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

Global existence and singularity of the Hill’s type lunar problem with strong potential

Abstract.

We characterize the fate of the solutions of Hill’s type lunar problem using the ideas of ground states from PDE. In particular, the relative equilibrium will be defined as the ground state, which satisfies some crucial energetic variational properties in our analysis. We study the dynamics of the solutions below, at, and (slightly) above the ground state energy threshold.

Yanxia Deng   111School of Mathematics (Zhuhai), Sun Yat-sen University, Zhuhai, Guangdong, China  dengyx53@mail.sysu.edu.cn, Slim Ibrahim222Department of Mathematics and Statistics, University of Victoria, Victoria, BC, Canada  ibrahims@uvic.ca

1. Introduction.

The three-body problem is a prototypical case in celestial mechanics. The system Sun-Earth-Moon can be considered as a typical example of the three-body problem. Using heuristic arguments about the relative size of various physical constants, Hill was able to give the equations for the motion of the moon as an approximation from the general three-body problem. The Hill’s lunar problem can be derived from the general three-body problem using symplectic scaling method [9]. We give a derivation of the main problem for homogeneous potential in the last section of the paper. A popular description of the Hill’s equations is to consider the motion of an infinitesimal body (the moon) which is attracted to a body (the earth) fixed at the origin. The infinitesimal body moves in a rotating coordinate system which rotates so that the positive xx-axis points towards an infinite body (the sun) which is infinitely far way. The ratio of the two infinite quantities is taken so that the gravitational attraction of the sun on the moon is finite.

In particular, if the position of the moon is given by (x,y)(x,y), the planar Hill’s equation with homogenous gravitational potential is given by

(1) {x¨2y˙=Vxy¨+2x˙=Vy,\begin{cases}\ddot{x}-2\dot{y}&=-V_{x}\\ \ddot{y}+2\dot{x}&=-V_{y},\end{cases}

where

(2) V(x,y)=α+22x2α+2rα,r=x2+y2,α>0V(x,y)=-\frac{\alpha+2}{2}x^{2}-\frac{\alpha+2}{r^{\alpha}},\quad r=\sqrt{x^{2}+y^{2}},\quad\alpha>0

is known as the effective potential. When α=1\alpha=1, this is the Newtonian Hill’s Lunar Problem; when α2\alpha\geq 2, we shall call it the Hill’s type lunar problem with strong potential.

This vector field is well-defined everywhere except at the origin (0,0)(0,0), which is the position of the earth. By the existence and uniqueness theorem of ODE, given q(0)=(x(0),y(0))(0,0)q(0)=(x(0),y(0))\neq(0,0) and q˙(0)2\dot{q}(0)\in\mathbb{R}^{2}, there exists a unique solution q(t)q(t) defined on the interval [0,Tmax)[0,T_{\mathrm{max}}), where TmaxT_{\mathrm{max}} is the maximal interval of existence.

Definition 1.

If Tmax<T_{\mathrm{max}}<\infty, then the solution is said to experience a singularity at TmaxT_{\mathrm{max}}; otherwise, we say the solution exists globally.

Since the ODE is locally Lipschitz in r0r\neq 0, blow-up is possible only by approaching the unique singularity, namely the collision. In particular, for the Hill’s equation, if Tmax<T_{\mathrm{max}}<\infty, then

(3) limtTmax(x(t),y(t))=(0,0),\begin{split}\lim\limits_{t\to T_{\mathrm{max}}}(x(t),y(t))=(0,0),\end{split}

that is, the singularity of the Hill’s equation is due to finite time collision at the origin.

The Hill’s equation admits the famous Jacobi integral which we shall refer to as the energy,

(4) E(x,y,x˙,y˙):=12(x˙2+y˙2)+V(x,y).E(x,y,\dot{x},\dot{y}):=\frac{1}{2}(\dot{x}^{2}+\dot{y}^{2})+V(x,y).

The effective potential V(x,y)V(x,y) has exactly two critical points L1:=(α1α+2,0)L_{1}:=(-\alpha^{\frac{1}{\alpha+2}},0) and L2:=(α1α+2,0)L_{2}:=(\alpha^{\frac{1}{\alpha+2}},0), which are known as the Lagrange points. The projection of the four-dimensional phase space onto the configuration (x,y)(x,y) space is called the Hill’s regions.

(5) c:={(x,y)|E(x,y,x˙,y˙)=c}={(x,y)|V(x,y)c}.\begin{split}\mathcal{H}_{c}:=\{(x,y)|E(x,y,\dot{x},\dot{y})=c\}=\{(x,y)|V(x,y)\leq c\}.\end{split}

The boundaries of the Hill’s regions are called zero velocity curves because they are the locus in the configuration space where the kinetic energy vanishes. A contour plot for V(x,y)V(x,y) in the (x,y)(x,y)-plane could be seen in Figure 1.

Refer to caption
Figure 1. The contour plot of V(x,y)V(x,y) with α=1\alpha=1. V(x,y)V(x,y) has two critical points L1:=(α1α+2,0)L_{1}:=(-\alpha^{\frac{1}{\alpha+2}},0) and L2:=(α1α+2,0)L_{2}:=(\alpha^{\frac{1}{\alpha+2}},0).

The structure of the Hill’s regions depends on the value of the energy. There are four distinct cases regarding the shape of the Hill’s regions:

  • (i)

    c<Ec<E^{*}: both necks are closed, so orbits inside will remain bounded in the configuration space or collide with the origin.

  • (ii)

    c=Ec=E^{*}: the threshold case.

  • (iii)

    E<c<0E^{*}<c<0: both necks are open, thus allowing orbits to enter the exterior region and escape from the system.

  • (iv)

    c0c\geq 0: motions over the entire configuration (x,y)(x,y) space is possible.

In Figure 2 we present the structure of the first and third possible Hill’s region for α=1\alpha=1; all the other α>0\alpha>0 have the same structure with varied values of L1,L2L_{1},L_{2}.

Refer to caption
Refer to caption
Figure 2. Hill’s regions c\mathcal{H}_{c} when α=1\alpha=1, E=4.5E^{*}=-4.5. The white domains correspond to the Hill’s regions, gray shaded domains indicate the energy forbidden regions. Left is below the ground state energy with c=4.6c=-4.6; right is above the ground state energy with c=4.4c=-4.4.

Consequently, ±Q:=(±α1α+2,0,0,0)\pm Q:=(\pm\alpha^{\frac{1}{\alpha+2}},0,0,0) are the only equilibria of (1). We will define ±Q\pm Q to be the ground state. In section 2, we will give a detailed analysis about the energetic variational properties satisfied by ±Q\pm Q as well as some insights on why would we call them the ground states. In particular, we will first define the ground state energy EE^{*}, which will be the minimum of the energy EE under the constraint W=0W=0, where

(6) W:=xVxyVy=(α+2)(x2αrα).\begin{split}W:=-xV_{x}-yV_{y}=(\alpha+2)(x^{2}-\frac{\alpha}{r^{\alpha}}).\end{split}

Note that EE^{*} coincides with the energy of the equilibria. See Figure 3 for the graph of V=EV=E^{*} and W=0W=0 in the (x,y)(x,y) plane. In particular, W<0W<0 corresponds to the inner bounded region, and W>0W>0 corresponds to the outer region. The rest of the paper is then devoted to study the dynamics of the solution of (1) with energy below, at, and (slightly) above the ground state energy threshold. Our main results are in the following theorems.

Theorem 1 (Dichotomy below the ground state).

For the Hill’s lunar problem, consider the sets:

(7) 𝒲+={Γ=(x,y,x˙,y˙)|E(Γ)<E,W(Γ)>0}𝒲={Γ=(x,y,x˙,y˙)|E(Γ)<E,W(Γ)0}\begin{split}\mathcal{W}_{+}&=\{\Gamma=(x,y,\dot{x},\dot{y})|E(\Gamma)<E^{*},W(\Gamma)>0\}\\ \mathcal{W}_{-}&=\{\Gamma=(x,y,\dot{x},\dot{y})|E(\Gamma)<E^{*},W(\Gamma)\leq 0\}\end{split}

then 𝒲+\mathcal{W}_{+} and 𝒲\mathcal{W}_{-} are invariant. Solutions in 𝒲+\mathcal{W}_{+} exist globally and solutions in 𝒲\mathcal{W}_{-} are bounded globally or collide with the origin in finite time. Moreover, for α2\alpha\geq 2, solutions in 𝒲\mathcal{W}_{-} all collide with the origin in finite time.

It is interesting to notice that for the strong potential where α2\alpha\geq 2, the system exhibits significantly different behavior from the weak potential case, as can be readily seen in the Kepler problem, where α=2\alpha=2 is a bifurcation critical value (cf. [3]). Theorem 1 suggests that there are simple smooth boundaries that distinguish colliding orbits from global ones for α2\alpha\geq 2 under some energy threshold, while for α<2\alpha<2 there are no simple boundaries and indeed they seem to be fractal as suggested by the numerical results in the paper [4]. Actually, it is well-known that the Newtonian Hill’s problem has traversal homoclinic intersections, hence chaotic (cf. [12][13]). In the current paper, we will focus on the strong potential case.

Theorem 2 (At the ground state energy).

For α2\alpha\geq 2, let

(8) 𝒲~+={Γ=(x,y,x˙,y˙)|E(Γ)=E,W(Γ)>0}𝒲~={Γ=(x,y,x˙,y˙)|E(Γ)=E,W(Γ)0}\begin{split}\tilde{\mathcal{W}}_{+}&=\{\Gamma=(x,y,\dot{x},\dot{y})|E(\Gamma)=E^{*},W(\Gamma)>0\}\\ \tilde{\mathcal{W}}_{-}&=\{\Gamma=(x,y,\dot{x},\dot{y})|E(\Gamma)=E^{*},W(\Gamma)\leq 0\}\end{split}

then they are invariant. Solutions in 𝒲~\tilde{\mathcal{W}}_{-} either have a finite time collision or approach the ground states as tt\to\infty, moreover, they approach the ground states only when they are on the stable manifolds of the ground states. Solutions in 𝒲~+\tilde{\mathcal{W}}_{+} exist globally.

When the energy is above the ground state energy, both bottle necks open up, and we no longer have invariant sets based on the sign of WW. But we can still describe the motion of the solutions when they are near the ground states and the bottle necks.

Theorem 3 (Above the ground state energy).

For α2\alpha\geq 2, there exists ϵ>0\epsilon>0, so that any solution ψ(t)\psi(t) with E(ψ)<E+ϵ2E(\psi)<E^{*}+\epsilon^{2} either stays inside a 2ϵ2\epsilon ball of the ground states, or ejected out of the ball. Moreover, if at the time of the exit of the ball, the sign of WW is negative, then the solution collides with the origin in finite time.

We remark that the finite time collision conclusion in Theorem 3 does not hold for the weak potential. In particular, for the Newtonian case, there are heteroclinic intersections between the two ground states, and solutions may exit the ball with W<0W<0 and then come back to the ball. For instance, see figure 7 of [12].

On the other hand, if at the time of exit the sign of WW is positive, we conjecture that the solution will escape to infinity. This is somehow supported by the results in the Newtonian case. In [7], the authors showed that the unstable manifolds for the ground states in the outer region (i.e. W>0W>0) goes forward to infinity for the Newtonian case. Their proof relies on the increasing of the argument along the orbits for a suitable complex coordinate system, which is the main theorem of [8]. However, for the strong potential, it seems that we don’t have the result of [8]. We will explain more about this at the end of section 5. Nonetheless, we did some numerical computations to give evidence of the no-return property (hence the escape) for the positive case for α2\alpha\geq 2, see the companion paper [4]. In this paper, we will leave the “no-return” property of the positive case (i.e. W>0W>0 at the exit) as a conjecture.

Also note that in the current paper, we did not describe the dynamics of the global solutions. This part is in a subsequent work in preparation [5]. The current paper is more on the characterization of the global existence and singularity of the solutions.

The methods of the paper are motivated from PDE, in particular, nonlinear dispersive equations (e.g. Klein-Gordon, NLS) as studied by Nakanishi-Schlag [10][11] and many others. Using this method, the authors have also given some conditional characterizations about the global existence and singularity of the general N-body problem in [3]. It is important to emphasize that Nakanishi-Schlag [10] extended the results of Theorem 3 to the case where WW is positive, using the infinite-dimensional aspect of the Klein-Gordon equation. Indeed, the striking difference between the finite-dimensional and infinite-dimensional models shows up. For example, small initial data leads to global solutions for the Klein-Gordon equation (cf. Lemma 2.2 [10]), which is not true in the Hill’s problem, as clearly seen in Theorem 1 for small initial data they collide.

The paper is organized as follows. In section 2, we give the definition of the ground state energy. In section 3, 4, 5, we give the proofs of the above Theorems. In the last section (appendix) we give a derivation of the Hill’s type lunar problem from the general three-body problem. In the paper, we shall use HLP to refer to the Hill’s lunar problem (1).

2. Defining the ground state

In PDE, the ground states are the solitons with the lowest energy. In the N-body problem, the solutions that play a similar role to the solitons are the relative equilibria, i.e. equilibrium under a uniform rotating frame. For the HLP, the only relative equilibria are given by the two Lagrange points, thus we shall expect them to be the ground states. In particular, we consider the following variational problem in 4\mathbb{R}^{4}:

(9) inf{E(x,y,x˙,y˙)|W=0},\inf\{E(x,y,\dot{x},\dot{y})|W=0\},

where W=xVxyVy=(α+2)(x2αrα)W=-xV_{x}-yV_{y}=(\alpha+2)(x^{2}-\frac{\alpha}{r^{\alpha}}).

For reasons why we consider the above type of variational problem, we refer the readers to [3].

Lemma 1.

Let E:=inf{E(x,y,x˙,y˙)|W=0}E^{*}:=\inf\{E(x,y,\dot{x},\dot{y})|W=0\}, then EE^{*} is finite and it’s achieved exactly by the relative equilibira ±Q=(±α1α+2,0,0,0)\pm Q=(\pm\alpha^{\frac{1}{\alpha+2}},0,0,0).

Proof.

Since E=12(x˙2+y˙2)+V(x,y)E=\frac{1}{2}(\dot{x}^{2}+\dot{y}^{2})+V(x,y), its minimum with the constraint W=0W=0 must satisfy x˙=y˙=0\dot{x}=\dot{y}=0. Next it’s easy to see the minimum of V(x,y)V(x,y) with W(x,y)=0W(x,y)=0 cannot be -\infty, thus we use the Lagrange multiplier V=λW\nabla V=\lambda\nabla W, the only solution of which is λ=0\lambda=0, i.e. the minimum is achieved at the critical points of VV. ∎

Definition 2 (Ground state).

We call EE^{*} the ground state energy of the HLP, and the relative equilibria ±Q\pm Q the ground states. Moreover, we have

E=E(±Q)=12(α+2)2ααα+2.E^{*}=E(\pm Q)=-\frac{1}{2}(\alpha+2)^{2}\alpha^{-\frac{\alpha}{\alpha+2}}.

Now, in order to study the fate of the solutions, we investigate the behavior of the distance between (x,y)(x,y) and the singular point (0,0)(0,0). In particular, let I:=12(x2+y2)I:=\frac{1}{2}(x^{2}+y^{2}) be the moment of inertia. If (x(t),y(t))(x(t),y(t)) is a solution to (1), then

(10) d2Idt2=x˙2+y˙2+2(xy˙x˙y)xVxyVy.\frac{d^{2}I}{dt^{2}}=\dot{x}^{2}+\dot{y}^{2}+2(x\dot{y}-\dot{x}y)-xV_{x}-yV_{y}.

Let

(11) K(x,y,x˙,y˙):=x˙2+y˙2+2(xy˙x˙y)xVxyVy,K(x,y,\dot{x},\dot{y}):=\dot{x}^{2}+\dot{y}^{2}+2(x\dot{y}-\dot{x}y)-xV_{x}-yV_{y},

then the sign of KK will describe the behavior of II along a solution, hence the fate. In the rest of this section, we give some variational properties of KK in terms of the ground state energy EE^{*}, which will play an important role in the proof of our main theorems. In particular, we prove the following proposition:

Proposition 1.

For α2\alpha\geq 2,

inf{E|K0,W0}=inf{E|K=0,W=0}=E.\inf\{E|K\geq 0,W\leq 0\}=\inf\{E|K=0,W=0\}=E^{*}.

The rest of this section is dedicated to the proof of Proposition 1. First we can show that for α2\alpha\geq 2, if EE is bounded and |K||K| is small, then (x,y)(x,y) must be away from the singular point (0,0)(0,0), see Lemma 2 and Lemma 3.

Lemma 2.

For α>2\alpha>2, there exists c>0c>0 such that if E(Γ)<1E(\Gamma)<1 and |K(Γ)|<c|K(\Gamma)|<c, then rcr\geq c.

