Next Article in Journal
Formal Developments for Lorentz-Violating Dirac Fermions and Neutrinos
Next Article in Special Issue
Kinematics Modeling and Analysis of Mid-Low Speed Maglev Vehicle with Screw and Product of Exponential Theory
Previous Article in Journal
On Some Iterative Numerical Methods for Mixed Volterra–Fredholm Integral Equations
Previous Article in Special Issue
RGB Inter-Channel Measures for Morphological Color Texture Characterization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Circular Hill Problem Regarding Arbitrary Disturbing Forces: The Periodic Solutions that are Emerging from the Equilibria

1
Departamento de Matemática Aplicada, Universidad de Salamanca, Casas del Parque, 2, 37008 Salamanca, Spain
2
Departamento de Matemáticas, Universidad de Castilla-La Mancha, Escuela Politécnica, Campus de Cuenca, 16071 Cuenca, Spain
*
Author to whom correspondence should be addressed.
Symmetry 2019, 11(10), 1196; https://doi.org/10.3390/sym11101196
Submission received: 6 September 2019 / Revised: 19 September 2019 / Accepted: 20 September 2019 / Published: 24 September 2019

Abstract

:
In this work, sufficient conditions for computing periodic solutions have been obtained in the circular Hill Problem with regard to arbitrary disturbing forces. This problem will be solved by means of using the averaging theory for dynamical systems as the main mathematical tool that has been applied in this work.
MSC:
37C27; 37J30; 37J25

