Next Article in Journal
An Exploratory Approach to the Adoption Process of Bitcoin by Business Executives
Next Article in Special Issue
Stochastic Computing Implementation of Chaotic Systems
Previous Article in Journal
Dynamics of Rössler Prototype-4 System: Analytical and Numerical Investigation
Previous Article in Special Issue
The Topological Entropy Conjecture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Zero-Hopf Bifurcation in a Generalized Genesio Differential Equation

1
Zouhair Diab Department of Mathematics and Computer Science, Larbi Tebessi University, Tebessa 12002, Algeria
2
Departamento de Matemáca Aplicada y Estadística, Universidad Politécnica de Cartagena, 30202 Cartagena, Región de Murcia, Spain
3
Centro Universitario de la Defensa, Academia General del Aire, Universidad Politécnica de Cartagena, 30720 Santiago de la Ribera, Región de Murcia, Spain
*
Author to whom correspondence should be addressed.
Mathematics 2021, 9(4), 354; https://doi.org/10.3390/math9040354
Submission received: 21 December 2020 / Revised: 30 January 2021 / Accepted: 2 February 2021 / Published: 10 February 2021
(This article belongs to the Special Issue Dynamical Systems and Their Applications Methods)

Abstract

:
The purpose of the present paper is to study the presence of bifurcations of zero-Hopf type at a generalized Genesio differential equation. More precisely, by transforming such differential equation in a first-order differential system in the three-dimensional space R 3 , we are able to prove the existence of a zero-Hopf bifurcation from which periodic trajectories appear close to the equilibrium point located at the origin when the parameters a and c are zero and b is positive.

1. Introduction and Statement of the Main Results

In [1], the authors analyzed the presence of zero-Hopf bifurcations for the ordinary differential equation of order three
x + a x ¨ + b x ˙ + c x x 2 = 0 ,
which is commonly known as the Genesio equation.
In this work, our aim is to study the existence of bifurcations of zero-Hopf type at the so-called generalized Genesio differential equation
x + a x ¨ + b x ˙ + c x + h ( x ) = 0 ,
where h ( x ) = k = 1 n ( 1 ) k x 2 k , where n is any natural number, and a , b , c R . Note that the map h ( x ) is defined in the maximum domain where the solutions of (2) are defined. Note that for n = 1 , we have the Genesio equation; for n = 2 , we have the simplest chaotic jerk equation and so on. We underline that the results stated in this paper can be mimetically reproduced for a smooth and even h ( x ) .
By defining of the variables y = x ˙ and z = y ˙ , the differential Equation (2) can be written as the system of nonlinear differential equations
x ˙ = y , y ˙ = z , z ˙ = c x b y a z k = 1 n ( 1 ) k x 2 k ,
Our endeavor is is to analyze the presence of zero-Hopf type equilibria and the existence of zero-Hopf bifurcations in the system (3).
Definition 1.
An zero-Hopf type equilibrium for a three-dimensional autonomous differential system of first-order is an isolated equilibrium point of such a system for which its linear part has eigenvalues one equal to zero and two purely imaginary, see [2] for more details.
The main approach to prove the presence of a zero-Hopf bifurcation is to transform the system object of our study into the normal form of a zero-Hopf bifurcation. However, the treatment that we have chosen for the system (3) is via the use of the averaging theory of dynamical systems. In Section 2, a summary of the main results which play a key role in our study are stated. The averaging method has already been used to study-Hopf and zero-Hopf bifurcations for other differential systems, see for instance [3,4,5,6]. Our main results are as follows:
Proposition 1.
The differential system (3) has a unique equilibrium of zero-Hopf type located at the origin when a = c = 0 and b > 0 .
Theorem 1.
Let us consider for the system (3), the following distribution of parameters a = ε α , b = ω 2 + ε β and c = ε γ , being ω > 0 and ε a sufficiently small parameter. Then, (3) has a zero-Hopf type bifurcation at the origin of coordinates in its equilibrium point for ε = 0 . In the case γ 2 α 2 ω 4 > 0 , (3) presents a unique periodic trajectory bifurcating from the origin. When γ = α ω 2 , (3) has at most 6 periodic orbits bifurcating from the origin.