Proof.

Suppose not, then there exists a sequence Γn4\Gamma_{n}\in\mathbb{R}^{4} such that E(Γn)<1E(\Gamma_{n})<1, |K(Γn)|<1n|K(\Gamma_{n})|<\frac{1}{n} and rn<1nr_{n}<\frac{1}{n}. Notice that

(12) E(Γ)=12(x˙2+y˙2)α+22x2α+2rα,E(\Gamma)=\frac{1}{2}(\dot{x}^{2}+\dot{y}^{2})-\frac{\alpha+2}{2}x^{2}-\frac{\alpha+2}{r^{\alpha}},

thus |K(Γn)|<1n|K(\Gamma_{n})|<\frac{1}{n} and rn<1nr_{n}<\frac{1}{n} imply that x˙n2+y˙n2(α+2)αnα\dot{x}_{n}^{2}+\dot{y}_{n}^{2}\sim(\alpha+2)\alpha n^{\alpha}, thus E(Γn)(α+2)(α21)nαE(\Gamma_{n})\sim(\alpha+2)(\frac{\alpha}{2}-1)n^{\alpha}\to\infty, contradiction. ∎

Notice that the bound E<1E<1 is not essential, in fact, the lemma can be extended to E<aE<a for any positive constant aa. For our purpose of the paper, aa doesn’t have to be positive. The energy we are interested in will only be slightly larger than that of EE^{*}, when α=2\alpha=2, E=42E^{*}=-4\sqrt{2}, thus we can state a similar lemma for α=2\alpha=2 as

Lemma 3.

For α=2\alpha=2, there exists c>0c>0 such that if E(Γ)<1E(\Gamma)<-1 and |K(Γ)|<c|K(\Gamma)|<c, then rcr\geq c.

Proof.

Suppose not, then there exists a sequence Γn4\Gamma_{n}\in\mathbb{R}^{4} such that E(Γn)<1E(\Gamma_{n})<-1, |K(Γn)|<1n|K(\Gamma_{n})|<\frac{1}{n} and rn<1nr_{n}<\frac{1}{n}. Notice that

(13) E(Γ)=12(x˙2+y˙2)2x24r2,E(\Gamma)=\frac{1}{2}(\dot{x}^{2}+\dot{y}^{2})-2x^{2}-\frac{4}{r^{2}},

and

(14) K(Γ)=(x˙y)2+(x+y˙)2+3x2y28r2,K(\Gamma)=(\dot{x}-y)^{2}+(x+\dot{y})^{2}+3x^{2}-y^{2}-\frac{8}{r^{2}},

thus |K(Γn)|<1n|K(\Gamma_{n})|<\frac{1}{n} and rn<1nr_{n}<\frac{1}{n} imply that

|x˙n2+y˙n28rn2|1n,|\dot{x}_{n}^{2}+\dot{y}_{n}^{2}-\frac{8}{r_{n}^{2}}|\lesssim\frac{1}{n},

thus E(Γn)0E(\Gamma_{n})\to 0, contradiction. ∎

Remark 1.

Similarly, we can show that for α2\alpha\geq 2, there exists c>0c>0, if E(Γ)<1E(\Gamma)<-1 and K(Γ)0K(\Gamma)\geq 0, then rcr\geq c.

2.1. Lagrange equation with one constraint

In this subsection, let’s consider the Lagrange multiplier equation: E=λK\nabla E=\lambda\nabla K with K=0K=0. In particular, we have

(15) {Vx=λ(2y˙VxxVxxyVyx),(a)Vy=λ(2x˙VyxVxyyVyy),(b)x˙=λ(2x˙2y),(c)y˙=λ(2y˙+2x).(d)\begin{cases}&V_{x}=\lambda(2\dot{y}-V_{x}-xV_{xx}-yV_{yx}),\qquad(a)\\ &V_{y}=\lambda(-2\dot{x}-V_{y}-xV_{xy}-yV_{yy}),\qquad(b)\\ &\dot{x}=\lambda(2\dot{x}-2y),\qquad(c)\\ &\dot{y}=\lambda(2\dot{y}+2x).\qquad(d)\end{cases}

First, for λ=0\lambda=0, the equilibria of (1) are solutions to (15). We will denote Γ0:=±Q\Gamma_{0}:=\pm Q as either of the two equilibria since they are symmetric, notice that the corresponding r0α+2=αr_{0}^{\alpha+2}=\alpha. Next, it’s easy to see that λ12\lambda\neq\frac{1}{2}, we then get

x˙=2λy12λ,y˙=2λx12λ.\dot{x}=\frac{-2\lambda y}{1-2\lambda},\quad\dot{y}=\frac{2\lambda x}{1-2\lambda}.

Plug into (15)(a)(b) and K=0K=0 we will get

(16) x[(α+2)4(α+1)λ212λ+α(α+2)(αλ1)rα+2]=0,(a)y[4λ212λ+α(α+2)(αλ1)rα+2]=0,(b)4λ(1λ)(x2+y2)(12λ)2+(α+2)(x2αrα)=0.(c)\begin{split}&x[\frac{(\alpha+2)-4(\alpha+1)\lambda^{2}}{1-2\lambda}+\frac{\alpha(\alpha+2)(\alpha\lambda-1)}{r^{\alpha+2}}]=0,\qquad(a)\\ &y[\frac{4\lambda^{2}}{1-2\lambda}+\frac{\alpha(\alpha+2)(\alpha\lambda-1)}{r^{\alpha+2}}]=0,\qquad(b)\\ &\frac{4\lambda(1-\lambda)(x^{2}+y^{2})}{(1-2\lambda)^{2}}+(\alpha+2)(x^{2}-\frac{\alpha}{r^{\alpha}})=0.\qquad(c)\end{split}

Case I: When xy0xy\neq 0, from (16)(a)(b) we get λ=12\lambda=-\frac{1}{2}. Thus we get four solutions for (15), we will denote any one of the four solutions as Γ1\Gamma_{1} since they are symmetric with r1α+2=α(α+2)2r_{1}^{\alpha+2}=\alpha(\alpha+2)^{2}, and

(17) x12=αr1α+3r124(α+2),y12=αr1α+(4α+5)r124(α+2),x˙1=y12,y˙1=x12.\quad x_{1}^{2}=\frac{\alpha}{r_{1}^{\alpha}}+\frac{3r_{1}^{2}}{4(\alpha+2)},\quad y_{1}^{2}=-\frac{\alpha}{r_{1}^{\alpha}}+\frac{(4\alpha+5)r_{1}^{2}}{4(\alpha+2)},\quad\dot{x}_{1}=\frac{y_{1}}{2},\quad\dot{y}_{1}=-\frac{x_{1}}{2}.
Lemma 4.

The energy E(Γ0)<E(Γ1)<0E(\Gamma_{0})<E(\Gamma_{1})<0.

Proof.

We have

(18) E(Γ0)=α+22r02α+2r0α=(α+2)22r0α,E(\Gamma_{0})=-\frac{\alpha+2}{2}r_{0}^{2}-\frac{\alpha+2}{r_{0}^{\alpha}}=-\frac{(\alpha+2)^{2}}{2r_{0}^{\alpha}},

and

(19) E(Γ1)=18r12α+22(αr1α+3r124(α+2))α+2r1α=(α+2)34r1α.E(\Gamma_{1})=\frac{1}{8}r_{1}^{2}-\frac{\alpha+2}{2}(\frac{\alpha}{r_{1}^{\alpha}}+\frac{3r_{1}^{2}}{4(\alpha+2)})-\frac{\alpha+2}{r_{1}^{\alpha}}=-\frac{(\alpha+2)^{3}}{4r_{1}^{\alpha}}.

Notice that we have used r0α+2=αr_{0}^{\alpha+2}=\alpha and r1α+2=α(α+2)2r_{1}^{\alpha+2}=\alpha(\alpha+2)^{2}. Both energies are negative. Let

f(α):=E(Γ1)E(Γ0)=12(α+2)2αα+2,f(\alpha):=\frac{E(\Gamma_{1})}{E(\Gamma_{0})}=\frac{1}{2}(\alpha+2)^{\frac{2-\alpha}{\alpha+2}},

we have f(α)=f(α)(α+2)2[2α4ln(α+2)]<0f^{\prime}(\alpha)=\frac{f(\alpha)}{(\alpha+2)^{2}}[2-\alpha-4\ln(\alpha+2)]<0 for all α>0\alpha>0. Since f(0)=1f(0)=1, we get f(α)<1f(\alpha)<1, hence E(Γ0)<E(Γ1)E(\Gamma_{0})<E(\Gamma_{1}). ∎

Case II: When x=0x=0, then r2=y2r^{2}=y^{2}, from (16)(b)(c) we will get

(20) {4λ212λ=α(α+2)(1αλ)rα+2,4λ(1λ)(12λ)2=α(α+2)rα+2.\begin{cases}&\frac{4\lambda^{2}}{1-2\lambda}=\frac{\alpha(\alpha+2)(1-\alpha\lambda)}{r^{\alpha+2}},\\ &\frac{4\lambda(1-\lambda)}{(1-2\lambda)^{2}}=\frac{\alpha(\alpha+2)}{r^{\alpha+2}}.\end{cases}

Thus

λ(12λ)=(1αλ)(1λ),\lambda(1-2\lambda)=(1-\alpha\lambda)(1-\lambda),
(2+α)λ2(2+α)λ+1=0,\Rightarrow(2+\alpha)\lambda^{2}-(2+\alpha)\lambda+1=0,

whose solutions are λ±=12(1±α2α+2)\lambda_{\pm}=\frac{1}{2}(1\pm\sqrt{\frac{\alpha-2}{\alpha+2}}) when α>2\alpha>2. Thus we get four solutions to (15), two for each of the λ±\lambda_{\pm}. We will denote any one of the four solutions as Γ2\Gamma_{2} (and Γ2±\Gamma_{2\pm} for λ±\lambda_{\pm}). For both of the λ±\lambda_{\pm}, we have r2α+2=α(α+2)(α2)4r_{2}^{\alpha+2}=\frac{\alpha(\alpha+2)(\alpha-2)}{4}. Using the relation x˙=2λy12λ\dot{x}=\frac{-2\lambda y}{1-2\lambda}, we can show that

Lemma 5.

For α>2\alpha>2, the energy E(Γ2)<0<E(Γ2+)E(\Gamma_{2-})<0<E(\Gamma_{2+}) and E(Γ1)<E(Γ2)E(\Gamma_{1})<E(\Gamma_{2-}).

Proof.

Notice that Γ2\Gamma_{2} has the form (0,y,x˙,0)(0,y,\dot{x},0), we have

(21) E(Γ2)=124λ2r22(12λ)2α+2r2α.E(\Gamma_{2})=\frac{1}{2}\cdot\frac{4\lambda^{2}r_{2}^{2}}{(1-2\lambda)^{2}}-\frac{\alpha+2}{r_{2}^{\alpha}}.

Plug in λ=12(1±α2α+2)\lambda=\frac{1}{2}(1\pm\sqrt{\frac{\alpha-2}{\alpha+2}}) and r2α+2=α(α+2)(α2)4r_{2}^{\alpha+2}=\frac{\alpha(\alpha+2)(\alpha-2)}{4}, we get

(22) E(Γ2±)=(α+2)28r2α[α(1±α2α+2)28α+2],E(\Gamma_{2\pm})=\frac{(\alpha+2)^{2}}{8r_{2}^{\alpha}}[\alpha(1\pm\sqrt{\frac{\alpha-2}{\alpha+2}})^{2}-\frac{8}{\alpha+2}],

thus we have E(Γ2)<0<E(Γ2+)E(\Gamma_{2-})<0<E(\Gamma_{2+}). To compare E(Γ2)E(\Gamma_{2-}) and E(Γ1)E(\Gamma_{1}), let

(23) g(α):=E(Γ2)E(Γ1)=r1α2r2α(α+2)[8α+2α(1α2α+2)2]=12(α+2)(4(α+2)α2)αα+2[8α+2α(1α2α+2)2].\begin{split}g(\alpha)&:=\frac{E(\Gamma_{2-})}{E(\Gamma_{1})}=\frac{r_{1}^{\alpha}}{2r_{2}^{\alpha}(\alpha+2)}[\frac{8}{\alpha+2}-\alpha(1-\sqrt{\frac{\alpha-2}{\alpha+2}})^{2}]\\ &=\frac{1}{2(\alpha+2)}\cdot(\frac{4(\alpha+2)}{\alpha-2})^{\frac{\alpha}{\alpha+2}}[\frac{8}{\alpha+2}-\alpha(1-\sqrt{\frac{\alpha-2}{\alpha+2}})^{2}].\end{split}

Through Mathematica, we know g(α)g(\alpha) is deceasing for α>2\alpha>2, limα2+g(α)=1\lim\limits_{\alpha\to 2^{+}}g(\alpha)=1 and limαg(α)=0\lim\limits_{\alpha\to\infty}g(\alpha)=0. Thus 0<g(α)<10<g(\alpha)<1 for all α>2\alpha>2 and E(Γ1)<E(Γ2)<0E(\Gamma_{1})<E(\Gamma_{2-})<0.∎

Case III: When y=0y=0, then r2=x2r^{2}=x^{2}, from (16)(a)(c) we will get

(24) {(α+2)4(α+1)λ212λ=α(α+2)(1αλ)rα+2,4(α+1)(λ2λ)+α+2(12λ)2=α(α+2)rα+2.\begin{cases}&\frac{(\alpha+2)-4(\alpha+1)\lambda^{2}}{1-2\lambda}=\frac{\alpha(\alpha+2)(1-\alpha\lambda)}{r^{\alpha+2}},\\ &\frac{4(\alpha+1)(\lambda^{2}-\lambda)+\alpha+2}{(1-2\lambda)^{2}}=\frac{\alpha(\alpha+2)}{r^{\alpha+2}}.\end{cases}

Thus

[(α+2)4(α+1)λ2](12λ)=(1αλ)[4(α+1)(λ2λ)+α+2],[(\alpha+2)-4(\alpha+1)\lambda^{2}](1-2\lambda)=(1-\alpha\lambda)[4(\alpha+1)(\lambda^{2}-\lambda)+\alpha+2],
λ[4(α+1)(α+2)(λ2λ)+α(α+4)]=0,\Rightarrow\lambda[4(\alpha+1)(\alpha+2)(\lambda^{2}-\lambda)+\alpha(\alpha+4)]=0,

the discriminant of the quadratic equation of λ\lambda is 16(α+1)(α+2)(2α)16(\alpha+1)(\alpha+2)(2-\alpha). Therefore the only solution is λ=0\lambda=0 when α>2\alpha>2 and they coincide with the equilibria Γ0\Gamma_{0}.

Summary: the solutions of (15) with K=0K=0 are as follows:

  1. 1.

    λ0=0\lambda_{0}=0, this corresponds to the equilibria of the HLP, and the configuration points are on the xx-axis. The radius r0α+2=αr_{0}^{\alpha+2}=\alpha.

  2. 2.

    λ1=12\lambda_{1}=-\frac{1}{2}, there are four configuration points, they are not on the axes. The radius r1α+2=α(α+2)2r_{1}^{\alpha+2}=\alpha(\alpha+2)^{2}.

  3. 3.

    λ2±=12(1±α2α+2)\lambda_{2\pm}=\frac{1}{2}(1\pm\sqrt{\frac{\alpha-2}{\alpha+2}}), there are two configuration points, they are on the yy-axis. The radius r2α+2=α(α+2)(α2)4r_{2}^{\alpha+2}=\frac{\alpha(\alpha+2)(\alpha-2)}{4}.

  4. 4.

    The values of EE for these solutions satisfy

    E(Γ0)<E(Γ1)<E(Γ2)<0<E(Γ2+).E(\Gamma_{0})<E(\Gamma_{1})<E(\Gamma_{2-})<0<E(\Gamma_{2+}).

2.2. Lagrange equation with two constraints

In this subsection, let’s consider the Lagrange multiplier equation: E=λK+μW\nabla E=\lambda\nabla K+\mu\nabla W with K=0,W=0K=0,W=0. In particular, we have

(25) {Vx=λ(2y˙VxxVxxyVyx)+μ(Vx+xVxx+yVyx)(a)Vy=λ(2x˙VyxVxyyVyy)+μ(Vy+xVxy+yVyy)(b)x˙=λ(2x˙2y)(c)y˙=λ(2y˙+2x).(d)\begin{cases}&V_{x}=\lambda(2\dot{y}-V_{x}-xV_{xx}-yV_{yx})+\mu(V_{x}+xV_{xx}+yV_{yx})\qquad(a)\\ &V_{y}=\lambda(-2\dot{x}-V_{y}-xV_{xy}-yV_{yy})+\mu(V_{y}+xV_{xy}+yV_{yy})\qquad(b)\\ &\dot{x}=\lambda(2\dot{x}-2y)\qquad(c)\\ &\dot{y}=\lambda(2\dot{y}+2x).\qquad(d)\end{cases}

First notice that λ12\lambda\neq\frac{1}{2}, then we have

x˙=2λy12λ,y˙=2λx12λ.\dot{x}=\frac{-2\lambda y}{1-2\lambda},\quad\dot{y}=\frac{2\lambda x}{1-2\lambda}.