1. Introduction

The three-body problem, as it is known, is a special case of the n—body problem where the motion of three point masses under their mutual gravitational interactions is described. This classical problem represents a large range of astronomical situations. Because a general solution of the problem exists but it is unusable in practice, some simplifications to it have been made. The Hill problem is an example of this. There are many studies on the Hill problem [1,2,3,4,5,6,7,8,9,10,11], and many modifications of this classic problem have appeared in the mathematical, astronomical, and mechanical literature. For example, in [6], Markellos and Roy introduced a photogravitational Hill problem as a simplified, non-trivial form of the photogravitational restricted three-body problem. In [8], Papadakis made a mainly numerical study of this problem on the planar case. On the other hand, in [2], the authors focused their attention on the case of a perturbed spatial Hill lunar problem using the averaging theory, and in [3], Chavineau and Mignard studied the trajectories of the Hill problem to describe the effect of solar perturbations on the relative motion of a binary asteroid. In a central force field, the binary objects case can be approached from potential [4].
Our main goal in this study was to find, using the averaging theory, sufficient conditions for the existence of periodic orbits that come from the Lagrangian points of the circular Hill problem with arbitrary disturbing forces.
Firstly, we assumed that the first two masses were small, compared to the third one (i.e., we assumed m 1 and m 2 were small as compared to M), and that the two small masses were close together, as compared to their separation from the third mass.
In the case of the circular Hill problem, the equations of the relative motion are [3,4]:
q ¨ 1 2 ω q ˙ 2 3 ω 2 q 1 + G m q 1 ( q 1 2 + q 2 2 + q 3 2 ) 3 / 2 = 0 , q ¨ 2 + 2 ω q ˙ 1 + G m q 2 ( q 1 2 + q 2 2 + q 3 2 ) 3 / 2 = 0 , q ¨ 3 + ω 2 q 3 + G m q 3 ( q 1 2 + q 2 2 + q 3 2 ) 3 / 2 = 0 ,
where G is the gravitational constant, m = m 1 + m 2 , and ω is the angular velocity of the motion of the barycenter of the m 1 - m 2 system around M.
The equilibrium solutions of the differential system (1) are given by
L 1 = G m 3 ω 2 1 / 3 , 0 , 0 , L 2 = G m 3 ω 2 1 / 3 , 0 , 0 .
In this work, we provide sufficient conditions for the existence of periodic solutions coming from the equilibrium point, L 1 . The same conditions are valid for the equilibrium, L 2 .
Making the change of variables x = q 1 G m 3 ω 2 1 / 3 , y = q 2 and z = q 3 , and linearizing, we can rewrite the unperturbed system of differential Equations (1) in the form:
x ¨ 9 ω 2 x 2 ω y ˙ = 0 , y ¨ + 2 ω x ˙ + 3 ω 2 y = 0 , z ¨ + 4 ω 2 z = 0 .
Using the averaging theory, we will provide a system of nonlinear equations whose simple zeros provide periodic solutions of the perturbed system:
x ¨ 9 ω 2 x 2 ω y ˙ = ε F 1 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) , y ¨ + 2 ω x ˙ + 3 ω 2 y = ε F 2 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) , z ¨ + 4 ω 2 z = ε F 3 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) ,
where the parameter ε is small, and F 1 , F 2 and F 3 are the smooth and periodic functions in the variable t in resonance p : q , ( p , q positive integers relatively prime) with some of the periodic solutions for ε = 0 .
The unperturbed system (2) has a unique singular point, the origin, with eigenvalues ± ω 1 i , ± ω 2 i , ± ω 3 with
ω 1 = 2 ω , ω 2 = ω 2 7 1 and ω 3 = ω 2 7 + 1 ,
and thus, the phase space x , x ˙ , y , y ˙ , z , z ˙ has two planes (excluding the origin) covered by periodic solutions, with periods
T 1 = 2 π ω 1 o r T 2 = 2 π ω 2 .
Accordingly, they belong to the plane associated to the eigenvectors with eigenvalues ± ω 1 i or ± ω 2 i , respectively. Now, we are going to study if and when the parameter ε is sufficiently small, and when the perturbed functions F i for i { 1 , 2 , 3 } have a period of either p T 1 / q or p T 2 / q , whether these solutions still exist for the perturbed system (3).
Let G = ( G 1 , G 2 ) be the Malkin bifurcation function (see [12]) for the system (3) and X 0 = ( X 1 0 , X 2 0 ) . Then:
G 1 ( X 0 ) = 1 p T 1 0 p T 1 cos ( ω 1 t ) F 1 * ( t ) d t , G 2 ( X 0 ) = 1 p T 1 0 p T 1 sin ( ω 1 t ) F 1 * ( t ) d t ,
where
F 1 * ( t ) = F 3
with F 3 = F 3 ( σ 1 1 ( t ) , , σ 6 1 ( t ) ) , and σ j 1 ( t ) = 0 , j = 1 , 4 ,
σ 5 1 ( t ) = X 1 sin ( 2 t ω ) X 2 cos ( 2 t ω ) 2 ω σ 6 1 ( t ) = X 1 cos ( 2 t ω ) + X 2 sin ( 2 t ω ) .
As usual, we are going to call a simple zero of the system G ( X 0 ) = 0 to any solution X 0 * = ( X 1 0 * , X 2 0 * ) , such that det G X 0 | X 0 = X 0 * 0 .
If X 0 * is a simple zero, then the T 1 -periodic solution of the unperturbed system with X 0 * as the initial value gives a T 1 -periodic solution of the perturbed system (see Appendix A).