2. Preliminaries

2.1. Results from Averaging Theory

Here, we present the main results on the second-order averaging theory of dynamical systems that will play a key role in the proof of our main results Theorem 1. For more information on this interesting theory and its application, see for instance [7] or [8] and references therein. For a proof of Theorem 2 that we are going to state, see Theorem 3.5.1 of Sanders and Verhulst [2], or [9] for a formulation in modern terminology.
Consider the differential system
x ˙ = ε F 1 ( t , x ) + ε 2 F 2 ( t , x ) + ε 3 Q ( t , x , ε ) , x 0 = x 0 ,
where F 1 , F 2 : R × Ω R n and Q : [ 0 , ] × Ω × 0 , ε 0 R n where Ω is a subset of R n open and F 1 , F 2 and Q are maps periodic of period T in the first variable. We set
F 10 ( x ) = 1 T 0 T F 1 ( t , x ) d t , F 20 ( x ) = 1 T 0 T D x F 1 ( t , x ) · y 1 ( t , x ) + F 2 ( t , x ) d t ,
where
y 1 ( t , x ) = 0 t F 1 ( s , x ) d s ,
Theorem 2.
In the previous conditions, assume that
(i) 
F 1 , F 2 , F 1 / x are locally Lipschitz in x and Q is twice differentiable with respect to ε .
(ii) 
Let V Ω be a bounded and open set, for each ε ε f , ε f \ { 0 } , there exists a ε V such that F 10 a ε + ε F 20 a ε = 0 and
d B F 10 + ε F 20 , V , a ε 0 .
Thus, for sufficiently small ε > 0 there exists a T-periodic solution φ ( · , ε ) holding φ ( 0 , ε ) = a ε .
Remark 1.
When we write d B F 10 + ε F 20 , V , a ε 0 means that the value of the Brouwer degree of the function F 10 : V R n at the fixed point a ε is not zero. A sufficient condition for having the previous property is that the Jacobian of the function F 10 + ε F 20 at a ε be non-zero. When F 10 0 , then the zeros of F 10 + ε F 20 are equivalent to the zeros of F 10 for sufficiently small ε. In this case, we can apply the so-called averaging theory of first-order dynamical systems. In the case F 10 = 0 but F 20 0 , the zeros of F 10 + ε F 20 are the ones of F 20 for sufficiently small ε. In such cases, we must apply second-order averaging theory.