Plug into K=0K=0,

4λ(1λ)(x2+y2)(12λ)2+(α+2)(x2αrα)=0,\frac{4\lambda(1-\lambda)(x^{2}+y^{2})}{(1-2\lambda)^{2}}+(\alpha+2)(x^{2}-\frac{\alpha}{r^{\alpha}})=0,

together with W=0W=0 we get λ=0,1\lambda=0,1. Simplify (25) (a)(b) we get

(26) {x[2(α+2)(λμ)+(α+2)+4λ212λ+α(α+2)(α(λμ)1)rα+2]=0(a)y[4λ212λ+α(α+2)(α(λμ)1)rα+2]=0.(b)\begin{cases}&x[2(\alpha+2)(\lambda-\mu)+(\alpha+2)+\frac{4\lambda^{2}}{1-2\lambda}+\frac{\alpha(\alpha+2)(\alpha(\lambda-\mu)-1)}{r^{\alpha+2}}]=0\qquad(a)\\ &y[\frac{4\lambda^{2}}{1-2\lambda}+\frac{\alpha(\alpha+2)(\alpha(\lambda-\mu)-1)}{r^{\alpha+2}}]=0.\qquad(b)\end{cases}

If xy0xy\neq 0, from (26) we will get μλ=12\mu-\lambda=\frac{1}{2}. But when λ=0,1\lambda=0,1, (26)(b) will never be satisfied unless y=0y=0. Therefore we must have y=0y=0 (since x0x\neq 0 from the constraint W=0W=0). Thus the only solution for the two-constraint Lagrange equation is Γ0\Gamma_{0}, i.e. ±Q\pm Q.

2.3. Proof of Proposition 1

First, K(±Q)=0K(\pm Q)=0, thus inf{E|K=0,W=0}=E\inf\{E|K=0,W=0\}=E^{*} is obvious from Lemma 1. Let’s study inf{E|K0,W0}\inf\{E|K\geq 0,W\leq 0\}, first notice that the infimum of EE must be achieved, see Lemma 2 and Lemma 3 and the remark after them. Let’s study the Lagrange multiplier equations with inequality constraints. Introduce new variables p,qp,q, then K0K\geq 0 and W0W\leq 0 are equivalent to K~=Kp2=0\tilde{K}=K-p^{2}=0 and W~=W+q2=0\tilde{W}=W+q^{2}=0. Then the Lagrange equation E=λK~+μW~\nabla E=\lambda\nabla\tilde{K}+\mu\nabla\tilde{W} coincide with (25) with two extra equations: 0=2λp,0=2μq0=2\lambda p,0=2\mu q.

Case I: when pq0pq\neq 0, then λ=μ=0\lambda=\mu=0 the the only possible solution for the Lagrange equation is Γ0\Gamma_{0}, but W=q20W=-q^{2}\neq 0, thus no solutions.

Case II: when p=0,q=0p=0,q=0, then K=0,W=0K=0,W=0, the solution is Γ0\Gamma_{0} as been studied in section 2.2.

Case III: when p=0,q0p=0,q\neq 0, then μ=0\mu=0 and K=0K=0, the equations coincide with (15), and we have shown that the infimum can only be attained by Γ0\Gamma_{0} in section 2.1.

Case IV: when p0,q=0p\neq 0,q=0, then λ=0\lambda=0. We get x˙=y˙=0\dot{x}=\dot{y}=0, and V=μW\nabla V=\mu\nabla W, together with W=0W=0, it is easy to show that the only possible solution is Γ0\Gamma_{0}, but K=p20K=p^{2}\neq 0, thus no solutions. Therefore

inf{E|K0,W0}=E(Γ0)=E.\inf\{E|K\geq 0,W\leq 0\}=E(\Gamma_{0})=E^{*}.

3. below the ground state energy

Consider four sets in the phase space with energy below EE^{*}:

(27) 𝒲++={𝒲|K(Γ)0,W(Γ)>0},𝒲+={𝒲|K(Γ)<0,W(Γ)>0},𝒲+={𝒲|K(Γ)0,W(Γ)0},𝒲={𝒲|K(Γ)<0,W(Γ)0},\begin{split}\mathcal{W}_{+}^{+}&=\{\mathcal{W}|K(\Gamma)\geq 0,W(\Gamma)>0\},\\ \mathcal{W}_{+}^{-}&=\{\mathcal{W}|K(\Gamma)<0,W(\Gamma)>0\},\\ \mathcal{W}_{-}^{+}&=\{\mathcal{W}|K(\Gamma)\geq 0,W(\Gamma)\leq 0\},\\ \mathcal{W}_{-}^{-}&=\{\mathcal{W}|K(\Gamma)<0,W(\Gamma)\leq 0\},\\ \end{split}

where 𝒲={Γ=(x,y,x˙,y˙)4|E(Γ)<E}\mathcal{W}=\{\Gamma=(x,y,\dot{x},\dot{y})\in\mathbb{R}^{4}|E(\Gamma)<E^{*}\}. Notice that the upper right sign of 𝒲\mathcal{W} corresponds to KK, and the lower right sign corresponds to WW. We have 𝒲+=𝒲++𝒲+\mathcal{W}_{+}=\mathcal{W}_{+}^{+}\cup\mathcal{W}_{+}^{-} and 𝒲=𝒲+𝒲\mathcal{W}_{-}=\mathcal{W}_{-}^{+}\cup\mathcal{W}_{-}^{-}.

Refer to caption
(a) Level curves of V(x,y)EV(x,y)\leq E^{*}
Refer to caption
(b) V=EV=E^{*} (blue) and W=0W=0 (orange)
Figure 3. We have α=3\alpha=3 for both graphs. (A) indicates the level curves of V(x,y)V(x,y) below EE^{*}, and VV decreases to -\infty when approaching the origin. (B) gives the level curves of V(x,y)=EV(x,y)=E^{*} (blue) and W(x,y)=0W(x,y)=0 (orange): the bounded region A in (B) is {V<E,W0}=Proj(𝒲)\{V<E^{*},W\leq 0\}=\mathrm{Proj}(\mathcal{W}_{-}) and the two unbounded regions B in (B) is {V<E,W>0}=Proj(𝒲+)\{V<E^{*},W>0\}=\mathrm{Proj}(\mathcal{W}_{+}), where Proj(𝒲±)\mathrm{Proj}(\mathcal{W}_{\pm}) denotes the projection of 𝒲±\mathcal{W}_{\pm} into the configuration space (x,y)(x,y).
Lemma 6.

The sets 𝒲+\mathcal{W}_{+} and 𝒲\mathcal{W}_{-} are invariant for HLP.

Proof.

It is enough to show 𝒲+={Γ4|E(Γ)<E,W(Γ)>0}\mathcal{W}_{+}=\{\Gamma\in\mathbb{R}^{4}|E(\Gamma)<E^{*},W(\Gamma)>0\} is invariant. Notice that the energy is conserved, so E<EE<E^{*} implies V(x,y)<EV(x,y)<E^{*} in the configuration space (x,y)(x,y). Therefore the invariance of 𝒲+\mathcal{W}_{+} can be seen from the comparison of the level curves of V(x,y)=(α+2)(x22+1rα)<EV(x,y)=-(\alpha+2)(\frac{x^{2}}{2}+\frac{1}{r^{\alpha}})<E^{*} and W(x,y)=(α+2)(x2αrα)=0W(x,y)=(\alpha+2)(x^{2}-\frac{\alpha}{r^{\alpha}})=0 as illustrated in Figure 3. In particular, the bounded region A in Figure 3(B) corresponds to the projection of 𝒲\mathcal{W}_{-} in the (x,y)(x,y) configuration space, and the unbounded regions B in Figure 3(B) corresponds to the projection of 𝒲+\mathcal{W}_{+}. ∎

From Lemma 6, we know immediately that solutions in 𝒲+\mathcal{W}_{+} exist globally, since (x,y)(x,y) stays away from the origin. Now we focus on the set 𝒲\mathcal{W}_{-}.

Lemma 7.

For α2\alpha\geq 2, the set 𝒲+\mathcal{W}_{-}^{+} is empty.

Proof.

Use Proposition 1. ∎

Now we are ready to show that solutions with initial conditions in 𝒲\mathcal{W}_{-}^{-} must have collisions.

Lemma 8.

For α2\alpha\geq 2, given any δ>0\delta>0, there exists κ(δ)>0\kappa(\delta)>0 such that if E(Γ)EδE(\Gamma)\leq E^{*}-\delta, W(Γ)0W(\Gamma)\leq 0 then |K(Γ)|κ(δ)|K(\Gamma)|\geq\kappa(\delta).

Proof.

Suppose there is δ0>0\delta_{0}>0, and a sequence {Γn}\{\Gamma_{n}\} satisfying E(Γn)Eδ0E(\Gamma_{n})\leq E^{*}-\delta_{0} and W(Γn)0W(\Gamma_{n})\leq 0, but |K(Γn)|1n|K(\Gamma_{n})|\leq\frac{1}{n}. By Lemma 2 and Lemma 3, there exists M>0M>0, such that when nMn\geq M, the assumptions |K(Γn)|<1n|K(\Gamma_{n})|<\frac{1}{n} and E(Γn)Eδ0E(\Gamma_{n})\leq E^{*}-\delta_{0} imply that rn1M.r_{n}\geq\frac{1}{M}. Together with W(Γn)0W(\Gamma_{n})\leq 0, then there exists a regular point Γ\Gamma_{\infty} such that limnΓn=Γ\lim\limits_{n\to\infty}\Gamma_{n}=\Gamma_{\infty}. Therefore K(Γ)=0,W(Γ)0K(\Gamma_{\infty})=0,W(\Gamma_{\infty})\leq 0 and E(Γ)Eδ0E(\Gamma_{\infty})\leq E^{*}-\delta_{0} lead to a contradiction. ∎

Proposition 2.

For α2\alpha\geq 2, solutions in 𝒲\mathcal{W}_{-}^{-} are singular, in particular, finite time collision.

Proof.

Let Γ(t)\Gamma(t) be a solution in 𝒲\mathcal{W}_{-}^{-} and δ=EE(Γ)>0\delta=E^{*}-E(\Gamma)>0, let I(t):=I(Γ(t))I(t):=I(\Gamma(t)) and K(t):=K(Γ(t))K(t):=K(\Gamma(t)) then by Lemma 8

d2I(t)dt2=K(t)κ(δ)<0.\frac{d^{2}I(t)}{dt^{2}}=K(t)\leq-\kappa(\delta)<0.

Thus I(t)12κ(δ)t2+I˙(0)t+I(0)I(t)\leq-\frac{1}{2}\kappa(\delta)t^{2}+\dot{I}(0)t+I(0), since I=12(x2+y2)I=\frac{1}{2}(x^{2}+y^{2}) is always nonnegative, this implies the maximal time of existence of Γ(t)\Gamma(t) must be finite, i.e. there is a finite time collision. ∎

Notice that 𝒲=𝒲\mathcal{W}_{-}^{-}=\mathcal{W}_{-}, thus the fate of solutions for the Hill’s lunar problem with strong potential (α2)(\alpha\geq 2) below the ground state energy is entirely determined by the sign of WW. In particular, solutions in the bounded region (W0)(W\leq 0) are singular, and solutions in the unbounded region (W>0)(W>0) are global, cf. Figure 3. For α<2\alpha<2, the sign of KK in 𝒲\mathcal{W}_{-} is not sign-definite, and in fact there are both collision and global solution in 𝒲\mathcal{W}_{-}. See the numerical investigation about this fact in our companion paper [4]. Thus we have completed the proof of Theorem 1.

4. At the ground state energy threshold

In this section, we study the case when E=EE=E^{*}. We use the similar definition for the four sets as in (27) but now with E(Γ)=EE(\Gamma)=E^{*}. We will denote

𝒲~:={Γ=(x,y,x˙,y˙)4|E(Γ)=E}.\tilde{\mathcal{W}}:=\{\Gamma=(x,y,\dot{x},\dot{y})\in\mathbb{R}^{4}|E(\Gamma)=E^{*}\}.

From Lemma 7 we know 𝒲~+={±Q}\tilde{\mathcal{W}}_{-}^{+}=\{\pm Q\}, i.e. the equilibria (ground states) of the system. We have 𝒲~={𝒲~|K<0,W0}\tilde{\mathcal{W}}_{-}^{-}=\{\tilde{\mathcal{W}}|K<0,W\leq 0\}, then

Lemma 9.

For α2\alpha\geq 2, the set 𝒲~\tilde{\mathcal{W}}_{-}^{-} is invariant, thus the set 𝒲~+\tilde{\mathcal{W}}_{+} is also invariant.

Proof.

When E=EE=E^{*} we have

𝒲~=𝒲~+𝒲~={E=E,W0},\tilde{\mathcal{W}}_{-}=\tilde{\mathcal{W}}_{-}^{+}\cup\tilde{\mathcal{W}}_{-}^{-}=\{E=E^{*},W\leq 0\},

we can show 𝒲~\tilde{\mathcal{W}}_{-} is invariant though figure 3 as well as the invariance of 𝒲~+={E=E,W>0}\tilde{\mathcal{W}}_{+}=\{E=E^{*},W>0\}. Furthermore, since 𝒲~+={±Q}\tilde{\mathcal{W}}_{-}^{+}=\{\pm Q\} is invariant, so is 𝒲~\tilde{\mathcal{W}}_{-}^{-}. ∎

Proposition 3.

For α2\alpha\geq 2, solutions in 𝒲~\tilde{\mathcal{W}}_{-}^{-} either have a finite time collision or approach the ground state as tt\to\infty.

Proof.

Let Γ(t)\Gamma(t) be a solution in 𝒲~\tilde{\mathcal{W}}_{-}^{-}, if K(t)K(t) is uniformly bounded below zero, i.e. there exists δ>0\delta>0 so that K(t)δK(t)\leq-\delta, then from the proof of Proposition 2, we know Γ(t)\Gamma(t) experiences a finite time collision. Now if K(t)K(t) is not uniformly bounded below zero, that is, there is a sequence of time {tn}\{t_{n}\} so that K(tn)0K(t_{n})\to 0^{-} as tnσt_{n}\to\sigma, where σ\sigma is the maximal time of existence. Since I¨(t)=K(t)<0\ddot{I}(t)=K(t)<0, I˙(t)\dot{I}(t) is decreasing, thus limtσI˙(t)\lim_{t\to\sigma}\dot{I}(t) exists and cannot be -\infty due to limnK(tn)=0\lim_{n\to\infty}K(t_{n})=0. Assume limtσI˙(t)=a\lim_{t\to\sigma}\dot{I}(t)=a, if a<0a<0, then I(t)I(t) exists for finite time, i.e. there is a collision. On the other hand, if a0a\geq 0, then I˙(t)>0\dot{I}(t)>0 for all time. That is, I(t)I(t) is always increasing, thus σ=\sigma=\infty. Since I(t)I(t) is always bounded in 𝒲~\tilde{\mathcal{W}}_{-}^{-}, aa must be zero. Thus limtK(t)=0\lim_{t\to\infty}K(t)=0 and the solution Γ(t)\Gamma(t) approaches the set {E=E,K=0,W0}={±Q}\{E=E^{*},K=0,W\leq 0\}=\{\pm Q\} as tt\to\infty.

Moreover, we know the ground states as equilibria are of saddle-center type, see section 5 and Figure 4. The solution in 𝒲~\tilde{\mathcal{W}}_{-}^{-} which approaches ±Q\pm Q as tt\to\infty only when it lies on the stable manifolds of ±Q\pm Q.

Proposition 4.

Solutions in 𝒲~+={E=E,W>0}\tilde{\mathcal{W}}_{+}=\{E=E^{*},W>0\} exists for all time.

Proof.

Because every solution is uniformly away from the origin. ∎

Thus we have completed the proof of Theorem 2.

5. Above the ground state energy

Following Nakanishi-Schlag [10][11], to study the fate above the ground state energy with E<E+ϵ2E<E^{*}+\epsilon^{2} we need to consider two scenarios: first, the stable behavior of solutions near the ground states on the center-stable manifold of ±Q\pm Q; second, the solutions that enter, but then again leave, a neighborhood of the ground states. These are the non-trapped trajectories. We will only describe the dynamics of non-trapped trajectories after they exit a ball of size 2ϵ2\epsilon, where the size is given under a suitable distance function relative to the ground states. We treat the region inside the 2ϵ2\epsilon ball around the ground states as a “black box” and we do not analyze at all, since we are either trapped or being ejected out of that ball, and then carried to much larger distances from the ground states.