Given an initial condition, we can use this result in order to find sufficient conditions for the existence of families of periodic orbits.
Theorem 1.
Suppose that the functions of the equations of motion of (3), F 1 , F 2 , and F 3 are smooth and periodic functions of the p T 1 / q period in the variable t, with p and q being positive integers that are relatively prime, and where the ratio of the frequencies ω 2 / ω 1 are not resonant with π. If ε 0 is sufficiently small, then for every simple zero that is not null, X 0 * , of the nonlinear system G ( X 0 ) = 0 , the system (3) has a periodic solution ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) such that ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) ( x ( t ) , y ( t ) , z ( t ) ) = ( σ 1 1 ( t ) , σ 3 1 ( t ) , σ 5 1 ( t ) ) | X 0 = X 0 * , when ϵ 0 —that is, when the periodic solution for the unperturbed system (2) travelled p times.
In Section 2, Theorem 1 will be proved using the averaging theory (see [13,14,15,16,17,18]).
In the next corollary, an application of Theorem 1 is shown, whose proof is implemented in Section 3.
Corollary 1.
Assume that f i ( t , x , x ˙ , y , y ˙ , z , z ˙ ) , i = 1 , 2 , 3 , are arbitrary, smooth functions, and
F 1 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = f 1 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) , F 2 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = f 2 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) , F 3 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = sin ( ω 1 t ) + ω z ˙ + f 3 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) x y z + 1 .
If ε 0 is sufficiently small, the perturbed system (3) has one periodic solution ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) such that, whenever ε 0 , ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) tends to the periodic solution of the unperturbed system ( x ( t ) , y ( t ) , z ( t ) ) = ( σ 1 1 ( t ) , σ 3 1 ( t ) , σ 5 1 ( t ) ) | X 0 = 0 , 4 3 ω 3 .
This corollary will be proved in Section 3.
Figure 1 shows the solution of the system (3) for the smooth function F 1 , F 2 and F 3 .
Let G ¯ = ( G 3 , G 4 ) be the Malkin bifurcation function (see [12]) for system (3) and Y 0 = ( Y 1 0 , Y 2 0 ) . Then:
G 3 ( Y 0 ) = 1 p T 2 0 p T 2 [ cos ( ω 2 t ) F 3 * ( t ) sin ( ω 2 t ) F 4 * ( t ) ] d t , G 4 ( Y 0 ) = 1 p T 2 0 p T 2 [ cos ( ω 2 t ) F 4 * ( t ) + sin ( ω 2 t ) F 3 * ( t ) ] d t ,
where
F 3 * ( t ) = 1 2 + 2 7 F 2 and F 4 * ( t ) = 1 2 2 7 1 7 F 1
with F i = F i ( σ 1 2 ( t ) , , σ 6 2 ( t ) ) , i = 1 , 2 , and
σ 1 2 ( t ) = 1 9 ω 4 7 Y 1 cos 2 7 1 ω t + Y 2 sin 2 7 1 ω t σ 2 2 ( t ) = 7 2 Y 2 cos 2 7 1 ω t Y 1 sin 2 7 1 ω t 2 7 1 σ 3 2 ( t ) = Y 2 cos 2 7 1 ω t Y 1 sin 2 7 1 ω t 2 7 1 ω σ 4 2 ( t ) = Y 1 cos 2 7 1 ω t + Y 2 sin 2 7 1 ω t
and σ j 2 ( t ) = 0 , j = 5 , 6 .
As has been stated before, if Y 0 * is a simple zero of G ¯ ( Y 0 * ) = 0 , then the T 2 -periodic solution of the unperturbed system with Y 0 * as the initial value gives a T 2 -periodic solution of the perturbed system (see Appendix A).
Theorem 2.
Suppose that the functions of the equations of motion of (3), F 1 , F 2 , and F 3 , are smooth and periodic functions of the p T 2 / q period in the variable t, with p and q positive integers relatively prime, and the ratio of the frequencies ω 2 / ω 1 not resonant with π. If ε 0 is sufficiently small, then for every simple and not null zero, Y 0 * of the nonlinear system G ¯ ( Y 0 ) = 0 , the system (3) has a periodic solution ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) such that ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) ( x ( t ) , y ( t ) , z ( t ) ) = ( σ 1 2 ( t ) , σ 3 2 ( t ) , σ 5 2 ( t ) ) | Y 0 = Y 0 * , when ϵ 0 —that is, when the periodic solution for the unperturbed system (2) travelled p times.
The proof of Theorem 2 is given in Section 2.
In the next corollary, an application of Theorem 2 is shown, whose proof is implemented in Section 3.
Corollary 2.
Let g ( t ) be the arbitrary and smooth function, and thus consider:
F 1 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = ( 1 + x z ˙ ) sin ( ω 2 t ) , F 2 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = x + y ˙ + 2 z , F 3 ( t , x , x ˙ , y , y ˙ , z , z ˙ ) = x 3 + y ˙ 2 + g ( t ) .
If ε 0 is sufficiently small, the perturbed system (3) has one periodic solution, ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) such that, whenever ε 0 , ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) tends to the periodic solution of the unperturbed system ( x ( t ) , y ( t ) , z ( t ) ) = ( σ 1 2 ( t ) , σ 3 2 ( t ) , σ 5 2 ( t ) ) | Y 0 = 14 7 7 w 7 7 + 4 7 w , 0 , ω 1 9 4 7 .
Corollary 2 will be proved in Section 3.
Figure 2 shows the solution of the system (3) for the smooth function, F 1 , F 2 and F 3 .