3. Proof of Proposition 1 and Theorem 1

Proof of Proposition 1. 
We are going to see that the linear part of the characteristic polynomial of system (3) at the equilibrium point x c = ( c , 0 , 0 ) is q ( λ ) = λ 3 + a λ 2 + b λ + c . We shall find the values of the parameters for which q has an eigenvalue equal to zero and a pair of purely imaginary roots, i.e., values for which q is of the form λ λ 2 + b with b > 0 .
To simplify, let b = ω 2 , with ω > 0 . Thus, if we assume q ( λ ) = λ λ 2 + ω 2 , we obtain that a = c = 0 and b > 0 . Therefore, when a = c = 0 and b > 0 , there exists an equilibrium point at the origin of unique coordinates of zero-Hopf type. Moreover, if we consider b = ω 2 , for ω > 0 , then the eigenvalues are equal to 0 and ± i ω , ending the proof of Proposition 1. □
Proof of Theorem 1. 
We are going to use the ideas of the second-order averaging theory of dynamical systems that we have briefly exposed in Section 2.1 (see Theorem 2) for analyzing the existence of equilibrium points of zero-Hopf type at the origin. Recall that such zero-Hopf type equilibrium points bifurcate some periodic orbit by moving the parameters a , b , c of system (3). In this order of ideas, let us consider a = ε α , b = ω 2 + ε β , c = ε γ with ε > 0 being a sufficiently small parameter. Then, the system (3) becomes
x ˙ = y , y ˙ = z , z ˙ = ε γ x ω 2 + ε β y ε α z k = 1 n ( 1 ) k x 2 k ,
The first step in order to write our differential system (6) in the normal form for applying the averaging theory is to write the linear part at the origin of system (6) when ε = 0 as its real Jordan normal form, which is of the form
0 ω 0 ω 0 0 0 0 0 .
Applying this linear change of variables x , y , z X , Y , Z where
x = Z ω X ω 2 , y = Y , z = ω X .
In the new variables X , Y , Z , system (6) becomes
X ˙ = γ ω 11 X α ω 13 X γ ω 10 Z Y ω 12 β ε / ω 13 + ω 6 X 6 6 ω 5 X 5 Z Y ω 14 20 ω 3 X 3 Z 3 + 15 ω 2 X 2 Z 4 + ω 10 X 2 2 ω 9 X Z + ω 8 Z 2 ω 8 X 4 + 4 ω 7 X 3 Z 6 ω 6 X 2 Z 2 + 4 ω 5 X Z 3 ω 4 Z 4 6 ω X Z 5 + 15 ω 4 X 4 Z 2 + Z 6 k = 4 n 1 k Z ω X ω 2 2 k ω 12 / ω 13 , Y ˙ = ω X , Z ˙ = γ ω 11 X α ω 13 X γ ω 10 Z Y ω 12 β ε / ω 12 6 ω 5 X 5 Z + 15 ω 4 X 4 Z 2 20 ω 3 X 3 Z 3 + 15 ω 2 X 2 Z 4 + ω 10 X 2 2 ω 9 X Z + ω 8 Z 2
ω 8 X 4 + 4 ω 7 X 3 Z 6 ω 6 X 2 Z 2 + 4 ω 5 X Z 3 ω 4 Z 4 + ω 6 X 6 6 ω X Z 5 + Z 6 k = 4 n 1 k Z ω X ω 2 2 k ω 12 / ω 12 .
Doing a re-scale of the variables ( X , Y , Z ) in the form ( X , Y , Z ) ( ε u , ε v , ε w ) , then system (8) is expressed as follows
u ˙ = ω v + γ ω 11 u α ω 13 u + ω 10 u 2 γ ω 10 w v ω 12 β 2 ω 9 u w + ω 8 w 2 ε / ω 13 + ω 8 u 4 ω 4 w 4 + 4 ω 7 u 3 w 6 ω 6 u 2 w 2 + 4 ω 5 u w 3 ε 3 / ω 13 + ( ω 6 u 6 6 ω 5 u 5 w + 15 ω 2 u 2 w 4 6 ω u w 5 + 15 ω 4 u 4 w 2 20 ω 3 u 3 w 3 + w 6 ) ε 5 / ω 13 1 ω k = 4 n 1 k w ω u ω 2 2 k ε 2 k 1 , v ˙ = ω u , w ˙ = γ ω 11 u α ω 13 u γ ω 10 w v ω 12 β + ω 10 u 2 2 ω 9 u w + ω 8 w 2 ε / ω 12 + ω 8 u 4 ω 4 w 4 + 4 ω 7 u 3 w 6 ω 6 u 2 w 2 + 4 ω 5 u w 3 ε 3 / ω 12 + ( ω 6 u 6 6 ω 5 u 5 w + 15 ω 2 u 2 w 4 6 ω u w 5 + 15 ω 4 u 4 w 2 20 ω 3 u 3 w 3 + w 6 ) ε 5 / ω 12 k = 4 n 1 k w ω u ω 2 2 k ε 2 k 1 .
The previous system can be written as follows
u ˙ = ω v + ( γ ω 11 u α ω 13 u + ω 10 u 2 γ ω 10 w v ω 12 β 2 ω 9 u w + ω 8 w 2 ) ε / ω 13 + ( ω 8 u 4 ω 4 w 4 + 4 ω 7 u 3 w 6 ω 6 u 2 w 2 + 4 ω 5 u w 3 ) ε 3 / ω 13 + ( ω 6 u 6 6 ω 5 u 5 w + 15 ω 2 u 2 w 4 6 ω u w 5 + 15 ω 4 u 4 w 2 20 ω 3 u 3 w 3 + w 6 ) ε 5 / ω 13 + O ε 7 , v ˙ = ω u , w ˙ = ( γ ω 11 u α ω 13 u γ ω 10 w v ω 12 β + ω 10 u 2 2 ω 9 u w + ω 8 w 2 ) ε / ω 12 + ( ω 8 u 4 ω 4 w 4 + 4 ω 7 u 3 w 6 ω 6 u 2 w 2 + 4 ω 5 u w 3 ) ε 3 / ω 12 + ( ω 6 u 6 6 ω 5 u 5 w + 15 ω 2 u 2 w 4 6 ω u w 5 + 15 ω 4 u 4 w 2 20 ω 3 u 3 w 3 + w 6 ) ε 5 / ω 12 + O ε 7 .
Now, we express the differential system (9) in cylindrical coordinates ( r , θ , w ) defined by u = r cos θ and v = r sin θ , and we obtain
r ˙ = cos θ ( γ ω 11 r cos θ α ω 13 r cos θ + ω 10 r 2 cos θ 2 γ ω 10 w r sin θ ω 12 β 2 ω 9 r cos θ w + ω 8 w 2 ) ε / ω 13 + cos θ ( ω 8 r 4 cos θ 4 ω 4 w 4 + 4 ω 7 r 3 cos θ 3 w 6 ω 6 r 2 cos θ 2 w 2 + 4 ω 5 r cos θ w 3 ) ε 3 / ω 13 + cos θ ( ω 6 r 6 cos θ 6 6 ω 5 r 5 cos θ 5 w + 15 ω 2 r 2 cos θ 2 w 4 6 ω r cos θ w 5 + 15 ω 4 r 4 cos θ 4 w 2 20 ω 3 r 3 cos θ 3 w 3 + w 6 ) ε 5 / ω 13 + O ε 7 , θ ˙ = ω ( sin θ ω 13 α r cos θ + sin θ ω 10 r 2 cos θ 2 sin θ γ ω 10 w + sin θ ω 8 w 2 + ω 12 r β cos θ 2 ω 12 r β 2 sin θ ω 9 r cos θ w + sin θ ω 11 γ r cos θ ) ε / r ω 13 ( 4 sin θ ω 7 r 3 cos θ 3 w 6 sin θ ω 6 r 2 cos θ 2 w 2 + 4 sin θ ω 5 r cos θ w 3 sin θ ω 4 w 4 sin θ ω 8 r 4 cos θ 4 ) ε 3 / r ω 13 ( sin θ ω 6 r 6 cos θ 6 6 sin θ ω 5 r 5 cos θ 5 w + 15 sin θ ω 2 r 2 cos θ 2 w 4 + 15 sin θ ω 4 r 4 cos θ 4 w 2 20 sin θ ω 3 r 3 cos θ 3 w 3 + sin θ w 6 6 sin θ ω r cos θ w 5 ) ε 5 / r ω 13 + + O ε 7 , w ˙ = ( γ ω 11 r cos θ α ω 13 r cos θ + ω 10 r 2 cos θ 2 γ ω 10 w r sin θ ω 12 β 2 ω 9 r cos θ w + ω 8 w 2 ) ε / ω 12 + ( ω 8 r 4 cos θ 4 ω 4 w 4 + 4 ω 7 r 3 cos θ 3 w 6 ω 6 r 2 cos θ 2 w 2 + 4 ω 5 r cos θ w 3 ) ε 3 / ω 12 + ( ω 6 r 6 cos θ 6 6 ω 5 r 5 cos θ 5 w + 15 ω 2 r 2 cos θ 2 w 4 6 ω r cos θ w 5 + 15 ω 4 r 4 cos θ 4 w 2 20 ω 3 r 3 cos θ 3 w 3 + w 6 ) ε 5 / ω 12 + O ε 7 .
In (10) we consider a new independent variable, and we obtain
d r d θ = ε F 11 θ , r , w + ε F 21 θ , r , w + O ε 2 , d w d θ = ε F 12 θ , r , w + ε F 22 θ , r , w + O ε 2 ,
where
F 11 θ , r , w = cos θ ( γ ω 3 r cos θ α ω 5 r cos θ + ω 2 r 2 cos θ 2 γ ω 2 w r sin θ ω 4 β 2 ω r cos θ w + w 2 ) / ω 6 , F 12 θ , r , w = γ ω 3 r cos θ α ω 5 r cos θ + ω 2 r 2 cos θ 2 γ ω 2 w r sin θ ω 4 β 2 ω r cos θ w + w 2 ) / ω 5 ,
and
F 21 θ , r , w = sin θ cos θ ( cos θ 4 ω 4 r 4 2 cos θ ω 5 γ 2 r w + cos θ 2 ω 6 γ 2 r 2 + 2 sin θ ω 9 r 2 β cos θ α + 6 cos θ ω 3 γ r w 2 2 cos θ 2 ω 8 γ r 2 α 6 cos θ 2 ω 4 r 2 γ w + 4 cos θ 2 ω 6 α r 2 w 2 cos θ ω 5 α r w 2 2 cos θ 3 ω 7 α r 3 4 cos θ 3 ω 3 r 3 w + w 4 + ω 8 r 2 β 2 4 cos θ ω r w 3 + ω 4 γ 2 w 2 + 2 cos θ ω 7 α r γ w 2 sin θ ω 4 w 2 r β 2 sin θ ω 6 r 3 β cos θ 2 + 2 cos θ 3 ω 5 γ r 3 + cos θ 2 ω 10 α 2 r 2 cos θ 2 ω 8 r 2 β 2 + 6 cos θ 2 ω 2 r 2 w 2 2 ω 2 w 3 γ 2 sin θ ω 7 r 2 β cos θ γ + 4 sin θ ω 5 r 2 β cos θ w + 2 sin θ ω 6 γ w r β ) / ω 12 r F 22 θ , r , w = sin θ cos θ ( cos θ 4 ω 4 r 4 2 cos θ ω 5 γ 2 r w + cos θ 2 ω 6 γ 2 r 2 + 2 sin θ ω 9 r 2 β cos θ α + 6 cos θ ω 3 γ r w 2 2 cos θ 2 ω 8 γ r 2 α 6 cos θ 2 ω 4 r 2 γ w + 4 cos θ 2 ω 6 α r 2 w 2 cos θ ω 5 α r w 2 2 cos θ 3 ω 7 α r 3 4 cos θ 3 ω 3 r 3 w + w 4 + ω 8 r 2 β 2 4 cos θ ω r w 3 + ω 4 γ 2 w 2 + 2 cos θ ω 7 α r γ w 2 sin θ ω 4 w 2 r β 2 sin θ ω 6 r 3 β cos θ 2 + 2 cos θ 3 ω 5 γ r 3 + cos θ 2 ω 10 α 2 r 2 cos θ 2 ω 8 r 2 β 2 + 6 cos θ 2 ω 2 r 2 w 2 2 ω 2 w 3 γ 2 sin θ ω 7 r 2 β cos θ γ + 4 sin θ ω 5 r 2 β cos θ w + 2 sin θ ω 6 γ w r β ) / ω 12 r
With the notation introduced in Section 2.1, t = θ , T = 2 π , x = ( r , w ) T and
F 1 θ , r , w = F 11 θ , r , w F 12 θ , r , w and F 2 θ , r , w = F 21 θ , r , w F 22 θ , r , w . F 10 r , w = F 101 r , w F 102 r , w and F 20 r , w = F 201 r , w F 201 r , w .
It can easily be checked that (11) satisfies all the assumptions of Theorem 2.
Computing the integrals (5), we have that
F 101 r , w = 1 2 π 0 2 π F 11 θ , r , w d θ = r 2 w + γ ω 2 α ω 4 2 ω 5 , F 102 r , w = 1 2 π 0 2 π F 12 θ , r , w d θ = 2 γ ω 2 w ω 2 r 2 2 w 2 2 ω 5 .
Note that F 101 r , w = F 102 r , w = 0 has only one solution ( r , w ) with r * > 0 , more precisely
r * = 2 ω 4 α 2 + 2 γ 2 ω 2 , w * = ω 2 γ α ω 2 2
The Jacobian evaluated at ( r , w ) has the form
det F 101 , F 101 ( r , w ) ( r , w ) = ( r * , w * ) = γ 2 α 2 ω 4 2 ω 6
which is nonzero by hypothesis.
From it, if γ 2 α 2 ω 4 > 0 the system (6) has a unique periodic orbit bifurcating from the origin.
Note that when γ = α ω 2 , the first-order averaging theory cannot be used. In order to apply the second-order averaging theory, we compute
F 11 r F 11 w F 12 r F 12 w 0 θ F 11 ( s , r , w ) d s 0 θ F 12 ( s , r , w ) d θ + F 21 ( θ , r , w ) F 22 ( θ , r , w ) .
Computing the integral of this expression at the interval [ 0 , 2 π ] and dividing by 2 π , we have
F 201 r , w = r r 2 π 2 β r ω 2 3 w β ω 2 ω 2 α w π + α ω 5 β / ω 8 . F 202 r , w = ( 7 β r 2 ω 5 + 6 ω 3 β w 2 4 ω 8 α β r + 12 w β r ω 4 8 w 3 π 6 ω 7 β α w 4 ω 8 α 2 w π + 2 ω 6 α r 2 π + 12 ω 4 α w 2 π ) / 4 ω 10 .
Let us now consider the system
F 201 r , w = 0 , F 202 r , w = 0 .
If we solve the first equation with respect to w, we obtain
w = r 2 π 2 β r ω 2 + ω 5 α β ω 2 α π ω + 3 β .
Substituting into the second equation gives
2 π 4 ω 13 2 α π ω + 3 β 3 r 6 12 π 3 β ω 11 2 α π ω + 3 β 3 r 5 + 3 α π 3 ω 8 2 α π ω + 3 β 2 3 / 2 β π 2 ω 9 2 α π ω + 3 β 2 + 6 α π 3 β ω 8 2 α π ω + 3 β 3 + 24 π 2 β 2 ω 9 2 α π ω + 3 β 3 r 4 + 12 α π 2 β ω 6 2 α π ω + 3 β 2 3 β π ω 7 2 α π ω + 3 β 16 π β 3 ω 7 2 α π ω + 3 β 3 + 6 β 2 π ω 7 2 α π ω + 3 β 2 + 24 α π 2 β 2 ω 6 2 α π ω + 3 β 3 r 3 + α π 2 ω 4 + α 2 π 2 ω 3 2 α π ω + 3 β 15 β 2 α π ω 4 2 α π ω + 3 β 2 7 β 4 ω 5 + 3 β α π 2 ω 4 2 α π ω + 3 β + 24 α π β 3 ω 4 2 α π ω + 3 β 3 + 6 α 2 π 2 β 2 ω 3 2 α π ω + 3 β 3 + 6 β 2 ω 5 2 α π ω + 3 β 6 α 2 π 2 β ω 3 2 α π ω + 3 β 2 6 β 3 ω 5 2 α π ω + 3 β 2 r 2 + β α ω 2 12 π α 2 β 3 ω 2 α π ω + 3 β 3 2 π α 2 β ω 2 α π ω + 3 β
+ 12 π α 2 β 2 ω 2 α π ω + 3 β 2 6 β 2 α ω 2 2 α π ω + 3 β + 6 β 3 α ω 2 2 α π ω + 3 β 2 r + 3 ω β 2 α 2 4 α π ω + 3 β 3 ω β 3 α 2 2 2 α π ω + 3 β 2 + ω 2 α 3 π β 2 α π ω + 3 β 3 ω 2 α 3 π β 2 2 α π ω + 3 β 2 + 2 ω 2 α 3 π β 3 2 α π ω + 3 β 3 = 0 .
The maximum number of positive solutions to this equation is 6.
Let r ¯ be one of these zeros, and r ¯ , w ¯ be a solution of system (12). To apply Theorem 2, the condition below is needed
det F 201 , F 202 ( r , w ) ( r , w ) = r ¯ , w ¯ 0 .
Thus, the solutions r ¯ , w ¯ of the system (12) holding (13) satisfy ( i ) and ( i i ) of Theorem 2. Thus, by the second-order averaging theory (11) has at most 6 periodic orbits. From it, we state that system (6) has at most 6 periodic orbits bifurcating from the origin ending the proof. □

4. Numerical Verification of the Analytic Results Obtained

We consider the generalizad Genesio differential equation
x . . . + a x . . + b x . + c x + k = 1 n . ( 1 ) k x 2 k = 0
with parameters a = ε α , b = ω + ε β , c = ε γ . The numerical experiments have been made for the following parameters
α = 1 , ω = 1 , β = 0.5 , γ = 1 , n = 5
and inital conditions x ( 0 ) , x . ( 0 ) = y ( 0 ) , x . . ( 0 ) = z ( 0 ) .
In the following pictures and tables (see Table 1, Table 2 and Table 3 and Figure 1 and Figure 2), we obtain the initial conditions and the period of four periodic solutions for different values of ε with six decimal number.
On the other hand, we observe period doubling increasing the value of the parameter ε .
The existence of a period doubling cascade numerically evidences the existence of chaos, as can be seen in the pictures of Orbits 5 to 8. We can estimate, by means of the Feigenbaum constant F 4.669 . . . , the value of the epsilon parameter ε * for which such a phenomenon can occur. The following numerical calculations are carried out in order to obtain such estimation.
ε * ε 1 + ε 2 ε 1 F F 1 = 0.7615 .

5. Conclusions

The application of the averaging theory to study the existence of zero-Hopf bifurcations for a generalized Genesio differential equation gave important results about the periodic structure of these equations. Our future work will be to apply this theory to study zero-Hopf bifurcations for autonomous or nonautonomous differential equations of order n, with n 4 .

Author Contributions

Z.D., J.L.G.G., J.A.V. have worked in an equal way in research, develop and writing the results here presented. All authors have read and agreed to the published version of the manuscript.

Funding

This paper have been partially supported by Ministerio de Ciencia, Innovación y Universidades grant number PGC2018-097198-B-I00 and Fundación Séneca de la Región de Murcia grant number 20783/PI/18.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Cardin, P.T.; Llibre, J. Transcritical and zero-Hopf bifurcations in the Genesio system. Nonlinear Dyn. 2016, 88, 547–553. [Google Scholar] [CrossRef] [Green Version]
  2. Sanders, J.A.; Verhulst, F. Averaging Methods in Nonlinear Dynamical Systems; Springer: New York, NY, USA, 2007. [Google Scholar]
  3. Buzzi, C.; Llibre, J.; Medrado, J. Hopf and zero-Hopf bifurcations in the Hindmarsh-Rose system. Nonlinear Dyn. 2016, 83, 1549–1556. [Google Scholar] [CrossRef] [Green Version]
  4. Castellanos, V.; Llibre, J.; Quilantan, I. Simultaneous periodic orbits bifurcating from two zero-Hopf equilibria in a tritrophic food chain model. J. Appl. Math. Phys. 2013, 1, 31–38. [Google Scholar] [CrossRef] [Green Version]
  5. Llibre, J. Periodic orbits in the zero-hopf bifurcation of the Rössler system. Rom. Astron. J. 2014, 24, 49–60. [Google Scholar]
  6. Llibre, J.; Oliveira, R.D.; Valls, C. On the integrability and the zero-Hopf bifurcation of a Chen-Wang differential system. Nonlinear Dyn. 2015, 80, 353–361. [Google Scholar] [CrossRef] [Green Version]
  7. Llibre, J.; Mereu, A.C.; Teixeira, M.A. Limit cycles of the generalized polynomial Liénard differential equations. Math. Proc. Camb. Philos. Soc. 2010, 148, 363–383. [Google Scholar] [CrossRef] [Green Version]
  8. Feddaoui, A.; Llibre, J.; Chemseddine, B.; Makhlouf, A. Periodic solutions for differential systems in ℝ3 and ℝ4. Appl. Math. Nonlinear Sci. 2020. [Google Scholar] [CrossRef]
  9. Buică, A.; Llibre, J. Averaging methods for finding periodic orbits via Brouwer degree. Bull. Sci. Math. 2004, 128, 7–22. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Orbit 1 to 4.
Figure 1. Orbit 1 to 4.
Mathematics 09 00354 g001
Figure 2. Pictures of Orbits 5 to 8.
Figure 2. Pictures of Orbits 5 to 8.
Mathematics 09 00354 g002aMathematics 09 00354 g002b
Table 1. Initial conditions and periods for ε = 0.0001 , 0.01 , 0.1 and 0.5 .
Table 1. Initial conditions and periods for ε = 0.0001 , 0.01 , 0.1 and 0.5 .
ε x ( 0 ) y ( 0 ) z ( 0 ) T
Orbit 1 0.0001 0.000035 0.000029 0.000024 6.281763
Orbit 2 0.01 0.000080 0.00061 0.00006 6.298837
Orbit 3 0.1 0.014919 0.017764 0.0116793 6.446567
Orbit 4 0.5 0.32917 0.0186038 0.184294 7.293240
Table 2. Initial conditions and periods for ε = 0.735 , 0.745 , 0.758 and 0.7605 .
Table 2. Initial conditions and periods for ε = 0.735 , 0.745 , 0.758 and 0.7605 .
ε x ( 0 ) y ( 0 ) z ( 0 ) T
Orbit 5 0.735 0.381394 0.0267376 0.402723 8.102655
Orbit 6 0.745 0.882003 0.0312616 0.305098 16.32508
Orbit 7 0.758 0.929685 0.0278435 0.268235 33.10797
Orbit 8 0.7605 0.938029 0.0243008 0.260449 66.43312
Table 3. Estimation of the ε parameter.
Table 3. Estimation of the ε parameter.
Bifurcationn ε n ε n ε n 1 F n
T 2 T 1 0.745
2 T 4 T 2 0.758 0.013
4 T 8 T 3 0.7605 0.0025 5.2 = F + 0.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Diab, Z.; Guirao, J.L.G.; Vera, J.A. Zero-Hopf Bifurcation in a Generalized Genesio Differential Equation. Mathematics 2021, 9, 354. https://doi.org/10.3390/math9040354

AMA Style

Diab Z, Guirao JLG, Vera JA. Zero-Hopf Bifurcation in a Generalized Genesio Differential Equation. Mathematics. 2021; 9(4):354. https://doi.org/10.3390/math9040354

Chicago/Turabian Style

Diab, Zouhair, Juan L. G. Guirao, and Juan A. Vera. 2021. "Zero-Hopf Bifurcation in a Generalized Genesio Differential Equation" Mathematics 9, no. 4: 354. https://doi.org/10.3390/math9040354

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