For the non-trapped trajectories, we study the Hill’s Lunar Problem concerning the properties:

1. “ejection lemma”, that is, solutions that do not remain close to the ground state for all positive times are ejected from any small neighborhood of it after some positive time. Moreover, the solution cannot return to that neighborhood, which is known as the “one-pass theorem” or “no-return” property.

2. The variational estimates of the ground state energy and K,WK,W.

In the PDE case, Nakanishi-Schlag [10] proved the existence of the center-stable (unstable etc.) invariant manifolds of the ground state in the infinity dimensional case. For the Hill’s Lunar problem, the phase space is of dimension four, and we know ±Q\pm Q are of center-saddle type. In particular, the matrix associated with the linearized Hill’s equation at each ground state have a pair of real eigenvalues of equal magnitude and opposite sign

(28) ±k:=±1236+36α+29α2+10α3+α4+(α2+3α2),\begin{split}\pm k:=\pm\frac{1}{\sqrt{2}}\sqrt{\sqrt{36+36\alpha+29\alpha^{2}+10\alpha^{3}+\alpha^{4}}+(\alpha^{2}+3\alpha-2)},\end{split}

and a pair of pure imaginary complex conjugate eigenvalues

(29) ±iω:=±i236+36α+29α2+10α3+α4(α2+3α2).\begin{split}\pm i\omega:=\pm\frac{i}{\sqrt{2}}\sqrt{\sqrt{36+36\alpha+29\alpha^{2}+10\alpha^{3}+\alpha^{4}}-(\alpha^{2}+3\alpha-2)}.\end{split}

By standard invariant manifolds theory (cf. [6]) in dynamical systems, the stable, unstable, center-stable, center-unstable manifolds exist for ±Q\pm Q. In Figure 4, we give the tangent line of the projection of the stable and unstable manifold into the configuration space at the Lagrange point L2L_{2}. The qualitative picture of the flow near the equilibrium resemble the linearized flow as we shall show in the following. In order to take full advantage of the Hamiltonian nature of the HLP, we will use symplectic coordinates in the following.

Refer to caption
Figure 4. Zero velocity curve (black solid) of the ground state energy, and tangent lines (blue and red) of the projection of the stable and unstable manifold into the configuration space at the Lagrange point L2L_{2}.

5.1. Symplectic algebra and Hill’s equations in Symplectic form

Let J=(0II0)J=\begin{pmatrix}0&I\\ -I&0\end{pmatrix} be the standard symplectic matrix in 2n\mathbb{R}^{2n} and Ω(ξ,η)=Jξη=ξTJTη\Omega(\xi,\eta)=J\xi\cdot\eta=\xi^{T}J^{T}\eta be the standard symplectic form on 2n\mathbb{R}^{2n}. The symplectic group, i.e. the linear transformations that preserves Ω\Omega, is

Sp(2n)={A2n×2n|ATJA=J}.Sp(2n)=\{A\in\mathbb{R}^{2n\times 2n}|A^{T}JA=J\}.

The infinitesimally symplectic group is

sp(2n)={A2n×2n|ATJ+JA=0}.sp(2n)=\{A\in\mathbb{R}^{2n\times 2n}|A^{T}J+JA=0\}.

Note that if Asp(2n)A\in sp(2n) then eASp(2n)e^{A}\in Sp(2n). For a linear Hamiltonian equation X˙=JSX\dot{X}=JSX where ST=SS^{T}=S, the matrix A=JSsp(2n)A=JS\in sp(2n) and etJSSp(2n)e^{tJS}\in Sp(2n) (cf. [1]).

Lemma 10.

If Asp(2n)A\in sp(2n), then Ω(Aξ,η)=Ω(ξ,Aη)\Omega(A\xi,\eta)=-\Omega(\xi,A\eta). Thus if ξ,η\xi,\eta are eigenvectors of AA with eigenvalues λ,μ\lambda,\mu, then Ω(ξ,η)=0\Omega(\xi,\eta)=0 provided λ+μ0\lambda+\mu\neq 0.

Proof.
Ω(Aξ,η)=ξTATJTη=ξTJAη=ξTJTAη=Ω(ξ,Aη).\Omega(A\xi,\eta)=\xi^{T}A^{T}J^{T}\eta=\xi^{T}JA\eta=-\xi^{T}J^{T}A\eta=-\Omega(\xi,A\eta).

Then second statement follows directly from this property. ∎

Consider the Hill’s equations with coordinates q=(x,y)q=(x,y) and p=(px,py)=(x˙y,y˙+x)p=(p_{x},p_{y})=(\dot{x}-y,\dot{y}+x), the Hamiltonian, i.e. the energy is

(30) E(x,y,px,py)=12[(px+y)2+(pyx)2]+V(x,y),\begin{split}E(x,y,p_{x},p_{y})=\frac{1}{2}[(p_{x}+y)^{2}+(p_{y}-x)^{2}]+V(x,y),\end{split}

and the function KK (cf. (11)) is

(31) K(x,y,px,py)=px2+py2+(α+1)x2y2α(α+2)rα.\begin{split}K(x,y,p_{x},p_{y})=p_{x}^{2}+p_{y}^{2}+(\alpha+1)x^{2}-y^{2}-\alpha(\alpha+2)r^{-\alpha}.\end{split}

The Hill’s equations (1) in Symplectic canonical form is

(32) q˙=Ep,p˙=Eq.\dot{q}=\frac{\partial E}{\partial p},\quad\dot{p}=-\frac{\partial E}{\partial q}.

That is,

(q˙p˙)=JE.\begin{pmatrix}\dot{q}\\ \dot{p}\end{pmatrix}=J\nabla E.

The ground states in the (q,p)(q,p) coordinates are

(33) ±Q~:=±(q0,0,0,q0),q0=α1α+2.\begin{split}\pm\tilde{Q}:=\pm(q_{0},0,0,q_{0}),\quad q_{0}=\alpha^{\frac{1}{\alpha+2}}.\end{split}

Linearize (32) at Q~\tilde{Q} one obtains

(q˙p˙)=J2E(Q~)(qp),\begin{pmatrix}\dot{q}\\ \dot{p}\end{pmatrix}=J\nabla^{2}E(\tilde{Q})\begin{pmatrix}q\\ p\end{pmatrix},

where

(34) 2E(Q~)=(1(α+2)200101+(α+2)1001101001).\nabla^{2}E(\tilde{Q})=\begin{pmatrix}1-(\alpha+2)^{2}&0&0&-1\\ 0&1+(\alpha+2)&1&0\\ 0&1&1&0\\ -1&0&0&1\end{pmatrix}.

Note that it is easy to check

(35) 2E(Q~)=2E(Q~),\begin{split}\nabla^{2}E(\tilde{Q})=\nabla^{2}E(-\tilde{Q}),\end{split}

which is extremely helpful in our analysis.

Let A=J2E(Q~)A=J\nabla^{2}E(\tilde{Q}), the eigenvalues of AA are ±k,±iω\pm k,\pm i\omega. Let ξ+\xi_{+} and ξ\xi_{-} be the eigenvectors of kk and k-k respectively, and we assume ξ+\xi_{+} and ξ\xi_{-} are normalized in the following sense:

(36) Ω(ξ+,ξ)=1,Ω(ξ,ξ+)=1,|ξ+|=|ξ|,\Omega(\xi_{+},\xi_{-})=1,\quad\Omega(\xi_{-},\xi_{+})=-1,\quad|\xi_{+}|=|\xi_{-}|,

where |||\cdot| denotes the Euclidean norm of a vector. It is obvious that

(37) Ω(ξ+,ξ+)=0,Ω(ξ,ξ)=0.\Omega(\xi_{+},\xi_{+})=0,\quad\Omega(\xi_{-},\xi_{-})=0.

See Figure 4 for the projection of ξ+,ξ\xi_{+},\xi_{-} to the configuration space at Q~\tilde{Q}.

Write (q,p)=Q~+X(q,p)=\tilde{Q}+X, where XX is the perturbation and we have

(38) X˙=J2E(Q~)X+N(X)=AX+N(X),\dot{X}=J\nabla^{2}E(\tilde{Q})X+N(X)=AX+N(X),

where N(X)N(X) stands for the non-linear terms. Decompose 4\mathbb{R}^{4} as EuEsEcE^{u}\bigoplus E^{s}\bigoplus E^{c} where Eu,s,cE^{u,s,c} are the eigenspace for k,k,±iωk,-k,\pm i\omega respectively. Write XX as follows:

(39) X=λ+(t)ξ++λ(t)ξ+γ(t),X=\lambda_{+}(t)\xi_{+}+\lambda_{-}(t)\xi_{-}+\gamma(t),

where

(40) γ(t)Ec,Ω(γ(t),ξ+)=Ω(γ(t),ξ)=0.\gamma(t)\in E^{c},\quad\Omega(\gamma(t),\xi_{+})=\Omega(\gamma(t),\xi_{-})=0.

One has λ±=±Ω(X,ξ)\lambda_{\pm}=\pm\Omega(X,\xi_{\mp}) and we can derive the differential equations for λ±(t)\lambda_{\pm}(t).

(41) dλ±dt(t)=±Ω(X˙,ξ)=±Ω(AX+N(X),ξ)=±Ω(AX,ξ)±Ω(N(X),ξ)=±kλ±(t)±Ω(N(X),ξ).\begin{split}\frac{d\lambda_{\pm}}{dt}(t)&=\pm\Omega(\dot{X},\xi_{\mp})\\ &=\pm\Omega(AX+N(X),\xi_{\mp})\\ &=\pm\Omega(AX,\xi_{\mp})\pm\Omega(N(X),\xi_{\mp})\\ &=\pm k\lambda_{\pm}(t)\pm\Omega(N(X),\xi_{\mp}).\end{split}

Thus

(42) dλ+dt(t)=kλ+(t)+N+(X),dλdt(t)=kλ(t)N(X),\begin{split}\frac{d\lambda_{+}}{dt}(t)=k\lambda_{+}(t)+N_{+}(X),\quad\frac{d\lambda_{-}}{dt}(t)=-k\lambda_{-}(t)-N_{-}(X),\end{split}

where

(43) N+(X)=Ω(N(X),ξ),N(X)=Ω(N(X),ξ+).\begin{split}N_{+}(X)=\Omega(N(X),\xi_{-}),\quad N_{-}(X)=\Omega(N(X),\xi_{+}).\end{split}

Moreover, we have

(44) λ±2=Ω(X,ξ)2|ξ+|2|X|2.\lambda^{2}_{\pm}=\Omega(X,\xi_{\mp})^{2}\leq|\xi_{+}|^{2}|X|^{2}.

5.2. Linearized energy norm

In this section we want to define the “linearized energy norm” for the perturbation XX. First, as a result of the symplectic orthogonality of ξ±\xi_{\pm} and γ\gamma, we have the following relationship:

Lemma 11.

The function γ(t)\gamma(t) in the decomposition (39) satisfies

Ω(γ,Aγ)|γ|2.\Omega(\gamma,A\gamma)\sim|\gamma|^{2}.
Proof.

Let η=η1+iη2\eta=\eta_{1}+i\eta_{2} be the eigenvector of iωi\omega, thus Ec=Span{η1,η2}E^{c}=\mathrm{Span}\{\eta_{1},\eta_{2}\} and we have Aη1=ωη2A\eta_{1}=-\omega\eta_{2}, Aη2=ωη1A\eta_{2}=\omega\eta_{1}. One can check that Ω(η1,η2)>0\Omega(\eta_{1},\eta_{2})>0, thus we shall assume η1,η2\eta_{1},\eta_{2} are normalized in the sense that

Ω(η1,η2)=1.\Omega(\eta_{1},\eta_{2})=1.

Thus ξ+,ξ,η1,η2\xi_{+},\xi_{-},\eta_{1},\eta_{2} form a symplectic basis. In particular, the matrix P=(ξ+η1ξη2)Sp(4)P=(\xi_{+}\,\,\eta_{1}\,\,\xi_{-}\,\,\eta_{2})\in Sp(4) and A=PBP1A=PBP^{-1} where

B=(k000000ω00k00ω00).B=\begin{pmatrix}k&0&0&0\\ 0&0&0&\omega\\ 0&0&-k&0\\ 0&-\omega&0&0\end{pmatrix}.

Since γEc\gamma\in E^{c}, we assume γ=aη1+bη2\gamma=a\eta_{1}+b\eta_{2}, thus P1γ=(0,a,0,b)P^{-1}\gamma=(0,a,0,b).

(45) Ω(γ,Aγ)=Ω(γ,PBP1γ)=Ω(P1γ,BP1γ)=ω(a2+b2).\begin{split}\Omega(\gamma,A\gamma)&=\Omega(\gamma,PBP^{-1}\gamma)\\ &=\Omega(P^{-1}\gamma,BP^{-1}\gamma)\\ &=\omega(a^{2}+b^{2}).\end{split}

The second equality is due to that PP is symplectic. The last equality is by direct computations. Since η1,η2,ω\eta_{1},\eta_{2},\omega are constant quantities, we conclude that Ω(γ,Aγ)|γ|2\Omega(\gamma,A\gamma)\sim|\gamma|^{2}. ∎

By Taylor expansion and analysis in section 5.1, we have

(46) E(Q~+X)E(Q~)=12XT2E(Q~)X+C(X)=12XTJTJ2E(Q~)X+C(X)=12Ω(X,AX)+C(X)=kλ+λ+12Ω(γ,Aγ)+C(X),\begin{split}E(\tilde{Q}+X)-E(\tilde{Q})&=\frac{1}{2}X^{T}\nabla^{2}E(\tilde{Q})X+C(X)\\ &=\frac{1}{2}X^{T}J^{T}J\nabla^{2}E(\tilde{Q})X+C(X)\\ &=\frac{1}{2}\Omega(X,AX)+C(X)\\ &=-k\lambda_{+}\lambda_{-}+\frac{1}{2}\Omega(\gamma,A\gamma)+C(X),\end{split}

note that C(X)=O(|X|3)C(X)=O(|X|^{3}). We define the linearized energy norm |X|E|X|_{E} by

(47) |X|E2:=k2(λ+2(t)+λ2(t))+12Ω(γ,Aγ).|X|_{E}^{2}:=\frac{k}{2}(\lambda_{+}^{2}(t)+\lambda_{-}^{2}(t))+\frac{1}{2}\Omega(\gamma,A\gamma).

Then we get that

(48) E(Q~+X)E(Q~)+k2(λ+(t)+λ(t))2|X|E2=O(|X|3).E(\tilde{Q}+X)-E(\tilde{Q})+\frac{k}{2}(\lambda_{+}(t)+\lambda_{-}(t))^{2}-|X|_{E}^{2}=O(|X|^{3}).
Lemma 12.

The function γ(t)\gamma(t) in the decomposition satisfies

|γ(t)||X(t)|,|\gamma(t)|\lesssim|X(t)|,

moreover,

|X(t)||X(t)|E.|X(t)|\sim|X(t)|_{E}.
Proof.

By the triangular inequality, Lemma 11 and (44).∎

5.3. Distance function from the ground state

We introduce the distance function relative to the ground states ±Q~\pm\tilde{Q}. Let

ψ=σ(Q~+X),X=λ+ξ++λξ+γ,\psi=\sigma(\tilde{Q}+X),\quad X=\lambda_{+}\xi_{+}+\lambda_{-}\xi_{-}+\gamma,

where λ±(t)\lambda_{\pm}(t) and γ(t)\gamma(t) are functions appearing in the decomposition for X(t)X(t) as in (39), and σ=±\sigma=\pm with “+” corresponds to the ground state Q~\tilde{Q} and “-” corresponds to the ground state Q~-\tilde{Q}, note that E(Q~)=E(Q~)E(\tilde{Q})=E(-\tilde{Q}).

From Lemma 12 and (48) we see that there exists a constant δE>0\delta_{E}>0 with the following property: for any solution ψ=σ(Q~+X)\psi=\sigma(\tilde{Q}+X) to (32) and any time tImax(ψ)t\in I_{max}(\psi) (i.e. the maximal interval of existence for ψ\psi) for which |X(t)|E4δE|X(t)|_{E}\leq 4\delta_{E},

(49) |E(ψ(t))E(Q~)+k2(λ+(t)+λ(t))2|X(t)|E2||X(t)|E210.|E(\psi(t))-E(\tilde{Q})+\frac{k}{2}(\lambda_{+}(t)+\lambda_{-}(t))^{2}-|X(t)|_{E}^{2}|\leq\frac{|X(t)|_{E}^{2}}{10}.