2. Proof of Theorems 1 and 3

The differential system (3) must be written as a first-order differential system defined in R 6 . For that, the variables ( x 1 , x 2 , y 1 , y 2 , z 1 , z 2 ) = ( x , x ˙ , y , y ˙ , z , z ˙ ) have been introduced in the following form, this being:
d x 1 d t = x 2 d x 2 d t = 9 ω 2 x 1 + 2 ω y 2 + ε F 1 ( x 1 , x 2 , y 1 , y 2 , z 1 , z 2 ) d y 1 d t = y 2 d y 2 d t = 2 ω x 2 3 ω 2 y 1 + ε F 2 ( x 1 , x 2 , y 1 , y 2 , z 1 , z 2 ) d z 1 d t = z 2 d z 2 d t = 4 ω 2 z 1 + ε F 3 ( x 1 , x 2 , y 1 , y 2 , z 1 , z 2 ) .
It must be remarked that if ε = 0 , the differential system (5) is equivalent to the differential system (2), well-known as the unperturbed system. Otherwise, when ε 0 , this is known as the perturbed system
The change of variables ( x ¯ 1 , x ¯ 2 , y ¯ 1 , y ¯ 2 , z ¯ 1 , z ¯ 2 ) ( x 1 , x 2 , y 1 , y 2 , z 1 , z 2 ) given by
x 1 x 2 y 1 y 2 z 1 z 2 = 0 0 7 4 9 w 0 4 + 7 9 w 4 + 7 9 w 0 0 0 2 7 1 + 2 7 2 + 7 1 + 2 7 2 + 7 1 + 2 7 0 0 0 1 1 + 2 7 w 1 1 + 2 7 w 1 1 + 2 7 w 0 0 1 0 1 1 0 1 2 w 0 0 0 0 1 0 0 0 0 0 x ¯ 1 x ¯ 2 y ¯ 1 y ¯ 2 z ¯ 1 z ¯ 2
writes the linear part of the differential system (5) in its real Jordan normal form, and this system in the new variables x ¯ = ( x ¯ 1 , x ¯ 2 , y ¯ 1 , y ¯ 2 , z ¯ 1 , z ¯ 2 ) becomes
x ¯ ˙ 1 = ω 1 x ¯ 2 + ε F 1 * x ¯ ˙ 2 = ω 1 x ¯ 1 + ε F 2 * y ¯ ˙ 1 = ω 2 y ¯ 2 + ε F 3 * y ¯ ˙ 2 = ω 2 y ¯ 1 + ε F 4 * z ¯ ˙ 1 = ω 3 z ¯ 1 + ε F 5 * z ¯ ˙ 2 = ω 3 z ¯ 2 + ε F 6 * ,
where
F 1 * = F 3 F 2 * = 0 F 3 * = 1 2 + 2 7 F 2 F 4 * = 1 2 2 7 1 7 F 1 F 5 * = 1 4 1 7 F 2 1 4 1 7 + 2 7 F 1 F 6 * = 1 4 1 7 F 2 + 1 4 1 7 + 2 7 F 1
with F i = F i ( σ 1 , σ 2 , σ 3 , σ 4 , σ 5 , σ 6 ) , and
σ 1 = 7 4 y ¯ 1 4 + 7 ( z ¯ 1 + z ¯ 2 ) 9 w σ 2 = 5 + 2 7 ( z ¯ 2 z ¯ 1 ) 2 7 5 y ¯ 2 3 σ 3 = 2 7 1 ( z ¯ 1 z ¯ 2 ) 1 + 2 7 y ¯ 2 3 3 w σ 4 = y ¯ 1 + z ¯ 1 + z ¯ 2 σ 5 = x ¯ 2 2 ω σ 6 = x ¯ 1 .
In order to prove Theorems 1 and 2, in the following lemma we are going to obtain the general expressions of the periodic orbits of the unperturbed system.
Lemma 1.
If ϵ = 0 , the T 1 —periodic solutions of the system of the differential Equations (7) are
x ¯ 1 ( t ) = X 1 0 cos ( ω 1 t ) + X 2 0 sin ( ω 1 t ) , x ¯ 2 ( t ) = X 2 0 cos ( ω 1 t ) X 1 0 sin ( ω 1 t ) , y ¯ 1 ( t ) = 0 , y ¯ 2 ( t ) = 0 , z ¯ 1 ( t ) = 0 , z ¯ 2 ( t ) = 0 ,
and the T 2 —periodic solutions are
x ¯ 1 ( t ) = 0 , x ¯ 2 ( t ) = 0 , y ¯ 1 ( t ) = Y 1 0 cos ( ω 2 t ) + Y 2 0 sin ( ω 2 t ) , y ¯ 2 ( t ) = Y 2 0 cos ( ω 2 t ) Y 1 0 sin ( ω 2 t ) , z ¯ 1 ( t ) = 0 , z ¯ 2 ( t ) = 0 .
Proof of Lemma 1.
If ε = 0 , the system (7) is a system of linear differential equations, and the proof follows easily. □
Proof of Theorem 1.
Since the positive integers p and q are relatively prime, and the functions F i , i = 1 , 2 , 3 , of the system (3) are p T 1 / q —periodic in the variable t, we can consider that the system (7) and the solutions (8) are p T 1 —periodic.
Now, we shall study what periodic solutions of the unperturbed system (7) with ε = 0 of the type (8) persist as periodic solutions, with the perturbed one for ε 0 being sufficiently small.
If we write the system (7) in the form
x ˙ ( t ) = G 0 ( t , x ¯ ) + ϵ G 1 ( t , x ¯ ) + ϵ 2 G 2 ( t , x ¯ , ϵ ) ,
where
x ¯ = x ¯ 1 x ¯ 2 y ¯ 1 y ¯ 2 z ¯ 1 z ¯ 2 , G 0 ( t , x ¯ ) = ω 1 x ¯ 2 ω 1 x ¯ 1 ω 2 y ¯ 2 ω 2 y ¯ 1 ω 3 z ¯ 1 ω 3 z ¯ 2 , G 1 ( t , x ¯ ) = F 1 * F 2 * F 3 * F 4 * F 5 * F 6 * , G 2 ( t , x ¯ , ε ) = 0 0 0 0 0 0 ,
we can apply Theorem A1 of the Appendix A. Our starting point is the description of the different elements which appear in this theorem for the particular case of the differential system (7). In this case, Ω = R 6 , k = 2 , and n = 6 . Now, let r 1 > 0 be arbitrarily small, and let r 2 > 0 be arbitrarily large. If V is the open and bounded subset of the plane x ¯ i = 0 , i = 3 , , 6 , of the form V = { ( X 1 0 , X 2 0 , 0 , 0 , 0 , 0 ) R 6 : r 1 < i = 1 2 ( X i 0 ) 2 < r 2 } . If α = ( X 1 0 , X 2 0 ) , then we identify V with the set { α R 2 : r 1 < | | α | | < r 2 } , where | | · | | is the Euclidean norm in R 2 . If we denote by Cl ( V ) , the closure of V, the function β : Cl ( V ) R 4 is β ( α ) = ( 0 , 0 , 0 , 0 ) . Using all these elements in our system, we have
