported by NSFC 11701217 and NSF 2017A030310181 of Guangdong Province(China).

References

  • [1] B. Arcet, J. Gine and V.G. Romanovski, Linearizability of planar polynomial Hamiltonian systems, Nonlinear Analysis: Real World Applications 63 (2022), 103422, 19.
  • [2] J. Basto-Goncalves, Linearization of resonant vector fields, Trans. Amer. Math. Soc. 362 (12) (2010), 6457-6476.
  • [3] C. Camacho, A. Lins Neto, and P. Sad, Topological invariants and equidesingularization for holomorphic systems, J. Differential Geom. 20 (1984), no. 1, 143-174.
  • [4] A. Cima, F. Mañosas, J. Villadelprat, Isochronicity for several classes of Hamiltonian systems, J. Differential Equations 157, 373-413 (1999).
  • [5] J. Cresson, J. Palafox, Isochronous centers of polynomial Hamiltonian systems and a conjecture of Jarque and Villadelprat, J. Differential Equations 266, 5713-5747 (2019).
  • [6] G. Fischer, Plane algebraic curves, Translated from the 1994 German original by Leslie Kay. Student Mathematical Library 15. American Mathematical Society, Providence, RI, 2001.
  • [7] L. Gavrilov, Isochronicity of plane polynomial Hamiltonian systems, Nonlinearity 10 (1997), 433-448.
  • [8] Y. Ilyashenko, S. Yakovenko, Lectures on analytic differential equations, Graduate Studies in Mathematics 86, American Mathematical Society, Providence, RI, 2008.
  • [9] X. Jarque and J. Villadelprat, Nonexistence of isochronous centers in planar polynomial Hamiltonian systems of degree four. J. Differential Equations 180, 334-373 (2002).
  • [10] F. Kirwan, Complex Algebraic Curves, London Mathematical Society, Student Text 23, Cambridge University Press, Cambridge, 1992.
  • [11] Llibre, J., Romanovski, V.G.: Isochronicity and linearizability of planar polynomial Hamiltonian systems. J. Differential Equations 259, 1649-1662 (2015).
  • [12] F. Mañosas and J. Villadelprat, Area-preserving normalizations for centers of planar Hamiltonian J. Differential Equations 179, 625-646 (2002).