Let χ\chi be a smooth function on \mathbb{R} such that χ(r)=1\chi(r)=1 for |r|1|r|\leq 1 and χ(r)=0\chi(r)=0 for |r|2|r|\geq 2. We define

dσ(ψ(t)):=|X(t)|E2+χ(|X(t)|E/2δE)C(ψ(t)),d_{\sigma}(\psi(t)):=\sqrt{|X(t)|_{E}^{2}+\chi(|X(t)|_{E}/2\delta_{E})C(\psi(t))},

where

C(ψ(t))=E(ψ(t))E(Q~)+k2(λ+(t)+λ(t))2|X(t)|E2,C(\psi(t))=E(\psi(t))-E(\tilde{Q})+\frac{k}{2}(\lambda_{+}(t)+\lambda_{-}(t))^{2}-|X(t)|_{E}^{2},

as appeared in (46). Moreover, we will let

dQ(ψ(t)):=min(d+(ψ(t)),d(ψ(t)))=dσ(ψ(t)),d_{Q}(\psi(t)):=\min(d_{+}(\psi(t)),d_{-}(\psi(t)))=d_{\sigma}(\psi(t)),

notice that σ\sigma is unique as long as dQ(ψ)d_{Q}(\psi) is less than half the distance of the two ground states.

It will be convenient to introduce the new parameters

λ1(t):=λ+(t)+λ(t)2,λ2(t):=λ+(t)λ(t)2,\lambda_{1}(t):=\frac{\lambda_{+}(t)+\lambda_{-}(t)}{2},\quad\lambda_{2}(t):=\frac{\lambda_{+}(t)-\lambda_{-}(t)}{2},

and we have

(50) dλ1dt(t)=kλ2(t)+12N1(X),dλ2dt(t)=kλ1(t)12N2(X),\begin{split}\frac{d\lambda_{1}}{dt}(t)=k\lambda_{2}(t)+\frac{1}{2}N_{1}(X),\quad\frac{d\lambda_{2}}{dt}(t)=k\lambda_{1}(t)-\frac{1}{2}N_{2}(X),\end{split}

where

(51) N1(X)=Ω(N(X),ξ++ξ),N2(X)=Ω(N(X),ξ+ξ).\begin{split}N_{1}(X)=\Omega(N(X),\xi_{+}+\xi_{-}),\quad N_{2}(X)=\Omega(N(X),\xi_{+}-\xi_{-}).\end{split}

Note that

(52) |N1(X)||X|2,|N2(X)||X|2.\begin{split}|N_{1}(X)|\lesssim|X|^{2},\quad|N_{2}(X)|\lesssim|X|^{2}.\end{split}

Some important properties about the function dQ(ψ(t))d_{Q}(\psi(t)) are in the following:

Lemma 13.

Assume that there exists an interval II on which

suptIdQ(ψ(t))δE.\mathrm{sup}_{t\in I}d_{Q}(\psi(t))\leq\delta_{E}.

Then, all of the following hold for all tIt\in I:

  • (i)

    12|X(t)|E2dQ(ψ(t))232|X(t)|E2,\frac{1}{2}|X(t)|^{2}_{E}\leq d_{Q}(\psi(t))^{2}\leq\frac{3}{2}|X(t)|^{2}_{E},

  • (ii)

    dQ(ψ(t))2=E(ψ(t))E(Q~)+2kλ12(t),d_{Q}(\psi(t))^{2}=E(\psi(t))-E(\tilde{Q})+2k\lambda_{1}^{2}(t),

  • (iii)

    ddtdQ(ψ(t))2=4k2λ1(t)λ2(t)+2kλ1(t)N1(X)\frac{d}{dt}d_{Q}(\psi(t))^{2}=4k^{2}\lambda_{1}(t)\lambda_{2}(t)+2k\lambda_{1}(t)N_{1}(X).

Proof.

The proof of (i) is directly from the definition of dQ(ψ)d_{Q}(\psi) and (49). To prove (ii), notice that from (i) and dQ(ψ(t))δEd_{Q}(\psi(t))\leq\delta_{E}, we have |X(t)|E<2δE|X(t)|_{E}<2\delta_{E}, thus

(53) dQ(ψ(t))2=|X(t)|E2+C(ψ(t))=E(ψ(t))E(Q~)+2kλ12(t).d_{Q}(\psi(t))^{2}=|X(t)|_{E}^{2}+C(\psi(t))=E(\psi(t))-E(\tilde{Q})+2k\lambda_{1}^{2}(t).

To show (iii), we differentiate (ii) with respect to tt on both sides:

(54) ddtdQ(ψ(t))2=E(ψ(t))ψ˙(t)+4kλ1λ1˙=E(ψ(t))JE(ψ(t))+4kλ1λ1˙=0+4kλ1[kλ2(t)+12N1(X)]=4k2λ1(t)λ2(t)+2kλ1(t)N1(X).\begin{split}\frac{d}{dt}d_{Q}(\psi(t))^{2}&=\nabla E(\psi(t))\cdot\dot{\psi}(t)+4k\lambda_{1}\dot{\lambda_{1}}\\ &=\nabla E(\psi(t))\cdot J\nabla E(\psi(t))+4k\lambda_{1}\dot{\lambda_{1}}\\ &=0+4k\lambda_{1}[k\lambda_{2}(t)+\frac{1}{2}N_{1}(X)]\\ &=4k^{2}\lambda_{1}(t)\lambda_{2}(t)+2k\lambda_{1}(t)N_{1}(X).\end{split}

Moreover, we can show that when the energy is at most slightly above the ground state energy, and the solution is near the ground state, then the distance function is dominated by |λ1||\lambda_{1}|.

Lemma 14 (Eigenmode dominance).

For any ψ4\psi\in\mathbb{R}^{4} such that

(55) E(ψ)<E+12dQ(ψ)2,dQ(ψ)δE,\begin{split}E(\psi)<E^{*}+\frac{1}{2}d_{Q}(\psi)^{2},\quad d_{Q}(\psi)\leq\delta_{E},\end{split}

one has dQ(ψ)|λ1|d_{Q}(\psi)\sim|\lambda_{1}|. Moreover, λ1\lambda_{1} has a fixed sign in each connected component of the region (55).

Proof.

We have

(56) |X|E2=E(ψ)E+k2(λ++λ)2C(ψ)E(ψ)E+2kλ12+|X|E210,\begin{split}|X|_{E}^{2}&=E(\psi)-E^{*}+\frac{k}{2}(\lambda_{+}+\lambda_{-})^{2}-C(\psi)\\ &\leq E(\psi)-E^{*}+2k\lambda_{1}^{2}+\frac{|X|_{E}^{2}}{10},\end{split}

thus

910|X|E2E(ψ)E+2kλ12.\frac{9}{10}|X|_{E}^{2}\leq E(\psi)-E^{*}+2k\lambda_{1}^{2}.

Using Lemma 13 (i), we obtain

(57) dQ(ψ)232|X|E256dQ(ψ)2+103kλ12,d_{Q}(\psi)^{2}\leq\frac{3}{2}|X|_{E}^{2}\leq\frac{5}{6}d_{Q}(\psi)^{2}+\frac{10}{3}k\lambda_{1}^{2},

hence

(58) dQ(ψ)220kλ12.d_{Q}(\psi)^{2}\leq 20k\lambda_{1}^{2}.

On the other hand,

(59) λ1212(λ+2+λ2)|ξ+|2|X|2dQ(ψ)2,\lambda_{1}^{2}\leq\frac{1}{2}(\lambda_{+}^{2}+\lambda_{-}^{2})\leq|\xi_{+}|^{2}|X|^{2}\lesssim d_{Q}(\psi)^{2},

where the last inequality is from Lemma 12 and Lemma 13 (i). Finally, inside the set (55) one can never have λ1=0\lambda_{1}=0, since that would mean both dQ(ψ)=0d_{Q}(\psi)=0 and E(ψ)<EE(\psi)<E^{*} which is impossible. ∎

5.4. The ejection process

Lemma 15 (Ejection Lemma).

There exists constants 0<δXδE0<\delta_{X}\leq\delta_{E} and C,TC_{*},T_{*} with the property: If ψ(t)\psi(t) is a local solution to (32) on [0,T][0,T] satisfying

(60) R:=dQ(ψ(0))δX,E(ψ)<E+12R2,R:=d_{Q}(\psi(0))\leq\delta_{X},\quad E(\psi)<E^{*}+\frac{1}{2}R^{2},

then we can extend ψ(t)\psi(t) as long as dQ(ψ(t))δXd_{Q}(\psi(t))\leq\delta_{X}. Furthermore, if there exists some t0(0,T)t_{0}\in(0,T) such that

(61) dQ(ψ(t))R,0<t<t0,d_{Q}(\psi(t))\geq R,\quad\forall 0<t<t_{0},

and let

TX:=inf{t[0,t0]:dQ(ψ(t))=δX}T_{X}:=\inf\{t\in[0,t_{0}]:d_{Q}(\psi(t))=\delta_{X}\}

where TX=t0T_{X}=t_{0} if dQ(ψ(t))<δXd_{Q}(\psi(t))<\delta_{X} on [0,t0][0,t_{0}], then for all t[0,TX]:t\in[0,T_{X}]:

  • (i)

    dQ(ψ(t))𝔰λ1(t)𝔰λ+(t)ektRd_{Q}(\psi(t))\sim\mathfrak{s}\lambda_{1}(t)\sim\mathfrak{s}\lambda_{+}(t)\sim e^{kt}R,

  • (ii)

    |λ(t)|+|γ(t)|R+R32|\lambda_{-}(t)|+|\gamma(t)|\lesssim R+R^{\frac{3}{2}},

  • (iii)

    𝔰K(ψ(t))dQ(ψ(t))CdQ(ψ(0))\mathfrak{s}K(\psi(t))\gtrsim d_{Q}(\psi(t))-C_{*}d_{Q}(\psi(0)), and 𝔰W(ψ(t))dQ(ψ(t))CdQ(ψ(0))\mathfrak{s}W(\psi(t))\gtrsim d_{Q}(\psi(t))-C_{*}d_{Q}(\psi(0)).

where 𝔰=+1\mathfrak{s}=+1 or 1-1. Moreover, dQ(ψ(t))d_{Q}(\psi(t)) is increasing for t[TR,TX]t\in[T_{*}R,T_{X}] and |dQ(ψ(t))R|R3|d_{Q}(\psi(t))-R|\lesssim R^{3} for t[0,TR]t\in[0,T_{*}R].

Proof.

First notice that when the solution is close to ±Q~\pm\tilde{Q} means it is away from the singular point, thus the solution extends as long as dQ(ψ(t))d_{Q}(\psi(t)) is small. Next, from the definition of TXT_{X} and (61) we have for any t[0,TX]t\in[0,T_{X}],

(62) RdQ(ψ(t))δX<δE,\begin{split}R\leq d_{Q}(\psi(t))\leq\delta_{X}<\delta_{E},\end{split}

thus from Lemma 13 and Lemma 14 for any t[0,TX]t\in[0,T_{X}],

(63) |γ(t)||X(t)||X(t)|EdQ(ψ(t))|λ1(t)|,\begin{split}|\gamma(t)|\lesssim|X(t)|\sim|X(t)|_{E}\sim d_{Q}(\psi(t))\sim|\lambda_{1}(t)|,\end{split}
(64) ddtdQ(ψ(t))2=4k2λ1(t)λ2(t)+2kλ1(t)N1(X).\begin{split}\frac{d}{dt}d_{Q}(\psi(t))^{2}=4k^{2}\lambda_{1}(t)\lambda_{2}(t)+2k\lambda_{1}(t)N_{1}(X).\end{split}

Hence for any t[0,TX]t\in[0,T_{X}],

(65) 0<R|λ1(t)|,\begin{split}0<R\lesssim|\lambda_{1}(t)|,\end{split}

which together with the continuity of λ1(t)\lambda_{1}(t) shows that for any t[0,TX]t\in[0,T_{X}]

(66) 𝔰:=sign[λ1(t)]=sign[λ1(0)].\begin{split}\mathfrak{s}:=\mathrm{sign}[\lambda_{1}(t)]=\mathrm{sign}[\lambda_{1}(0)].\end{split}

The ejection condition (61) implies ddtdQ(ψ(t))2|t=00\frac{d}{dt}d_{Q}(\psi(t))^{2}|_{t=0}\geq 0. Since |N1(X)||X|2λ12|N_{1}(X)|\lesssim|X|^{2}\lesssim\lambda_{1}^{2}, we deduce from (64) that 𝔰λ2(0)|λ1(0)|2\mathfrak{s}\lambda_{2}(0)\gtrsim-|\lambda_{1}(0)|^{2} and so λ+(0)λ1(0)𝔰R\lambda_{+}(0)\sim\lambda_{1}(0)\sim\mathfrak{s}R.

Integrating the equation for λ±\lambda_{\pm} yields

(67) |λ±(t)e±ktλ±(0)|0tek(ts)|N±(X(s))|𝑑s0tek(ts)|λ1(s)|2𝑑s,\begin{split}|\lambda_{\pm}(t)-e^{\pm kt}\lambda_{\pm}(0)|\lesssim\int_{0}^{t}e^{k(t-s)}|N_{\pm}(X(s))|ds\lesssim\int_{0}^{t}e^{k(t-s)}|\lambda_{1}(s)|^{2}ds,\end{split}

thus by continuity in time we deduce that as long as Rekt1Re^{kt}\ll 1, we have

(68) λ1(t)λ+(t)𝔰Rekt,|λ±(t)e±ktλ±(0)|R2e2kt.\begin{split}\lambda_{1}(t)\sim\lambda_{+}(t)\sim\mathfrak{s}Re^{kt},\quad|\lambda_{\pm}(t)-e^{\pm kt}\lambda_{\pm}(0)|\lesssim R^{2}e^{2kt}.\end{split}

Thus completes the proof of (i) and that |λ(t)|R+R2e2kt|\lambda_{-}(t)|\lesssim R+R^{2}e^{2kt}. Now to estimate |γ(t)||\gamma(t)|, we introduce the nonlinear energy projected to {ξ+,ξ}\{\xi_{+},\xi_{-}\} plane (cf. (46)):

(69) E{ξ+,ξ}(t):=E(Q~+λ+ξ++λξ)E(Q~)=kλ+λ+C(λ+ξ++λξ).\begin{split}E_{\{\xi_{+},\xi_{-}\}}(t):=E(\tilde{Q}+\lambda_{+}\xi_{+}+\lambda_{-}\xi_{-})-E(\tilde{Q})=-k\lambda_{+}\lambda_{-}+C(\lambda_{+}\xi_{+}+\lambda_{-}\xi_{-}).\end{split}

Denote ξ(t):=λ+(t)ξ++λ(t)ξ\xi(t):=\lambda_{+}(t)\xi_{+}+\lambda_{-}(t)\xi_{-}, we get

(70) |ddtE{ξ+,ξ}(t)|=|E(Q~+ξ)ξ˙|=|(2E(Q~)ξ+O(|ξ|2))ξ˙||2E(Q~)ξξ˙|+|ξ|2|ξ˙|=|Ω(ξ˙,Aξ)|+|ξ|2|ξ˙|=k|λ˙+λ+λ+λ˙|+|ξ|2|ξ˙|=k|λN+(X)+λ+N(X)|+|ξ|2|ξ˙|(|λ|+|λ+|)3+(|λ|+|λ+|)2(|λ˙+|+|λ˙|)(|λ|+|λ+|)3+(|λ|+|λ+|)2|γ(t)|2.\begin{split}|\frac{d}{dt}E_{\{\xi_{+},\xi_{-}\}}(t)|&=|\nabla E(\tilde{Q}+\xi)\cdot\dot{\xi}|\\ &=|(\nabla^{2}E(\tilde{Q})\xi+O(|\xi|^{2}))\cdot\dot{\xi}|\\ &\lesssim|\nabla^{2}E(\tilde{Q})\xi\cdot\dot{\xi}|+|\xi|^{2}|\dot{\xi}|\\ &=|\Omega(\dot{\xi},A\xi)|+|\xi|^{2}|\dot{\xi}|\\ &=k|\dot{\lambda}_{+}\lambda_{-}+\lambda_{+}\dot{\lambda}_{-}|+|\xi|^{2}|\dot{\xi}|\\ &=k|\lambda_{-}N_{+}(X)+\lambda_{+}N_{-}(X)|+|\xi|^{2}|\dot{\xi}|\\ &\lesssim(|\lambda_{-}|+|\lambda_{+}|)^{3}+(|\lambda_{-}|+|\lambda_{+}|)^{2}(|\dot{\lambda}_{+}|+|\dot{\lambda}_{-}|)\\ &\lesssim(|\lambda_{-}|+|\lambda_{+}|)^{3}+(|\lambda_{-}|+|\lambda_{+}|)^{2}|\gamma(t)|^{2}.\end{split}

Hence

(71) |ddt(E(ψ(t))E{ξ+,ξ}(t))|(|λ|+|λ+|)3+(|λ|+|λ+|)2|γ(t)|2.\begin{split}|\frac{d}{dt}(E(\psi(t))-E_{\{\xi_{+},\xi_{-}\}}(t))|\lesssim(|\lambda_{-}|+|\lambda_{+}|)^{3}+(|\lambda_{-}|+|\lambda_{+}|)^{2}|\gamma(t)|^{2}.\end{split}

Moreover,

(72) E(ψ(t))E(Q~)E{ξ+,ξ}(t)=12Ω(γ,Aγ)+O(|X|3)+O(|ξ|3)|γ|2+O(|X|3)+O(|ξ|3),\begin{split}E(\psi(t))-E(\tilde{Q})-E_{\{\xi_{+},\xi_{-}\}}(t)&=\frac{1}{2}\Omega(\gamma,A\gamma)+O(|X|^{3})+O(|\xi|^{3})\\ &\sim|\gamma|^{2}+O(|X|^{3})+O(|\xi|^{3}),\end{split}

thus by (72) (71) and (68) we get the desired estimate in (ii) for |γ||\gamma|. The equation of λ2\lambda_{2} implies 𝔰λ2(t)R(ekt1)R2\mathfrak{s}\lambda_{2}(t)\gtrsim R(e^{kt}-1)-R^{2}, hence there is T1T_{*}\sim 1 such that 𝔰λ2R\mathfrak{s}\lambda_{2}\gtrsim R and ddtdQ(ψ)>0\frac{d}{dt}d_{Q}(\psi)>0 for tTRt\geq T_{*}R. For t[0,TR]t\in[0,T_{*}R], we have |ddtdQ(ψ)|R2|\frac{d}{dt}d_{Q}(\psi)|\lesssim R^{2} and so |dQ(ψ)R|R3|d_{Q}(\psi)-R|\lesssim R^{3}.