Z = { z α = ( α , β ( α ) ) , α Cl ( V ) } = = { ( X 1 0 , X 2 0 , 0 , 0 , 0 , 0 ) R 6 : r 1 i = 1 2 ( X i 0 ) 2 r 2 }
Now, for each z α Z , we consider the p T 1 —periodic solution
x ( t , z α ) = ( X 1 ( t ) , X 2 ( t ) , 0 , 0 , 0 , 0 )
given by (8).
The fundamental matrix M z α ( t ) of the linear differential system (7) with ε = 0 , that is, associated to the p T 1 —periodic solution z α = ( X 1 0 , X 2 0 , 0 , 0 , 0 , 0 ) and such that M z α ( 0 ) = I d in R 6 , is
M z α ( t ) = M ( t ) =                                           
= cos ( ω 1 t ) sin ( ω 1 t ) 0 0 0 0 sin ( ω 1 t ) cos ( ω 1 t ) 0 0 0 0 0 0 cos ( ω 2 t ) sin ( ω 2 t ) 0 0 0 0 sin ( ω 2 t ) cos ( ω 2 t ) 0 0 0 0 0 0 e ω 3 t 0 0 0 0 0 0 e ω 3 t ,
where M ( t ) does not depend on the particular periodic solution x ( t , z α , 0 ) .
Computing the matrix M = M 1 ( 0 ) M 1 ( p T 1 ) , we get
M = 2 sin 2 ( p π ) sin ( 2 p π ) 0 0 0 0 sin ( 2 p π ) 2 sin 2 ( p π ) 0 0 0 0 0 0 2 sin 2 p π ω 2 ω 1 sin 2 p π ω 2 ω 1 0 0 0 0 sin 2 p π ω 2 ω 1 2 sin 2 p π ω 2 ω 1 0 0 0 0 0 0 1 e 2 p π ω 3 ω 1 0 0 0 0 0 0 1 e 2 p π ω 3 ω 1 ,
and M satisfies the assumptions of statement (ii) of Theorem A1, since the determinant
2 sin 2 p π ω 2 ω 1 sin 2 p π ω 2 ω 1 0 0 sin 2 p π ω 2 ω 1 2 sin 2 p π ω 2 ω 1 0 0 0 0 1 e 2 p π ω 3 ω 1 0 0 0 1 e 2 p π ω 3 ω 1 =
= 16 sinh 2 p π ω 3 ω 1 sin 2 p π ω 2 ω 1 0 ,
owing to the ratio of the frequencies, is non-resonant with π . In other words, all the assumptions of Theorem A1 are satisfied by the system (7).
In our particular system, if x ¯ = ( x ¯ 1 , x ¯ 2 , y ¯ 1 , y ¯ 2 , z ¯ 1 , z ¯ 2 ) , the map ξ : R 6 R 2 is given by ξ ( x ¯ ) = ( x ¯ 1 , x ¯ 2 ) , and if we calculate the function G ( X 1 0 , X 2 0 )
G ( X 1 0 , X 2 0 ) = G ( α ) = ξ 1 p T 1 0 p T 1 M z α 1 ( t ) G 1 ( t , x ¯ ( t , z α , 0 ) ) d t ,
we obtain G ( X 0 ) = ( G 1 ( X 0 ) , G 2 ( X 0 ) ) , where the functions G 1 , G 2 are given in (4). Then, by means of Theorem A1 of the Appendix A, for every simple zero X 0 * V of the system of nonlinear functions G ( X 0 ) = 0 , there is a periodic solution ( x ¯ 1 ( t , ε ) , x ¯ 2 ( t , ε ) , y ¯ 1 ( t , ε ) , y ¯ 2 ( t , ε ) , z ¯ 1 ( t , ε ) , z ¯ 2 ( t , ε ) ) of system (7), so that
( x ¯ 1 ( 0 , ε ) , x ¯ 2 ( 0 , ε ) , y ¯ 1 ( 0 , ε ) , y ¯ 2 ( 0 , ε ) , z ¯ 1 ( 0 , ε ) , z ¯ 2 ( 0 , ε ) ) ε 0 ( X 1 0 * , X 2 0 * , 0 , 0 , 0 , 0 ) .
Going back through the change of variables (6), the periodic solution ( x 1 ( t , ε ) , x 2 ( t , ε ) , y 1 ( t , ε ) , y 2 ( t , ε ) , z 1 ( t , ε ) , z 2 ( t , ε ) ) of system (7) has been obtained, so that
x 1 ( t , ε ) x 2 ( t , ε ) y 1 ( t , ε ) y 2 ( t , ε ) z 1 ( t , ε ) z 2 ( t , ε ) 0 0 0 0 X 1 sin ( 2 t ω ) X 2 cos ( 2 t ω ) 2 ω X 1 cos ( 2 t ω ) + X 2 sin ( 2 t ω ) when ε 0 .
Consequently, a periodic solution ( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) of system (3) has been achieved, so that
( x ( t , ε ) , y ( t , ε ) , z ( t , ε ) ) 0 0 X 1 sin ( 2 t ω ) X 2 cos ( 2 t ω ) 2 ω when ε 0 .
This completes the proof of Theorem 1. □
Proof of Theorem 2.
The proof is analogous to the proof of Theorem 1 changing the roles of T 1 for T 2 , and we obtain the periodic solution
4 7 Y 1 cos 2 7 1 t ω + Y 2 sin 2 7 1 t ω 9 ω 7 2 Y 2 cos 2 7 1 t ω Y 1 sin 2 7 1 t ω 2 7 1 Y 2 cos 2 7 1 t ω Y 1 sin 2 7 1 t ω 2 7 1 ω Y 1 cos 2 7 1 t ω + Y 2 sin 2 7 1 t ω 0 0 when ε 0 .
In consequence, a periodic solution ( x , y , z ) ( t , ε ) of system (3) has been obtained, so that
( x , y , z ) ( t , ε ) 4 7 Y 1 cos 2 7 1 t ω + Y 2 sin 2 7 1 t ω 9 ω Y 2 cos 2 7 1 t ω Y 1 sin 2 7 1 t ω 2 7 1 ω 0 when ε 0 .