Finally, we expand KK around Q~\tilde{Q}:

(73) K(Q~+X)=K(Q~)X+O(|X|2).\begin{split}K(\tilde{Q}+X)=\nabla K(\tilde{Q})\cdot X+O(|X|^{2}).\end{split}

Let Q0:=(q0,0,0,q0)Q_{0}:=(-q_{0},0,0,q_{0}) then from (31) and (34) one can check that

(74) K(Q~)=((α2+4α+2)q0,0,0,2q0)=2E(Q~)Q0.\begin{split}\nabla K(\tilde{Q})=((\alpha^{2}+4\alpha+2)q_{0},0,0,2q_{0})=\nabla^{2}E(\tilde{Q})Q_{0}.\end{split}

Thus

(75) K(Q~+X)=2E(Q~)Q0X+O(|X|2)=Ω(Q0,AX)+O(|X|2)=kλ+Ω(Q0,ξ+)kλΩ(Q0,ξ)+Ω(Q0,Aγ)+O(|X|2).\begin{split}K(\tilde{Q}+X)&=\nabla^{2}E(\tilde{Q})Q_{0}\cdot X+O(|X|^{2})\\ &=\Omega(Q_{0},AX)+O(|X|^{2})\\ &=k\lambda_{+}\Omega(Q_{0},\xi_{+})-k\lambda_{-}\Omega(Q_{0},\xi_{-})+\Omega(Q_{0},A\gamma)+O(|X|^{2}).\end{split}

We can choose the normalized eigenvector ξ+,ξ\xi_{+},\xi_{-} (cf. (36)) so that Ω(Q0,ξ+)>0\Omega(Q_{0},\xi_{+})>0, then we obtain 𝔰K(ψ(t))(ektC)R\mathfrak{s}K(\psi(t))\gtrsim(e^{kt}-C_{*})R from (75) (68) and the estimates on |λ(t)|+|γ(t)||\lambda_{-}(t)|+|\gamma(t)| in (ii) as we have proved. The proof for WW is similar, and this time notice that

(76) W(Q~)=((α+2)2q0,0,0,0)=2E(Q~)Q~,\begin{split}\nabla W(\tilde{Q})=((\alpha+2)^{2}q_{0},0,0,0)=-\nabla^{2}E(\tilde{Q})\tilde{Q},\end{split}

and one can check that the previous chosen ξ+\xi_{+} will satisfy Ω(Q~,ξ+)>0-\Omega(\tilde{Q},\xi_{+})>0. ∎

Remark 2.

Note that for the ejection condition (61) to be satisfied, one can consider for instance λ(0)=0,γ(0)=0,λ+(0)=R0\lambda_{-}(0)=0,\gamma(0)=0,\lambda_{+}(0)=R_{0}, then differential equation of (λ+,λ,γ)(\lambda_{+},\lambda_{-},\gamma) shows that λ+(t)\lambda_{+}(t) grows exponentially.

Remark 3.

Note that the ejection Lemma also works backward in time, this time it’s the stable manifold, i.e. the eigenvalue k-k dominates. Since λ1=λ++λ2\lambda_{1}=\frac{\lambda_{+}+\lambda_{-}}{2}, we get the same statements as in the lemma for backward time.

Remark 4.

The ejection lemma holds for all α>0\alpha>0. However, to show the one-pass theorem we need some crucial variational estimates about the sign of KK, and that estimate only works for α2\alpha\geq 2.

5.5. The variational estimates

Let

dL(Γ):=min{dist((x,y),L1),dist((x,y),L2)}d_{L}(\Gamma):=\min\{\mathrm{dist}((x,y),L_{1}),\mathrm{dist}((x,y),L_{2})\}

be the distance in the configuration space with respect to the Lagrange points. The following lemma implies that the sign of W,KW,K cannot change outside of a neighborhood of L1,L2L_{1},L_{2} that is not too small; the size here depends on the amount by which the energy of the solution exceeds that of L1,L2L_{1},L_{2}.

Refer to caption
Figure 5. The black curve is the zero velocity curve for E(Γ)=E+cE(\Gamma)=E^{*}+c, i.e. V(x,y)=E+cV(x,y)=E^{*}+c. When c=ϵ(δ)2c=\epsilon(\delta)^{2} is small enough, the value of |W||W| is uniformly away from 0, provided dL(Γ)>δd_{L}(\Gamma)>\delta, i.e. outside the red dashed circles.
Lemma 16.

For the strong force α2\alpha\geq 2, for any δ>0\delta>0, there exist ϵ(δ),κ(δ)>0\epsilon(\delta),\kappa(\delta)>0 such that for any Γ4\Gamma\in\mathbb{R}^{4} satisfying

(77) E(Γ)<E+ϵ(δ)2,dL(Γ)δ,E(\Gamma)<E^{*}+\epsilon(\delta)^{2},\quad d_{L}(\Gamma)\geq\delta,

one has either

W(Γ)κ(δ)andK(Γ)κ(δ),W(\Gamma)\leq-\kappa(\delta)\quad\textrm{and}\quad K(\Gamma)\leq-\kappa(\delta),

or

W(Γ)κ(δ).W(\Gamma)\geq\kappa(\delta).
Proof.

First, notice that there exist ϵ(δ),κ(δ)>0\epsilon(\delta),\kappa(\delta)>0 so that if (77) is satisfied then |W(Γ)|κ(δ)|W(\Gamma)|\geq\kappa(\delta), see figure 5. We prove the first case by contradiction. Suppose there is δ0>0\delta_{0}>0 and a sequence Γn\Gamma_{n} with

dL(Γn)δ0,E(Γn)<E+1n,K(Γn)>1nd_{L}(\Gamma_{n})\geq\delta_{0},\quad E(\Gamma_{n})<E^{*}+\frac{1}{n},\quad K(\Gamma_{n})>-\frac{1}{n}

and there exists N0,κ0>0N_{0},\kappa_{0}>0 so that W(Γn)κ0W(\Gamma_{n})\leq-\kappa_{0} for all nN0n\geq N_{0}. By lemma 2, lemma 3 and remark 1, we know there is a regular point Γ\Gamma_{\infty} so that ΓnΓ\Gamma_{n}\to\Gamma_{\infty}. Thus

dL(Γ)δ0,E(Γ)E,K(Γ)0,W(Γ)κ0,d_{L}(\Gamma_{\infty})\geq\delta_{0},\quad E(\Gamma_{\infty})\leq E^{*},\quad K(\Gamma_{\infty})\geq 0,\quad W(\Gamma_{\infty})\leq-\kappa_{0},

which is impossible since inf{E|W0,K0}=E\inf\{E|W\leq 0,K\geq 0\}=E^{*} and is only achieved by ±Q\pm Q. ∎

The variational estimate in Lemma 16 uses the distance function in the configuration space (x,y)(x,y). We also have the variational estimate if we use the distance function dQ(ψ(t))d_{Q}(\psi(t)) in the phase space (q,p)(q,p).

Lemma 17.

For the strong force α2\alpha\geq 2, for any δ>0\delta>0, there exist ϵ(δ),κ(δ)>0\epsilon(\delta),\kappa(\delta)>0 such that for any ψ4\psi\in\mathbb{R}^{4} satisfying

(78) E(ψ)<E+ϵ(δ)2,dQ(ψ)δ,E(\psi)<E^{*}+\epsilon(\delta)^{2},\quad d_{Q}(\psi)\geq\delta,

one has either

W(ψ)κ(δ)andK(ψ)κ(δ),W(\psi)\leq-\kappa(\delta)\quad\textrm{and}\quad K(\psi)\leq-\kappa(\delta),

or

W(ψ)κ(δ).W(\psi)\geq\kappa(\delta).
Proof.

Use the fact that dQ(ψ(t))|X|=min(|ψ(t)Q~|,|ψ(t)+Q~|)d_{Q}(\psi(t))\sim|X|=\min(|\psi(t)-\tilde{Q}|,|\psi(t)+\tilde{Q}|) and an obvious modification of the proof as in Lemma 16. ∎

We remark that in the above lemma, when W>0W>0, the sign of KK can be both positive and negative with the condition (78). Nonetheless, Lemma 17 and the Ejection Lemma 15 still enable us to define a sign function away from ±Q~\pm\tilde{Q} by combining those of λ1\lambda_{1} and WW.

Lemma 18 (The sign function).

For the strong force α2\alpha\geq 2, let δS:=δX2C>0\delta_{S}:=\frac{\delta_{X}}{2C_{*}}>0 where δX\delta_{X} and C>1C_{*}>1 are constants from Lemma 15. Let 0<δδS0<\delta\leq\delta_{S} and

(79) (δ):={ψ4|E(ψ)<E+min(dQ2(ψ)/2,ϵ(δ)2)},\begin{split}\mathcal{H}_{(\delta)}:=\{\psi\in\mathbb{R}^{4}|E(\psi)<E^{*}+\min(d_{Q}^{2}(\psi)/2,\epsilon(\delta)^{2})\},\end{split}

where ϵ(δ)\epsilon(\delta) is given by Lemma 17. Then there exists a unique continuous function 𝔖:(δ){±1}\mathfrak{S}:\mathcal{H}_{(\delta)}\to\{\pm 1\} satisfying

(80) {ψ(δ),dQ(ψ)δE𝔖(ψ)=sign[λ1],ψ(δ),dQ(ψ)δ𝔖(ψ)=sign[W(ψ)],\begin{split}\begin{cases}\psi\in\mathcal{H}_{(\delta)},d_{Q}(\psi)\leq\delta_{E}&\Longrightarrow\quad\mathfrak{S}(\psi)=\mathrm{sign}[\lambda_{1}],\\ \psi\in\mathcal{H}_{(\delta)},d_{Q}(\psi)\geq\delta&\Longrightarrow\quad\mathfrak{S}(\psi)=\mathrm{sign}[W(\psi)],\end{cases}\end{split}

where we set sign[0]=+1\mathrm{sign}[0]=+1.

Proof.

The proof is the same as that in Nakanishi-Schlag ([10], Lemma 4.9). In particular, Lemma 14 implies that λ1\lambda_{1} is continuous for dQ(ψ)δEd_{Q}(\psi)\leq\delta_{E}, and Lemma 17 implies that sign[W(ψ)]\mathrm{sign}[W(\psi)] is continuous for dQ(ψ)δd_{Q}(\psi)\geq\delta. Thus it suffices to show that they coincide at dQ(ψ)=δS[δ,δX]d_{Q}(\psi)=\delta_{S}\in[\delta,\delta_{X}] in (δ)\mathcal{H}_{(\delta)}. Let ψ(t)\psi(t) be a solution of the HLP with ψ(0)(δ),dQ(ψ(0))=δS\psi(0)\in\mathcal{H}_{(\delta)},d_{Q}(\psi(0))=\delta_{S}, and such that the ejection condition (61) is satisfied. The ejection condition is easy to achieve, for instance see Remark 2. Then the Ejection Lemma implies that ψ(t)\psi(t) stays in (δ)\mathcal{H}_{(\delta)} and that sign[λ1(t)]\mathrm{sign}[\lambda_{1}(t)] is constant, until dQ(ψ(t))d_{Q}(\psi(t)) reaches δX\delta_{X}, at which time sign[W(ψ(t))]\mathrm{sign}[W(\psi(t))] is the same as sign[λ1(t)]\mathrm{sign}[\lambda_{1}(t)], because 2CδSδX2C_{*}\delta_{S}\leq\delta_{X}. Since sign[W(ψ)]\mathrm{sign}[W(\psi)] is constant for dQ(ψ)δd_{Q}(\psi)\geq\delta, we conclude that sign[W(ψ(t))]=sign[λ1(t)]\mathrm{sign}[W(\psi(t))]=\mathrm{sign}[\lambda_{1}(t)] from the beginning t=0t=0. ∎

5.6. The one-pass theorem

In this section we prove the crucial no-return property of the solutions for α2\alpha\geq 2. Indeed, thanks to Lemma 17 we are able to prove this for the case when 𝔖=1\mathfrak{S}=-1. When 𝔖=+1\mathfrak{S}=+1, we lose the control of the sign of KK, and the method from Nakanishi-Schlag seem to fail. We conjecture that the no-return property still holds for 𝔖=+1\mathfrak{S}=+1 and provide another direction about the proof of the no-return property for this case. We remark that the no-return property means that there are no homoclinic or heteroclinic orbits relative to ±Q~\pm\tilde{Q}. The following theorem implies that there are no heteroclinic orbits between the two equilibria, moreover, when a trajectory exits the 2ϵ2\epsilon ball of Q~\tilde{Q} or Q~-\tilde{Q} with 𝔖=1\mathfrak{S}=-1, i.e. ejecting into the region W<0W<0, then it will collide with the origin. This is very different from the Newtonian Hill’s Lunar problem, where an orbit enters the region W<0W<0 may still leave it through one of the bottle necks. See Figure 7 in [12].

Theorem 4 (One-pass theorem).

For the strong force α2\alpha\geq 2, there exist constants 0<ϵR10<\epsilon_{*}\ll R_{*}\ll 1 with the property: for any solution ψ\psi of the HLP (32) on the maximal interval [0,Tmax)[0,T_{\mathrm{max}}) satisfying

E(ψ)<E+ϵ2,dQ(ψ(0))<R,E(\psi)<E^{*}+\epsilon^{2},\quad d_{Q}(\psi(0))<R,

for some ϵ(0,ϵ]\epsilon\in(0,\epsilon_{*}], R(2ϵ,R]R\in(2\epsilon,R_{*}], and let

(81) Ttrap:=sup{t0|dQ(ψ(t))<R+R2,t[0,t)},\begin{split}T_{\mathrm{trap}}:=\sup\{t_{*}\geq 0|d_{Q}(\psi(t))<R+R^{2},\forall\,\,t\in[0,t_{*})\},\end{split}

then one has the following:

  1. (i)

    if Ttrap=TmaxT_{\mathrm{trap}}=T_{\mathrm{max}}, then ψ\psi is “trapped” at ±Q~\pm\tilde{Q} and Tmax=T_{\mathrm{max}}=\infty;

  2. (ii)

    if Ttrap<TmaxT_{\mathrm{trap}}<T_{\mathrm{max}}, and 𝔖(Ttrap)=1\mathfrak{S}(T_{\mathrm{trap}})=-1, then 𝔖(t)\mathfrak{S}(t) does not change on [Ttrap,Tmax)[T_{\mathrm{trap}},T_{\mathrm{max}}), and dQ(ψ(t))R+R2d_{Q}(\psi(t))\geq R+R^{2} for all Ttrapt<TmaxT_{\mathrm{trap}}\leq t<T_{\mathrm{max}}. Moreover, the solution has a finite time collision.

The proof relies on a virial type argument (I¨=K\ddot{I}=K) and is similar to that of PDE, cf. Theorem 4.1 in [11], Theorem 4.11 in [10] and Theorem 7.1 in [2]. In the paper, we follow closely with that in [11]. We choose R+R2R+R^{2} for the dichotomy because the distance function may exhibit oscillations on the order of O(R3)O(R^{3}), and so we need some room to ensure a true ejection from the small neighborhood. The most work of the proof is on the no-return statement, as the finite time collision then follows easily. Indeed, the finite time collision in the 𝔖=1\mathfrak{S}=-1 region follows from W<κ<0,W<-\kappa<0, hence K<κ<0K<-\kappa<0 and we can use the argument as in Proposition 2. Thus the one-pass theorem implies Theorem 3.

We now turn to the proof of the no-return statement. We may assume that ψ\psi does not stay very close to ±Q~\pm\tilde{Q} for all t>0t>0, so that we can apply the Ejection Lemma 15 at some time t=Ttrap>0t_{*}=T_{\mathrm{trap}}>0. The idea is to combine the hyperbolic structure of Lemma 15 close to ±Q~\pm\tilde{Q} with the variational structure in Lemma 17 away from ±Q~\pm\tilde{Q}, in order to control the virial identity through K(ψ)K(\psi). We choose δ>0\delta_{*}>0 as the distance threshold between the two regions in 4\mathbb{R}^{4}: for dQ(ψ)<δd_{Q}(\psi)<\delta_{*} we use the hyperbolic estimate in Lemma 15, and for dQ(ψ)>δd_{Q}(\psi)>\delta_{*} we use the variational estimate in Lemma 17. So δ,ϵ,R\delta_{*},\epsilon_{*},R_{*} should satisfy

(82) ϵRδδS,ϵϵ(δ).\begin{split}\epsilon_{*}\ll R_{*}\ll\delta_{*}\ll\delta_{S},\quad\epsilon_{*}\leq\epsilon(\delta_{*}).\end{split}

We shall impose further smallness conditions on δ,R,ϵ\delta_{*},R_{*},\epsilon_{*}. Afterward, RR_{*} and then ϵ\epsilon_{*} need to be made even smaller in order to satisfy the above conditions, depending on δ\delta_{*}.

Suppose towards a contradiction that ψ\psi solves the HLP (32) on [0,Tmax)[0,T_{\mathrm{max}}) in 4\mathbb{R}^{4} satisfying for some 0<T1<T2<T3<Tmax0<T_{1}<T_{2}<T_{3}<T_{\mathrm{max}} and all t(T1,T3)t\in(T_{1},T_{3}),

(83) dQ(ψ(0))<R=dQ(ψ(T1))=dQ(ψ(T3))<dQ(ψ(t)),dQ(ψ(T2))R+R2,\begin{split}d_{Q}(\psi(0))<R=d_{Q}(\psi(T_{1}))=d_{Q}(\psi(T_{3}))<d_{Q}(\psi(t)),\quad d_{Q}(\psi(T_{2}))\geq R+R^{2},\end{split}

as well as E(ψ)<E+ϵ2E(\psi)<E^{*}+\epsilon^{2} for some ϵ(0,ϵ]\epsilon\in(0,\epsilon_{*}] and R(2ϵ,R]R\in(2\epsilon,R_{*}].

Lemma 18 implies that 𝔰:=𝔖(ψ(t)){±1}\mathfrak{s}:=\mathfrak{S}(\psi(t))\in\{\pm 1\} is well-defined and constant on T1tT3T_{1}\leq t\leq T_{3}.

We apply the Ejection Lemma 15 first from t=T1t=T_{1} forward in time. Then by the lemma, there exists T1(T1,T1+TR)T_{1}^{\prime}\in(T_{1},T_{1}+T_{*}R) such that dQ(ψ(t))d_{Q}(\psi(t)) increases for t>T1t>T_{1}^{\prime} until it reaches δX\delta_{X}, and dQ(ψ(T1))=R+O(R3)<dQ(ψ(T2))δXd_{Q}(\psi(T_{1}^{\prime}))=R+O(R^{3})<d_{Q}(\psi(T_{2}))\ll\delta_{X}. Hence T1<T1<T2T_{1}<T_{1}^{\prime}<T_{2}, and by the lemma there is T1′′(T1,T3)T_{1}^{\prime\prime}\in(T_{1}^{\prime},T_{3}) such that dQ(ψ(t))d_{Q}(\psi(t)) increases exponentially on (T1,T1′′)(T_{1}^{\prime},T_{1}^{\prime\prime}), dQ(ψ(T1′′))=δXd_{Q}(\psi(T_{1}^{\prime\prime}))=\delta_{X} and on (T1,T1′′)(T_{1},T_{1}^{\prime\prime}),

(84) dQ(ψ(t))ek(tT1)R,𝔰K(ψ(t))(ek(tT1)C)R.\begin{split}d_{Q}(\psi(t))\sim e^{k(t-T_{1})}R,\quad\mathfrak{s}K(\psi(t))\gtrsim(e^{k(t-T_{1})}-C_{*})R.\end{split}

We can argue in the same way from t=T3t=T_{3} backward in time to obtain a time interval (T3′′,T3)(T1′′,T3)(T_{3}^{\prime\prime},T_{3})\subset(T_{1}^{\prime\prime},T_{3}), so that dQ(ψ(T3′′))=δXd_{Q}(\psi(T_{3}^{\prime\prime}))=\delta_{X},

(85) dQ(ψ(t))ek(T3t)R,𝔰K(ψ(t))(ek(T3t)C)R,t(T3′′,T3),\begin{split}d_{Q}(\psi(t))\sim e^{k(T_{3}-t)}R,\quad\mathfrak{s}K(\psi(t))\gtrsim(e^{k(T_{3}-t)}-C_{*})R,\quad t\in(T_{3}^{\prime\prime},T_{3}),\end{split}

and dQ(ψ(t))d_{Q}(\psi(t)) is decreasing in the region dQ(ψ(t))R+R2d_{Q}(\psi(t))\geq R+R^{2}. Moreover, from any τ(T1′′,T3′′)\tau\in(T_{1}^{\prime\prime},T_{3}^{\prime\prime}) where dQ(ψ(τ))<δd_{Q}(\psi(\tau))<\delta_{*} is a local minimum, we can apply the Ejection Lemma both forward and backward in time, thereby obtaining an open interval 𝐈τ(T1′′,T3′′)\mathbf{I}_{\tau}\subset(T_{1}^{\prime\prime},T_{3}^{\prime\prime}) so that dQ(ψ(𝐈τ))={δX}d_{Q}(\psi(\partial\mathbf{I}_{\tau}))=\{\delta_{X}\},

(86) dQ(ψ(t))ek|tτ|dQ(ψ(τ)),𝔰K(ψ(t))(ek|tτ|C)dQ(ψ(τ)),t𝐈τ,\begin{split}d_{Q}(\psi(t))\sim e^{k|t-\tau|}d_{Q}(\psi(\tau)),\quad\mathfrak{s}K(\psi(t))\gtrsim(e^{k|t-\tau|}-C_{*})d_{Q}(\psi(\tau)),\quad t\in\mathbf{I}_{\tau},\end{split}

and dQ(ψ(t))d_{Q}(\psi(t)) is monotone in the region dQ(ψ(t))2dQ(ψ(τ))d_{Q}(\psi(t))\geq 2d_{Q}(\psi(\tau)), which is the reason for 𝐈τ(T1′′,T3′′)\mathbf{I}_{\tau}\subset(T_{1}^{\prime\prime},T_{3}^{\prime\prime}). Moreover, the monotonicity away from τ\tau implies that any two intervals 𝐈τ1\mathbf{I}_{\tau_{1}} and 𝐈τ2\mathbf{I}_{\tau_{2}} for distinct local minimal points τ1\tau_{1} and τ2\tau_{2} are either disjoint or identical. Therefore, we have obtained disjoint open subintervals 𝐈1,,𝐈n(T1,T3)\mathbf{I}_{1},\cdots,\mathbf{I}_{n}\subset(T_{1},T_{3}) with n2n\geq 2, where we have either (84), (85) or (86) with τ=τj𝐈j\tau=\tau_{j}\in\mathbf{I}_{j}.

At the remaining times

(87) t𝐈:=(T1,T3)j=1n𝐈j,\begin{split}t\in\mathbf{I}^{\prime}:=(T_{1},T_{3})\setminus\bigcup_{j=1}^{n}\mathbf{I}_{j},\end{split}

we have dQ(ψ(t))δd_{Q}(\psi(t))\geq\delta_{*}, so that we can apply Lemma 17 to obtain for t𝐈t\in\mathbf{I}^{\prime},

(88) {W(ψ(t))κ(δ),K(ψ(t))κ(δ)(𝔰=1),W(ψ(t))κ(δ)(𝔰=+1).\begin{split}\begin{cases}W(\psi(t))\leq-\kappa(\delta_{*}),K(\psi(t))\leq-\kappa(\delta_{*})&(\mathfrak{s}=-1),\\ W(\psi(t))\geq\kappa(\delta_{*})&(\mathfrak{s}=+1).\end{cases}\end{split}

We shall call the 𝐈j\mathbf{I}_{j}’s the “hyperbolic” intervals, and 𝐈\mathbf{I}^{\prime} the “variational” intervals. See Figure 6.

ttdQd_{Q}δX\delta_{X}δ\delta_{*}R+R2R+R^{2}RR0T1T_{1}T1T_{1}^{\prime}T2T_{2}T1′′T_{1}^{\prime\prime}τ\tauT3′′T_{3}^{\prime\prime}T3T_{3}𝐈τ\mathbf{I}_{\tau}
Figure 6. Behavior of dQ(ψ(t))d_{Q}(\psi(t)) in (T1,T3)(T_{1},T_{3}) when there is a local minimum at τ\tau with dQ(ψ(τ))<δd_{Q}(\psi(\tau))<\delta_{*}. The red intervals indicate the “hyperbolic” region and the blue intervals indicate the “variational” region. The curve is smooth.

5.6.1. Virial estimate in the finite time collision case 𝔰=1\mathfrak{s}=-1

Notice that the virial identity is

(89) d2dt2I(ψ(t))=K(ψ(t)).\begin{split}\frac{d^{2}}{dt^{2}}I(\psi(t))=K(\psi(t)).\end{split}

Multiply (89) by 𝔰=1\mathfrak{s}=-1 and integrate over (T1,T3)(T_{1},T_{3}). Combining (84)-(88) we obtain

(90) |I˙(ψ(T3))I˙(ψ(T1))|j=1n𝐈j(ek|tτj|C)dQ(ψ(τj))𝑑t+𝐈κ(δ)𝑑tnδXδX.\begin{split}|\dot{I}(\psi(T_{3}))-\dot{I}(\psi(T_{1}))|&\gtrsim\sum_{j=1}^{n}\int_{\mathbf{I}_{j}}(e^{k|t-\tau_{j}|}-C_{*})d_{Q}(\psi(\tau_{j}))dt+\int_{\mathbf{I}^{\prime}}\kappa(\delta_{*})dt\\ &\gtrsim n\delta_{X}\geq\delta_{X}.\end{split}

On the other hand,

(91) I˙=xx˙+yy˙=x(px+y)+y(pyx)=xpx+ypy,\begin{split}\dot{I}=x\dot{x}+y\dot{y}=x(p_{x}+y)+y(p_{y}-x)=xp_{x}+yp_{y},\end{split}

thus I˙(±Q~)=0\dot{I}(\pm\tilde{Q})=0. At t=T1,T3t=T_{1},T_{3} we have dQ(ψ(t))=Rd_{Q}(\psi(t))=R, therefore

(92) |I˙(ψ(T3))I˙(ψ(T1))|RRδX.\begin{split}|\dot{I}(\psi(T_{3}))-\dot{I}(\psi(T_{1}))|\lesssim R\leq R_{*}\ll\delta_{X}.\end{split}

Comparing this bound with (90) leads to a contradiction. In conclusion, the solution ψ(t)\psi(t) cannot return to the RR-ball from the 𝔰=1\mathfrak{s}=-1 side.

5.6.2. No-return property for the case 𝔰=+1\mathfrak{s}=+1

In this case, we lose the sign definiteness of KK in the variational region thus the above virial type argument seems to fail for 𝔰=+1\mathfrak{s}=+1. However, we conjecture that the no-return property should still hold. This is somehow supported by the results in the Newtonian case. In [7], the authors showed that the unstable manifolds for the equilibria in the outer region (i.e. W>0W>0) goes forward to infinity for the Newtonian case. Their proof relies on the increasing of the argument θ\theta along the orbits for a suitable complex coordinate system for the general restricted three-body problem (R3BP), which is the main theorem of [8]. In particular, McGehee [8] showed that all orbits (with energy slightly above L2L_{2}) of the Newtonian R3BP leaving a vicinity of one of the equilibira proceed around an invariant annulus before returning to that vicinity. The Hill’s lunar problem is a limiting case of the R3BP, and that is how Llibre-Martínez-Simó (cf. section 4 [7]) showed the unstable manifolds in the outer region for L1,L2L_{1},L_{2} must go to infinity (no-return). The idea is roughly sketched in Figure 7.

Refer to caption
Refer to caption
Figure 7. Left: main technical result of McGehee [8] for the R3BP. Right: the unstable manifolds of the equilibria of the HLP in the outer region Llibre-Martínez-Simó (cf. section 4 [7])

Initially, we attempted to show the same result for the R3BP with α2\alpha\geq 2 as in [8], but it seems that this is not the case. Instead of going around an invariant annulus (such an invariant annlus may not exist for strong potential R3BP) in the configuration space, all solutions leaving a neighborhood of the L2L_{2} very likely will collide with one of the primary bodies in finite time. In particular, we conjecture that for the strong potential R3BP, all solutions with energy below the second Lagrange point (this energy is the threshold energy when one has a bounded component of the Hill’s region for the R3BP) should collide with one of the primary bodies in finite time if they started in the bounded component of its Hill’s region. Assuming this conjecture, then by passing to a limit similar to that of [7], we can show that the unstable manifolds of the equilibira in the outer region (W>0)(W>0) of the strong potential HLP go to infinity, hence completing the no-return property for 𝔰=+1\mathfrak{s}=+1.

We remark that characterizing the fates of the solutions for the strong potential R3BP itself is a very interesting problem, and it is a more complicated problem than the HLP. It is our hope that more attention could be put to this problem for a broader community.

acknowledgement

The authors are grateful to Takafumi Akahori and Kenji Nakanishi for helpful discussions. The first author is supported by Sun Yat-sen University start-up grant No. 74120-18841213. The second author is supported by the NSERC grant No. 371637-2019.

Appendix: Defining the Hill’s lunar problem

Following [9], we introduce scaled symplectic coordinates into the general three body problem with α\alpha-potential so that the Hill’s lunar equations are the equations of the first approximation. We remark that there is no essential difficulty in extending [9] to the α\alpha-potential, we include this section for completeness.

Consider a uniform rotating frame with frequency equal to one with reference to a fixed inertial frame and let x0,x1,x2;y0,y1,y22x_{0},x_{1},x_{2};y_{0},y_{1},y_{2}\in\mathbb{R}^{2} be the position and momentum vectors relative to the rotating frame of three particles of masses m0,m1,m2m_{0},m_{1},m_{2}. The Hamiltonian defining the equations of motion of the three particles with α\alpha-potential is

(93) H=i=02(|yi|22mixiTJyi)i<jmimj|xixj|α,H=\sum\limits_{i=0}^{2}(\frac{|y_{i}|^{2}}{2m_{i}}-x_{i}^{T}Jy_{i})-\sum\limits_{i<j}\frac{m_{i}m_{j}}{|x_{i}-x_{j}|^{\alpha}},

where J=(0110)J=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}. We shall refer to the particles of mass m0,m1m_{0},m_{1} and m2m_{2} as the earth, moon and sun, respectively. Since the center of mass is preserved and we want to scale the distances between the earth and moon, we use the Jacobi coordinates, which is a symplectic change of coordinates:

(94) u0=(m0+m1+m2)1(m0x0+m1x1+m2x2),u1=x1x0,u2=x2(m0+m1)1(m0x0+m1x1),v0=y0+y1+y2,v1=(m0+m1)1(m0y1m1y0),v2=(m0+m1+m2)1[(m0+m1)y2m2(y0+y1)],\begin{split}u_{0}&=(m_{0}+m_{1}+m_{2})^{-1}(m_{0}x_{0}+m_{1}x_{1}+m_{2}x_{2}),\\ u_{1}&=x_{1}-x_{0},\\ u_{2}&=x_{2}-(m_{0}+m_{1})^{-1}(m_{0}x_{0}+m_{1}x_{1}),\\ v_{0}&=y_{0}+y_{1}+y_{2},\\ v_{1}&=(m_{0}+m_{1})^{-1}(m_{0}y_{1}-m_{1}y_{0}),\\ v_{2}&=(m_{0}+m_{1}+m_{2})^{-1}[(m_{0}+m_{1})y_{2}-m_{2}(y_{0}+y_{1})],\end{split}

and obtain

(95) H=i=02(|vi|22MiuiTJvi)m0m1|u1|αm1m2|u2ν0u1|αm0m2|u2+ν1u1|α,H=\sum\limits_{i=0}^{2}(\frac{|v_{i}|^{2}}{2M_{i}^{\prime}}-u_{i}^{T}Jv_{i})-\frac{m_{0}m_{1}}{|u_{1}|^{\alpha}}-\frac{m_{1}m_{2}}{|u_{2}-\nu_{0}^{\prime}u_{1}|^{\alpha}}-\frac{m_{0}m_{2}}{|u_{2}+\nu_{1}^{\prime}u_{1}|^{\alpha}},

where

(96) M0=m0+m1+m2,M1=(m0+m1)1m0m1,M2=(m0+m1+m2)1(m0+m1)m2,ν0=(m0+m1)1m0,ν1=(m0+m1)1m1.\begin{split}&M_{0}^{\prime}=m_{0}+m_{1}+m_{2},\qquad M_{1}^{\prime}=(m_{0}+m_{1})^{-1}m_{0}m_{1},\\ &M_{2}^{\prime}=(m_{0}+m_{1}+m_{2})^{-1}(m_{0}+m_{1})m_{2},\\ &\nu_{0}^{\prime}=(m_{0}+m_{1})^{-1}m_{0},\qquad\nu_{1}^{\prime}=(m_{0}+m_{1})^{-1}m_{1}.\end{split}

Since HH is independent of u0u_{0} (the center of mass), its conjugate variable v0v_{0} (total linear momentum) is an integral. As usual, we take u0=v0=0u_{0}=v_{0}=0, thus we shall proceed with the Hamiltonian defined in (95) with the summation index from i=1i=1 to 2. We will make various assumptions on the size of various quantities until we are led to a definition of the equation of the first approximation for lunar theory. Following [9], we will use the common arrow notation in scaling problems.

The first assumption is that the earth and moon have approximately the same mass and are small compared to the mass of the sun, thus we let

(97) m0=ϵ2γμ0,m1=ϵ2γμ1,m2=μ2,m_{0}=\epsilon^{2\gamma}\mu_{0},\qquad m_{1}=\epsilon^{2\gamma}\mu_{1},\qquad m_{2}=\mu_{2},

where ϵ\epsilon is a small positive parameter and γ\gamma is a positive number to be chosen later. Since the masses of m0m_{0} and m1m_{1} are of order ϵ2γ\epsilon^{2\gamma}, so will be their momenta provided their velocities are of order 1. We make the substitutions v1ϵ2γv1,v2ϵ2γv2v_{1}\to\epsilon^{2\gamma}v_{1},v_{2}\to\epsilon^{2\gamma}v_{2} in (95). With this symplectic change of variables with multiplier ϵ2γ\epsilon^{2\gamma} the Hamiltonian becomes

(98) H=H1+H2+O(ϵ2γ),H1=|v1|22M1u1TJv1ϵ2γμ0μ1|u1|α,H2=|v2|22M2u2TJv2μ1μ2|u2ν0u1|αμ0μ2|u2+ν1u1|α,\begin{split}H&=H_{1}+H_{2}+O(\epsilon^{2\gamma}),\\ H_{1}&=\frac{|v_{1}|^{2}}{2M_{1}}-u_{1}^{T}Jv_{1}-\frac{\epsilon^{2\gamma}\mu_{0}\mu_{1}}{|u_{1}|^{\alpha}},\\ H_{2}&=\frac{|v_{2}|^{2}}{2M_{2}}-u_{2}^{T}Jv_{2}-\frac{\mu_{1}\mu_{2}}{|u_{2}-\nu_{0}u_{1}|^{\alpha}}-\frac{\mu_{0}\mu_{2}}{|u_{2}+\nu_{1}u_{1}|^{\alpha}},\end{split}

where

(99) M1=(μ0+μ1)1μ0μ1,M2=μ0+μ1,ν0=(μ0+μ1)1μ0,ν1=(μ0+μ1)1μ1.\begin{split}&M_{1}=(\mu_{0}+\mu_{1})^{-1}\mu_{0}\mu_{1},\qquad M_{2}=\mu_{0}+\mu_{1},\\ &\nu_{0}=(\mu_{0}+\mu_{1})^{-1}\mu_{0},\qquad\nu_{1}=(\mu_{0}+\mu_{1})^{-1}\mu_{1}.\end{split}

Note that the O(ϵ2γ)O(\epsilon^{2\gamma}) only depends on |v1||v_{1}| and |v2||v_{2}|.

The next assumption is that the distance between the earth and moon (|u1||u_{1}|) is small relative to the distance between the sun and the center of mass of the earth-moon system (|u2||u_{2}|). We make the change of variables u1ϵ2δu1u_{1}\to\epsilon^{2\delta}u_{1}, where δ\delta is a positive number to be chosen later. Following [9], we use the Legendre polynomials to expand the α\alpha-potential terms in H2H_{2}. In particular, if u,uu,u^{\prime} are two vectors with |u|>|u||u|>|u^{\prime}|, then

(100) 1|uu|=k=0|u|k|u|k+1Pk(cosθ),\frac{1}{|u-u^{\prime}|}=\sum\limits_{k=0}^{\infty}\frac{|u^{\prime}|^{k}}{|u|^{k+1}}P_{k}(\cos\theta),

where PkP_{k} is the kk-th order Legendre polynomial and θ\theta is the angle between u,uu,u^{\prime}. Use the Taylor series (1+x)α=1+αx+α(α1)2x2+(1+x)^{\alpha}=1+\alpha x+\frac{\alpha(\alpha-1)}{2}x^{2}+\cdots, and the assumption that |u2|>>|u1||u_{2}|>>|u_{1}| we have

(101) μ1μ2|u2ν0u1|α+μ0μ2|u2+ν1u1|α=μ2(μ0+μ1)|u2|α+μ2(μ1ν02+μ0ν12)|u1|2|u2|α+2[αP2(cosθ)+α(α1)2P12(cosθ)]+h.o.t.\begin{split}\frac{\mu_{1}\mu_{2}}{|u_{2}-\nu_{0}u_{1}|^{\alpha}}&+\frac{\mu_{0}\mu_{2}}{|u_{2}+\nu_{1}u_{1}|^{\alpha}}=\frac{\mu_{2}(\mu_{0}+\mu_{1})}{|u_{2}|^{\alpha}}+\\ &\frac{\mu_{2}(\mu_{1}\nu_{0}^{2}+\mu_{0}\nu_{1}^{2})|u_{1}|^{2}}{|u_{2}|^{\alpha+2}}[\alpha P_{2}(\cos\theta)+\frac{\alpha(\alpha-1)}{2}P_{1}^{2}(\cos\theta)]+h.o.t.\end{split}

where θ\theta is the angle between u1u_{1} and u2u_{2} and h.o.t.h.o.t. stands for higher order terms. Let

(102) H3=|v2|22M2u2TJv2μ2(μ0+μ1)|u2|α,H_{3}=\frac{|v_{2}|^{2}}{2M_{2}}-u_{2}^{T}Jv_{2}-\frac{\mu_{2}(\mu_{0}+\mu_{1})}{|u_{2}|^{\alpha}},

then H3H_{3} is the Hamiltonian of the Kepler problem, where a fixed body of mass μ2\mu_{2} is located at the origin and another body of mass μ0+μ1\mu_{0}+\mu_{1} moves in a rotating frame and is attracted to the fixed body by the α\alpha-potential. One can think of the fixed body as the sun and the other body as the union of the earth and moon.

The third and final assumption we shall make is that the center of mass of the earth-moon system moves on a nearly circular orbit about the sun. Since H3H_{3} is the Hamiltonian of a Kepler problem in rotating coordinates, one of the circular orbits becomes a circle of critical points for H3H_{3}. In particular, H3H_{3} has a critical point u2=a,v2=M2Jau_{2}=a,v_{2}=-M_{2}Ja for any constant vector aa satisfying |a|α+2=αμ2|a|^{\alpha+2}=\alpha\mu_{2}. Introduce coordinates,

Z=(u2v2)Z=\begin{pmatrix}u_{2}\\ v_{2}\end{pmatrix}

and a constant vector

Z0=(aM2Ja)Z_{0}=\begin{pmatrix}a\\ -M_{2}Ja\end{pmatrix}

so that H3H_{3} is a function of ZZ and H3(Z0)=0\nabla H_{3}(Z_{0})=0. By Taylor’s theorem

(103) H3(Z)=H3(Z0)+12(ZZ0)TS(ZZ0)+O(|ZZ0|3),H_{3}(Z)=H_{3}(Z_{0})+\frac{1}{2}(Z-Z_{0})^{T}S(Z-Z_{0})+O(|Z-Z_{0}|^{3}),

where SS is the Hessian of H3H_{3} evaluated at Z0Z_{0}. Since constants are lost in the formation of the equations of motion we shall ignore the constant H3(Z0)H_{3}(Z_{0}) in our further discussions. Thus we make the change of variables ZZ0ϵβXZ-Z_{0}\to\epsilon^{\beta}X, where β\beta is again a positive number to be chosen.

So far, starting with (98) we have proposed the following change of variables u1ϵ2δu1u_{1}\to\epsilon^{2\delta}u_{1} and ZZ0ϵβXZ-Z_{0}\to\epsilon^{\beta}X. In order to have a symplectic change of variables (of multiplier ϵ2β\epsilon^{2\beta}) we must make the further change v1ϵ2(βδ)v1v_{1}\to\epsilon^{2(\beta-\delta)}v_{1}. Thus we propose the following symplectic change of variables in (98):

(104) u1ϵ2δu1,v1ϵ2(βδ)v1,ZZ0ϵβX.u_{1}\to\epsilon^{2\delta}u_{1},\quad v_{1}\to\epsilon^{2(\beta-\delta)}v_{1},\quad Z-Z_{0}\to\epsilon^{\beta}X.

The first restriction is that the kinetic energy and potential energy in H1H_{1} should be of the same order of magnitude as mentioned in [9], and this leads to the restriction that 2β=(2α)δ+γ2\beta=(2-\alpha)\delta+\gamma. The second restriction following Hill is that the second term in (101) should be of the same order of magnitude as the terms in H1H_{1} and this leads to the condition 2δ=β2\delta=\beta. The smallest positive integer solutions for β\beta and δ\delta are δ=1\delta=1, β=2\beta=2 and γ=α+2\gamma=\alpha+2. With this choice of scale factors the Hamiltonian becomes

(105) H=|v1|22M1u1TJv1μ0μ1|u1|αμ2(μ1ν02+μ0ν12)|u1|2αμ2[αP2(cosθ)+α(α1)2P12(cosθ)]+12XTSX+O(ϵ2).\begin{split}H=\frac{|v_{1}|^{2}}{2M_{1}}-u_{1}^{T}Jv_{1}-\frac{\mu_{0}\mu_{1}}{|u_{1}|^{\alpha}}-\frac{\mu_{2}(\mu_{1}\nu_{0}^{2}+\mu_{0}\nu_{1}^{2})|u_{1}|^{2}}{\alpha\mu_{2}}&[\alpha P_{2}(\cos\theta)+\frac{\alpha(\alpha-1)}{2}P_{1}^{2}(\cos\theta)]\\ &+\frac{1}{2}X^{T}SX+O(\epsilon^{2}).\end{split}

In order to reduce the the number of constants in (105) we shall make one further scaling of the variables. We introduce new variables ξ\xi and η\eta to eliminate the subscripts and use the fact that P1(x)=xP_{1}(x)=x and P2(x)=12(3x21)P_{2}(x)=\frac{1}{2}(3x^{2}-1). Also we choose a=((αμ2)1α+2,0)a=((\alpha\mu_{2})^{\frac{1}{\alpha+2}},0) so that the abscissa points at the sun. Make the symplectic change of coordinates (with multiplier (μ0+μ1α+2)2α+2M1(\frac{\mu_{0}+\mu_{1}}{\alpha+2})^{\frac{2}{\alpha+2}}M_{1}):

(106) u1=(μ0+μ1α+2)1α+2ξ,v1=(μ0+μ1α+2)1α+2M1η,X=(μ0+μ1α+2)1α+2M11/2Yu_{1}=(\frac{\mu_{0}+\mu_{1}}{\alpha+2})^{\frac{1}{\alpha+2}}\xi,\quad v_{1}=(\frac{\mu_{0}+\mu_{1}}{\alpha+2})^{\frac{1}{\alpha+2}}M_{1}\eta,\quad X=(\frac{\mu_{0}+\mu_{1}}{\alpha+2})^{\frac{1}{\alpha+2}}M_{1}^{1/2}Y

so that (105) becomes

(107) H=|η|22ξTJηα+2|ξ|α12((α+2)ξ12|ξ|2)+12YTSY+O(ϵ2).H=\frac{|\eta|^{2}}{2}-\xi^{T}J\eta-\frac{\alpha+2}{|\xi|^{\alpha}}-\frac{1}{2}((\alpha+2)\xi_{1}^{2}-|\xi|^{2})+\frac{1}{2}Y^{T}SY+O(\epsilon^{2}).

Hill proposed to construct a lunar theory defined by the Hamiltonian

(108) H=|η|22ξTJηα+2|ξ|α12((α+2)ξ12|ξ|2)H^{\prime}=\frac{|\eta|^{2}}{2}-\xi^{T}J\eta-\frac{\alpha+2}{|\xi|^{\alpha}}-\frac{1}{2}((\alpha+2)\xi_{1}^{2}-|\xi|^{2})

and the equations defined by (108) are known as Hill’s lunar equations. Use the notation ξ=(x,y)\xi=(x,y) and η=(px,py)\eta=(p_{x},p_{y}) for the position and momentum variables, then the effective potential

V(x,y)=H12[(px+y)2+(pyx)2]=α+22x2α+2(x2+y2)α2.V(x,y)=H^{\prime}-\frac{1}{2}[(p_{x}+y)^{2}+(p_{y}-x)^{2}]=-\frac{\alpha+2}{2}x^{2}-\frac{\alpha+2}{(x^{2}+y^{2})^{\frac{\alpha}{2}}}.

References

  • [1] R. Abraham, J. Marsden: Foundations of Mechanics. Second Edition, Amer. Math. Soc. (2008)
  • [2] T. Akahori, S. Ibrahim, H. Kikuchi, H. Nawa: Global dynamics above the ground state energy for the combined power type nonlinear Schrodinger equations with energy critical growth at low frequencies. https://arxiv.org/abs/1510.08034v2(2019)
  • [3] Y. Deng, S. Ibrahim: Global Existence and Singularity of the N-body Problem with Strong Force. Qual. Theory Dyn. Syst. 19, 49. (2020)
  • [4] Y. Deng, S. Ibrahim: Numerical investigation on the Hill’s type lunar problem with homogeneous potential. arxiv.org/abs/1910.02512
  • [5] Y. Deng, S. Ibrahim, K. Nakanishi: Dynamics classification for the supercritical Hill problem. in preparation
  • [6] M.W. Hirsch, C.C. Pugh, M. Shub: Invariant manifolds. Springer-Verlag (1977)
  • [7] J. Llibre, R. Martínez, C. Simó: Transversality of the Invariant Manifolds Associated to the Lyapunov Family of Periodic Orbits near L2L_{2} in the Restricted Three-body Problem. J. Diff. Eq. 58, 104-156 (1985)
  • [8] R. P. McGehee: Some Homoclinic Orbits for the Restricted Three-body Problem. Ph.D. thesis, University of Wisconsin (1969)
  • [9] K. R. Meyer, D. S. Schmidt: Hill’s Lunar Equations and the Three-Body Problem. J. Diff. Eq. 44, 263-272 (1982)
  • [10] K. Nakanishi, W. Schlag: Invariant manifolds and dispersive Hamiltonian evolution equations. European Mathematics Society (2011)
  • [11] K. Nakanishi, W. Schlag: Global dynamics above the ground state energy for the cubic NLS equation in 3D. Calc. Var. and PDE 44, 1-45 (2012)
  • [12] H. Waalkens, A. Burbanks, S. Wiggins: Escape from planetary neighborhoods. Mon. Not. R. Astron. Soc. 361, 763-775 (2005)
  • [13] Z. Xia: Melnikov Method and Transversal Homoclinic Points in the Restricted Three-body Problem. Journal of Differential Equations 96, 170-184 (1992)