3. Proof of Corollaries 1 and 2

Proof of Corollary 1.
Under the assumptions of Corollary 1, we have the function
G 1 ( X 0 ) = 1 p T 1 0 p T 1 cos ( ω 1 t ) F 1 * ( t ) d t = 3 8 X 1 0 ( X 1 0 ) 2 + ( X 2 0 ) 2 ω , G 2 ( X 0 ) = 1 p T 1 0 p T 1 sin ( ω 1 t ) F 1 * ( t ) d t = 1 8 [ 4 + 3 ( X 1 0 ) 2 X 2 0 ω + 3 ( X 2 0 ) 3 ω ]
and
det G X 0 X 0 * = ( X 1 0 , X 2 0 ) = 3 8 9 ω 2 2 3 0 , X 1 0 , X 2 0 R , ω 0
Therefore, the solution X 0 * = 0 , 4 3 ω 3 of the nonlinear system G ( X 0 ) = 0 is simple, and by Theorem 1 we only have one periodic solution for the system of this corollary. □
Proof of Corollary 2.
Under the assumptions of Corollary 2, we have the function
G 3 ( Y 0 ) = 1 p T 2 0 p T 2 cos ( ω 2 t ) F 3 * ( t ) sin ( ω 2 t ) F 4 * ( t ) d t = = w 7 + 4 7 Y 1 0 + 14 7 7 7 Y 1 0 28 w G 4 ( Y 0 ) = 1 p T 2 0 p T 2 cos ( ω 2 t ) F 4 * ( t ) + sin ( ω 2 t ) F 3 * ( t ) d t = = 1 28 Y 2 0 7 w + 4 7 + 7 ,
and, if ω 1 9 ( 4 7 ) ,
det G ¯ Y 0 Y 0 * = ( Y 1 0 , Y 2 0 ) = 1 112 1 w 2 2 4 + 7 w + 8 7 + 23 0 .
Therefore, the solution Y 0 * = 14 7 7 w 7 7 + 4 7 w , 0 of the nonlinear system G ¯ ( Y 0 ) = 0 is simple. By means of Theorem 1, the system of this Corollary has only this periodic solution. □

Author Contributions

All the authors have contributed in an equal way to the research in this paper.

Funding

This work has been partially supported by FEDER OP2014-2020 of Castilla-La Mancha (Spain) grant number 2019-GRIN-27168, Fundación Séneca de la Región de Murcia (Spain) grant number 20783/PI/18 and the grant of Ministerio de Ciencia, Innovación y Universidades from Spain reference PGC2018-097198-B-I00.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Basic Results on Averaging Theory

In this appendix we present the basic result from the averaging theory that we shall need for proving the main results of this paper.
We consider the problem of the bifurcation of T —periodic solutions from a differential system of the form
x ˙ ( t ) = G 0 ( t , x ) + ϵ G 1 ( t , x ) + ϵ 2 G 2 ( t , x , ϵ ) ,
with ϵ 0 sufficiently small. Here the functions G 0 , G 1 : R × Ω R n and G 2 : R × Ω × ( ϵ 0 , ϵ 0 ) R n are C 2 functions, T —periodic in the first variable, and Ω is an open subset of R n . The main assumption is that the unperturbed system
x ˙ ( t ) = G 0 ( t , x ) ,
has a submanifold of periodic solutions. A solution of this problem is given using the averaging theory.
Malkin (see [13] and references therein) studied the bifurcation of T–periodic solutions in the T–periodic system x ˙ = G 0 ( t , x ) + ε G 1 ( t , x , ε ) , whose unperturbed system has a family of T–periodic solutions with initial conditions given by a smooth function β : R k R n , and proved that if the bifurcation function
G ( α ) = 0 T < u 1 ( t , α ) , g ( t , x ( t , β ( α ) ) , 0 ) > < u k ( t , α ) , g ( t , x ( t , β ( α ) ) , 0 ) > d t ,
where u i , i = 1 , , k , are k linearly independent T-periodic solutions of the adjoint linearized differential system, has a simple zero α 0 such that det ( G ) | α = α 0 0 , then for any ε > 0 sufficiently small, the system x ˙ = G 0 ( t , x ) + ε G 1 ( t , x , ε ) has an unique T–periodic solution x ε such that x ε ( 0 ) β ( α 0 ) as ε 0 .
This can be rephrased as follows: Let x ( t , z , ϵ ) be the solution of the system (A2) such that x ( 0 , z , ϵ ) = z . We write the linearization of the unperturbed system along a periodic solution x ( t , z , 0 ) as
y ˙ = D x G 0 ( t , x ( t , z , 0 ) ) y .
In what follows we denote by M z ( t ) some fundamental matrix of the linear differential system (A3), and by ξ : R k × R n k R k the projection of R n onto its first k coordinates; i.e., ξ ( x 1 , , x n ) = ( x 1 , , x k ) .
We assume that there exists a k-dimensional submanifold Z of Ω filled with T —periodic solutions of (A2). Then an answer to the problem of bifurcation of T —periodic solutions from the periodic solutions contained in Z for system (A1) is given in the following result.
Theorem A1.
Let V be an open and bounded subset of R k , and let β : Cl ( V ) R n k be a C 2 function. We assume that
(i) 
Z = { z α = ( α , β ( α ) ) , α Cl ( V ) } Ω and that for each z α Z the solution x ( t , z α ) of (A2) is T-periodic;
(ii) 
for each z α Z there is a fundamental matrix M z α ( t ) of (A3) such that the matrix M z α 1 ( 0 ) M z α 1 ( T ) has in the upper right corner the k × ( n k ) zero matrix, and in the lower right corner a ( n k ) × ( n k ) matrix Δ α with det ( Δ α ) 0 .
We consider the function G : Cl ( V ) R k
G ( α ) = ξ ( 1 T 0 T M z α 1 ( t ) G 1 ( t , x ( t , z α , 0 ) ) d t ) .
If there exists a V with G ( a ) = 0 and det ( ( d G / d α ) ( a ) ) 0 , then there is a T —periodic solution x ( t , ϵ ) of system (A1) such that x ( 0 , ϵ ) z a as ϵ 0 .
For a proof of Theorem A1 see [12,19,20] for shorter proof.

References

  1. Abouelmagd, E.I.; Guirao, J.L.G. On the perturbed restricted three-body problem. Appl. Math. Nonlinear Sci. 2016, 1, 123–144. [Google Scholar] [CrossRef] [Green Version]
  2. de Bustos, M.T.; Guirao, J.L.G.; Vera, J.A. The spatial Hill Lunar problem: Periodic solutions emerging from equilibria. Dynam. Syst. 2017, 32, 340–353. [Google Scholar] [CrossRef]
  3. Chauvineau, B.; Mignard, F. Dynamics of Binary Asteroids. I. Hill’s Case. Icarus 1990, 83, 360–381. [Google Scholar] [CrossRef]
  4. Chauvineau, B. The generalized Hill problem. Case of an external field of force deriving from a central potential. Celest. Mech. Dyn. Astron. 1991, 51, 119–129. [Google Scholar] [CrossRef]
  5. Llibre, J. On the central configurations of the n-body problem. Appl. Math. Nonlinear Sci. 2017, 2, 509–518. [Google Scholar] [CrossRef]
  6. Markellos, V.V.; Roy, A.E.; Velgakis, M.J.; Kanavos, S.S. A Photogravitational Hill Problem and radiation effects on Hill Stability of Orbits. Astrophys. Space Sci. 2000, 271, 293–301. [Google Scholar] [CrossRef]
  7. Michalodimitrakis, M. Hill’s problem: Families of three dimentional periodic orbits (part I). Astrophys. Space Sci. 1980, 68, 253–268. [Google Scholar] [CrossRef]
  8. Papadakis, K.E. The Planar Photogravitational Hill Problem. Int. J. Bifur. Chaos Appl. Sci. Engrg. 2006, 16, 1809–1821. [Google Scholar] [CrossRef]
  9. Perdiou, A.E.; Markellos, V.V.; Douskos, C.N. The Hill Problem with oblate secondary: Numerical exploration. Earth Moon Planets 2005, 97, 127–145. [Google Scholar] [CrossRef]
  10. Pérez-Chavela, E.; Tamayo, C. Relative Equilibria in the 4-Vortex Problem Bifurcating from an Equilateral Triangle Configuration. Appl. Math. Nonlinear Sci. 2016, 1, 301–310. [Google Scholar] [CrossRef] [Green Version]
  11. Simó, C.; Stuchi, T.J. Central stable/unstable manifolds and the destruction of KAM tori in the planar Hill problem. Physics D 2000, 140, 1–32. [Google Scholar] [CrossRef]
  12. Malkin, I.G. Some Problems of the Theory of Nonlinear Oscillations; Gosudarstv. Izdat. Tehn.–Teor. Lit.: Moscow, Russia, 1956. [Google Scholar]
  13. Buica, A.; García, I. Periodic solutions of some perturbed symmetric Euler top. Topol. Methods Nonlinear Anal. 2010, 36, 91–100. [Google Scholar]
  14. de Bustos, M.T.; Guirao, J.L.G.; Vera, J.A.; Vigo-Aguiar, J. Periodic orbits and C1-integrability in the planar Stark-Zeeman problem. J. Math. Phys. 2012, 53, 082701. [Google Scholar] [CrossRef]
  15. de Bustos, M.T.; Guirao, J.L.G.; Llibre, J.; Vera, J.A. New families of periodic orbits for a galactic potential. Chaos Solitons Fractals 2016, 82, 97–102. [Google Scholar] [CrossRef]
  16. de Bustos, M.T.; López, M.A.; Martínez, R. On the periodic auto-oscillations of an electric circuit with periodic imperfections on its variables. Appl. Math. Inf. Sci. 2013, 7, 2105–2111. [Google Scholar] [CrossRef]
  17. de Bustos, M.T.; López, M.A.; Martínez, R. On the periodic solutions of a linear chain of three identical atoms. Nonlinear Dyn. 2014, 76, 893–903. [Google Scholar] [CrossRef]
  18. Llibre, J.; Rodrigues, A. On the periodic orbits of Hamiltonian systems. J. Math. Phys. 2010, 51, 042704. [Google Scholar] [CrossRef]
  19. Roseau, M. Vibrations Non Linéaires et Théorie de la Stabilité; Springer Tracts in Natural Philosophy; Springer: Berlin, Germany, 1966; Volume 8. [Google Scholar]
  20. Buica, A.; Francoise, J.P.; Llibre, J. Periodic solutions of nonlinear periodic differential systems with a small parameter. Commun. Pure Appl. Anal. 2007, 6, 103–111. [Google Scholar]
Figure 1. Solutions of the differential system (3) for the smooth functions F 1 = 1 x x ˙ 2 sin ( ω 1 t ) y , F 2 = sin ( ω 1 t ) x ˙ 2 y z + z 2 , and F 3 = sin ( ω 1 t ) + ω z ˙ + x ˙ x y z + 1 for Corollary 1 with ε > 0 , w = 1 , ω 1 = 2 ω , t = [ 0 , 2 π / ω 1 ] , and X 0 * = 0 , 4 3 ω 3 .
Figure 1. Solutions of the differential system (3) for the smooth functions F 1 = 1 x x ˙ 2 sin ( ω 1 t ) y , F 2 = sin ( ω 1 t ) x ˙ 2 y z + z 2 , and F 3 = sin ( ω 1 t ) + ω z ˙ + x ˙ x y z + 1 for Corollary 1 with ε > 0 , w = 1 , ω 1 = 2 ω , t = [ 0 , 2 π / ω 1 ] , and X 0 * = 0 , 4 3 ω 3 .
Symmetry 11 01196 g001
Figure 2. Solutions of the differential system (3) for the smooth functions F 1 = ( 1 + x z ˙ ) sin ( ω 2 t ) , F 2 = x + y ˙ + 2 z and F 3 = x 3 + y ˙ 2 + t for Corollary 2 with ε > 0 , w = 1 , ω 2 = ω 2 7 1 , t = [ 0 , 2 π / ω 2 ] and Y 0 * = 14 7 7 w 7 7 + 4 7 w , 0 .
Figure 2. Solutions of the differential system (3) for the smooth functions F 1 = ( 1 + x z ˙ ) sin ( ω 2 t ) , F 2 = x + y ˙ + 2 z and F 3 = x 3 + y ˙ 2 + t for Corollary 2 with ε > 0 , w = 1 , ω 2 = ω 2 7 1 , t = [ 0 , 2 π / ω 2 ] and Y 0 * = 14 7 7 w 7 7 + 4 7 w , 0 .
Symmetry 11 01196 g002

Share and Cite

MDPI and ACS Style

de Bustos, M.T.; López, M.A.; Martínez, R. The Circular Hill Problem Regarding Arbitrary Disturbing Forces: The Periodic Solutions that are Emerging from the Equilibria. Symmetry 2019, 11, 1196. https://doi.org/10.3390/sym11101196

AMA Style

de Bustos MT, López MA, Martínez R. The Circular Hill Problem Regarding Arbitrary Disturbing Forces: The Periodic Solutions that are Emerging from the Equilibria. Symmetry. 2019; 11(10):1196. https://doi.org/10.3390/sym11101196

Chicago/Turabian Style

de Bustos, M. Teresa, Miguel A. López, and Raquel Martínez. 2019. "The Circular Hill Problem Regarding Arbitrary Disturbing Forces: The Periodic Solutions that are Emerging from the Equilibria" Symmetry 11, no. 10: 1196. https://doi.org/10.3390/sym11101196

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop