Next Article in Journal
Integrated Order Picking and Multi-Skilled Picker Scheduling in Omni-Channel Retail Stores
Previous Article in Journal
A High–Order WENO Scheme Based on Different Numerical Fluxes for the Savage–Hutter Equations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Number of Limit Cycles Bifurcating from an Elementary Centre of Hamiltonian Differential Systems

1
School of Mathematics, Hangzhou Normal University, Hangzhou 310036, China
2
Department of Mathematics, Shanghai Normal University, Shanghai 200234, China
*
Author to whom correspondence should be addressed.
Mathematics 2022, 10(9), 1483; https://doi.org/10.3390/math10091483
Submission received: 25 March 2022 / Revised: 20 April 2022 / Accepted: 22 April 2022 / Published: 29 April 2022
(This article belongs to the Topic HAT: Hamiltonian Systems—Applications and Theory)

Abstract

:
This paper studies the number of small limit cycles produced around an elementary center for Hamiltonian differential systems with the elliptic Hamiltonian function H = 1 2 y 2 + 1 2 x 2 2 3 x 3 + a 4 x 4 ( a 0 ) under two types of polynomial perturbations of degree m, respectively. It is proved that the Hamiltonian system perturbed in Liénard systems can have at least [ 3 m 1 4 ] small limit cycles near the center, where m 101 , and that the related near-Hamiltonian system with general polynomial perturbations can have at least m + [ m + 1 2 ] 2 small-amplitude limit cycles, where m 16 . Furthermore, in any of the cases, the bounds for limit cycles can be reached by studying the isolated zeros of the corresponding first order Melnikov functions and with the help of Maple programs. Here, [ · ] represents the integer function.
MSC:
34C07; 34E10; 37G15; 37M20

1. Introduction and Statement of the Main Results

The famous Hilbert’s 16th problem is one of the 23 problems posed by the German mathematician David Hilbert [1] in 1900. The second part of Hilbert’s 16th problem is finding the maximum number of limit cycles in a planar polynomial vector field of degree m and investigating their relative positions. Up to now, this problem also remains unsolved, even for m = 2 . In 1977, Arnold [2] proposed a weaker version on Hilbert’s 16th Problem and that can be stated as follows: Consider a polynomial 1-form ω :   = Q ( x , y ) d x P ( x , y ) d y with real polynomials P ( x , y ) and Q ( x , y ) of degree m in x and y. Then, the problem is finding the maximum number of isolated zeroes of Abelian integrals Γ h ω , where Γ h ’s are the compact level curves of polynomials with a given degree. The so-called Poincaré–Pontrjagin theorem shows that the number of isolated zeros of the Abelian integrals is a lower bound of the maximum number of limit cycles of a near-Hamiltonian system of the form
x ˙ = H y ( x , y ) + ε P ( x , y ) , y ˙ = H x + ε Q ( x , y ) ,
where the Hamiltonian H ( x , y ) is a real polynomial of degree n + 1 . The first order Melnikov function (the Abelian integrals) relevant to the perturbed system (1) is
M ( h ) = Γ h Q ( x , y ) d x P ( x , y ) d y ,
where Γ h is a real oval H ( x , y ) = h . Though finding the number of zeros of the Abelian integrals is a weaker problem of the Hilbert’s 16th problem, it is still difficult; see [3].
Smale proposed another restricted version of the Hilbert’s 16th problem, which is to study the maximum number of limit cycles in polynomial Liénard system
x ˙ = y , y ˙ = g ( x ) y f ( x )
with g ( x ) = x and their relative positions, which is called Smale’s 13th problem [4]. For the classical Liénard system, Lins et al. [5] showed it has [ m 2 ] limit cycles locally and conjectured that the number is possibly the maximum number for the global case. Later, De Maesschalck and Dumortier [6] proved it has at least [ m 2 ] + 1 limit cycles for m 5 by a geometric singular perturbation theory.
Smale’s 13th problem was also extended to the generalized Liénard system; see [7,8,9,10,11,12] and the references therein. Most of those cases study the maximum number of small amplitude limit cycles bifurcating from a center or a focus. Llibre et al. [13] exhibited that the generalized Liénard system (3) can have at least [ m + n 1 2 ] small amplitude limit cycles by averaging theory of first, second, and third order. For the system (3) with a quadratic polynomial g ( x ) , the authors in [14,15] proved that the Hopf cyclicity is [ 2 m + 1 3 ] . For the system (3) with a cubic polynomial g ( x ) , Christopher and Lynch [14] showed that the Hopf cyclicity is 2 [ 3 ( m + 2 ) 8 ] for 1 < m 50 using Lyapunov quantities being the coefficients of the monomials ( x 2 + y 2 ) i in the total derivative of the Lyapunov function along trajectories associated with system (3).
The origin is an elementary center for a Hamiltonian system x ˙ = H y ( x , y ) , y ˙ = H x ( x , y ) if the origin is a singularity and the eigenvalues of the matrix ( H y , H x ) ( x , y ) ( 0 , 0 ) are pure imaginary.
Gavrilov and Iliev [8] studied the limit cycles of polynomial near-Hamiltonian systems (1) with the Hamiltonian
H = 1 2 y 2 + 1 2 x 2 2 3 x 3 + a 4 x 4 , a 0 , 8 9 ,
which have four topological structures: a < 0 saddle-loop, 0 < a < 1 eight loop, a = 1 cuspidal loop, and a > 1 global center. Tian et al. [11] considered the generalized Liénard system (3) where g ( x ) = H x ( x , y ) satisfy (4) with a = 8 9 , and proved that the Hopf cyclicity at the origin is [ 3 m + 2 4 ] using involution.
The paper is concerned about the Hamiltonian system with the Hamiltonian (4) having four topological structures, and finds a lower bound for the maximum number of limit cycles appearing from a center under perturbations using a different method. Consider a generalized Liénard system of the form
x ˙ = y , y ˙ = g ( x ) + ε y f ( x ) ,
where f ( x ) and g ( x ) are polynomials of degree m and 3, respectively. Our first result is the following.
Theorem 1.
When a = 1 , 1 2 , 1 or 2, there exist Liénard differential systems of the form (5) with g ( x ) = x ( 1 2 x + a x 2 ) , which have [ 3 m + 2 4 ] small amplitude limit cycles bifurcating from the origin with m = deg ( f ) 100 .
Here, we study a weakened version (5) of the generalized Liénard system (3). The number of limit cycles from a center that we obtain is the same as the Hopf cyclicity in [11] with a = 8 9 , and is attainable in each case of four topological structures of the unperturbed Hamiltonian system. In addition, Wei et al. [12] investigated the Liénard system (5) via the first order Melnikov function, whose the unperturbed system has the Hamiltonian one
H ( x , y ) = 1 2 y 2 1 2 x 2 + 1 4 x 4 ,
including two elementary centers and a double homoclinic loop, and which can have 3 [ m 2 5 ] + 1 + mod ( mod ( m 2 , 5 ) , 2 ) limit cycles inside a single homoclinic loop, with [ m 2 5 ] + mod ( mod ( m 2 , 5 ) , 4 ) small limit cycles near the elementary center inside the loop. The sign mod ( a , b ) stands for the remainder of a divided by b. It is easy to see that the number of limit cycles from the center is better than that of [12].
Get back to the weakened Hilbert’s 16th problem, and look at Newtonian mechanical problems restricting the Hamiltonian to the form H ( x , y ) = y 2 + F n + 1 ( x ) . For the Hamiltonian H ( x , y ) = y 2 + x 4 x 2 λ x having two elementary centers and a double loop to a hyperbolic saddle, Petrov [16] in 1990 proved that the number of zeros of elliptic integrals (2) corresponding to the near-Hamiltonian system (1) in a single loop is not more than m + [ m 3 2 ] , and system (1) exists such that the integrals have at least m + [ m 3 2 ] zeros. Zhao and Zhang [17] in 1999 provided an upper bound 7 m + 5 of the number of isolated zeros of Abelian integrals associated with the system (1) whose Hamiltonian H ( x , y ) = y 2 + F n + 1 ( x ) of degree 4 has at least one center. Liu [18] in 2003 affirmed in the case of the Hamiltonian H ( x , y ) = y 2 + x 4 x 2 λ x that the total number of zeros of the integrals over two periodic annuli inside an eight-loop does not exceed 4 [ m + 1 2 ] 1 , and that the number in the periodic annulus outside the eight-loop does not exceed 4 [ m + 1 2 ] + 1 . Tian and Han [19] exhibited that system (1) with the Hamiltonian (6) can have [ 7 m 6 3 ] limit cycles obtained by studying the isolated zeros of Abelian integrals for m = 3 , 5 , 7 , 9 , being comprised of m 1 limit cycles inside each of a single loop and [ m 3 ] limit cycles outside the double loop. In this paper, we continue to study limit cycles for the cubic Hamiltonian system under m degree polynomial perturbations via zeros of the Abelian integrals. Our second result is the following.
Theorem 2.
When a = 1 , 1 2 , 1 or 2, near-Hamiltonian systems of the form (1) exist with the Hamiltonian (4) which have m + [ m 3 2 ] small amplitude limit cycles surrounding the origin with m 16 .
We note that the number of limit cycles is gained by the isolated zeros of the Abelian integrals in accordance with that of [16]. These limit cycles are located within a homoclinic loop in [16], but in this paper only near an elementary center for each case of four topological structures of the unperturbed Hamiltonian system.
The number of limit cycles in each of the above two theorems is realizable by calculations with the help of Maple programs in Section 3. These programs are theoretically efficient for all m’s. Thus, we conjecture that conclusions in Theorems 1 and 2 hold for all m’s.
In addition, the Maple programs are operational for all a, and one can obtain the lower bound of the number of limit cycles for the near-Hamiltonian system (1) or the generalized Liénard system (5) with the Hamiltonian (4) provided a.
To study small limit cycles produced around an elementary center under small perturbations, one can use Melnikov functions [20,21,22], focus values [23,24], or the averaging method [25,26]. In order to prove Theorem 1, we shall study the Melnikov function for system (5) and present another algorithm to compute the coefficients of its corresponding asymptotic expansion, which will be shown in Theorem 2.
This paper is organized as follows: Section 2 shows some preliminary work to prove our main results. In Section 3, we will prove Theorems 1 and 2. And the section will provide two Maple programs used to compute the rank of Jacobian matrices.

2. Preliminaries

Consider an analytic near-Hamiltonian system of the form:
x ˙ = H y ( x , y ) + ε P 0 ( x , y , δ ) , y ˙ = H x ( x , y ) + ε Q 0 ( x , y , δ ) ,
where H , P 0 , Q 0 C ω ( R 2 ) , ε 0 is a small perturbation parameter and δ D R m is a vector valued parameter with D a compact subset. Suppose that the unperturbed system (7) | ε = 0 has an elementary center C enclosed by a period annulus U. Set h c :   = H ( C ) , and I :   = { h = H ( x , y ) : ( x , y ) U } .
It is well known that the first order Melnikov function M ( h , δ ) near an elementary center for the system (7) is analytic at the end point h c , see [27,28]. One can have the following lemma.
Lemma 1.
For the analytic near-Hamiltonian system (7) whose unperturbed system has an elementary center, the first order Melnikov function near the center has the expansion
M ( h , δ ) = j 0 C j 0 ( δ ) ( h h c ) j + 1 , h U ( h c , τ ) I ¯ ,
where 0 < τ 1 , U ( h c , τ ) stands for a neighborhood of h c , and
C 0 0 = 2 π β P 0 x + Q 0 y | ( C , δ ) , ε = 0
with the eigenvalues ± i β ( β > 0 ) of the center.
Han et al. [21] in 2009 established an algorithm to compute higher order coefficients C j 0 in the first order Melnikov function by developing a Maple program. Later, Tian and Han [19] and Wei and Zhang [29] characterized definitely all coefficients with the help of a homoclinic loop. The characteristics still keep in the case of an elementary center without other conditions.
For the analytic near-Hamiltonian system (7) with analytic perturbations P i and Q i ,   i N , instead of P 0 and Q 0 , one has a new near-Hamiltonian system
x ˙ = H y ( x , y ) + ε P i ( x , y , δ ) , y ˙ = H x ( x , y ) + ε Q i ( x , y , δ ) .
It follows from Lemma 1 that the first order Melnikov function near the center C has the expansion of the form
M i ( h , δ ) = j 0 C j i ( δ ) ( h h c ) j + 1 , h U ( h c , τ ) I ¯ ,
with C 0 i = 2 π β P i x + Q i y | ( C , δ ) , ε = 0 .
Set
Δ l i : = δ D | C j i ( δ ) = 0 , j = 0 , 1 , , l .
The feature of the coefficients is stated as follows.
Lemma 2.
For the analytic near-Hamiltonian system (7), whose unperturbed system has an elementary center, assume that there exist analytic functions P i ( x , y , δ ) and Q i ( x , y , δ ) for i = 1 , 2 , , r , such that, for δ Δ 0 i 1 , the following equality holds:
P i 1 x + Q i 1 y ( x , y , δ ) = H ( x , y ) x P i ( x , y , δ ) + H ( x , y ) y Q i ( x , y , δ ) ,
in a neighborhood of the center. Then, the coefficients of the first order Melnikov functions (8) of system (7) satisfy
C i 0 | Δ i 1 0 = 1 ( i + 1 ) ! C 0 i ,
where C 0 i ’s are given in (9).
In [19], (10) was introduced to study the coefficients in the corresponding expansions of the first order Melnikov functions for bifurcations of limit cycles around an elementary center and a homoclinic loop. This idea was extended to the investigation of limit cycles near a homoclinic loop with a nilpotent singularity in [29] and near a heteroclinic loop in [30], respectively. In Lemma 2, we only focus on finding small-amplitude limit cycles near a center. The next lemma can be found in [27] or Corollary 2.4.1 in [31].
Lemma 3.
For the analytic near-Hamiltonian system (7) whose unperturbed system has an elementary center, supposing that a positive integer n and δ 0 D R s with D bounded and s n exist, such that
C i 0 ( δ 0 ) = 0 , i = 0 , , n 1 , C n 0 ( δ 0 ) 0 , and det ( C 0 0 , C 1 0 , C 2 0 , , C n 1 0 ) ( δ 1 , , δ n ) ( δ 0 ) 0 ,
where δ = ( δ 1 , δ 2 , , δ s ) ; then, for any ε 0 > 0 and any neighborhood V of the elementary center, system (7) in V has precisely n limit cycles for some ( ε , δ ) satisfying 0 < ε < ε 0 and | δ δ 0 | < ε 0 .
Hopf bifurcation by the feature in Lemma 2 follows immediately. Set
C i :   =   C 0 0 , for i = 0 , and C i :   =   C i 0 | Δ i 1 0 , for i > 0 .
Theorem 3.
Let the assumptions in Lemma 2 hold. If δ 0 R s exists, and a positive integer n such that
C i ( δ 0 ) = 0 , i = 0 , , n 1 , C n ( δ 0 ) 0 , and rank ( C 0 , C 1 , C 2 , , C n 1 ) ( δ 1 , , δ s ) ( δ 0 ) = n ,
then system (7) has exactly n limit cycles near an elementary center for some ( ε , δ ) near ( 0 , δ 0 ) .
Proof. 
To study isolated zeros of the first order Melnikov function (Abelian integrals) (2), we display the relations between C i and C i 0 , of the coefficients in the Melnikov function.
It follows from (11) that C 0 0 = C 0 , and consequently for i 1
C i 0 = C i 0 | Δ i 1 , 0 + O | C 0 0 , C 1 0 , , C i 1 0 | = C i + O | C 0 0 , C 1 0 , , C i 1 0 | .
By induction, one obtains that O | C 0 0 , C 1 0 , , C i 1 0 | = O | C 0 , C 1 , , C i 1 | for each i 1 . Namely,
C i 0 = C i + O | C 0 , C 1 , , C i 1 | ,
satisfying the conditions in Lemma 3. We finish the proof. □
We note that the number of limit cycles in Theorem 3 is obtained from the perspective of theoretical analysis. The theorem can be applied to the Liénard system (5) or the near-Hamiltonian system (1) with a specific unperturbed Hamiltonian system having an elementary center. However, it is difficult to show definite expressions for the perturbations such that systems (5) or (1) have a certain number of limit cycles, since we can take C 0 , C 1 , , C n 1 as free parameters such that
| C j | | C j + 1 | 1 and C j · C j + 1 < 0 ,   j = 0 , 1 , , n 2 .
In addition, then
| C j 0 | | C j + 1 0 | 1 and C j 0 · C j + 1 0 < 0 ,   j = 0 , 1 , , n 2 ,
hold theoretically. However, in practice, it is not easy to guarantee the relation | C j 0 | | C j + 1 0 | due to the term O | C 0 , C 1 , , C j 1 | .
To simplify calculations for some specific systems, we show the following corollary further.
Corollary 1.
Let the assumptions in Lemma 2 hold. If there exists a positive integer n such that
rank ( C 0 , C 1 , C 2 , , C n ) ( δ 1 , , δ s ) = n + 1 ,
where δ = ( δ 1 , δ 2 , , δ s ) R s , then system (7) has n limit cycles near the elementary center for some ( ε , δ ) .

3. Proofs of Theorems 1 and 2

The Hamiltonian (4) with a 0 of the unperturbed system (5) | ε = 0 or (1) | ε = 0 consists of an elementary center at the origin.
Proof 
(Proof of Theorem 1). Consider the Liénard differential system (5) with a polynomial f m = i = 0 m b i x i . By Lemma 1, one has
C 0 = 2 π b 0 .
Using Taylor expansion at x = 0 yields
1 1 2 x + a x 2 = j = 0 x j ( 2 a x ) j = l = 0 x l S l ,
where S l = j = [ l + 1 2 ] l j ! ( a ) l j · 2 2 j l ( l j ) ! ( 2 j l ) ! .
Since P 0 = 0 and Q 0 = f m ( x ) y . In the light of the condition Δ 0 0 in Theorem 2, being f m ( 0 ) = 0 , the equality 10 holds by taking
P 1 = f m ( x ) | Δ 0 0 x ( 1 2 x + a x 2 ) and Q 1 = 0 ,
which gives
P 1 = l = 0 b l 1 x l
in a neighborhood of zero, where
b l 1 = j = 0 l b j + 1 S l j , for 0 l m 1 , j = 0 m 1 b j + 1 S l j , for l m ,
with S l given in (13).
Set
P i 1 :   = l = 0 b l i 1 x l and Q i 1 :   = 0
for i 2 in a neighborhood of zero. It follows from the condition Δ 0 i 1 and the equality (10) that
P i = l = 0 b l i x l and Q i = 0
in a neighborhood of zero, where
b l i = j = 0 l ( j + 2 ) b j + 2 i 1 S l j
for all l’s.
By induction, one has for 1 i [ m 2 ]
b 1 i = j = 1 i 1 A j B i ( b 1 , b 2 , , b 2 i ) T ,
and, for i > [ m 2 ]
b 1 i = j = 1 i 1 A j B i ( b 1 , b 2 , , b m ) T ,
where A 1 = ( 2 S 1 , 3 S 0 ) , A j ’s are the ( 2 j 1 ) × ( 2 j ) matrices
A j = 2 S 2 3 S 1 4 S 0 0 0 2 S 3 3 S 2 4 S 1 5 S 0 0 2 S 4 3 S 3 4 S 2 5 S 1 0 2 S 2 j 1 3 S 2 j 2 4 S 2 j 3 5 S 2 j 4 ( 2 j + 1 ) S 0 ,
for 2 j i 1 , and B 1 = ( S 1 , S 0 ) , B i ’s are the ( 2 i 2 ) × ( 2 i ) matrices
B i = S 2 S 1 S 0 0 0 S 3 S 2 S 1 S 0 0 S 4 S 3 S 2 S 1 0 S 2 i 1 S 2 i 2 S 2 i 3 S 2 i 4 S 0 ,
for 2 i [ m 2 ] , and B i ’s are the ( 2 i 2 ) × m matrices
B i = S 2 S 1 S 0 0 0 S 3 S 2 S 1 S 0 0 S 4 S 3 S 2 S 1 0 S m 1 S m 2 S m 3 S m 4 S 0 S m S m 1 S m 2 S m 3 S 1 S 2 i 1 S 2 i 2 S 2 i 3 S 2 i 4 S 2 i m ,
for i > [ m 2 ] with S l given by (13). Noting that the expression j = 1 i 1 A j in (14) or (15) represents an identity matrix for i = 1 .
According to Lemma 2 and the expressions (9) and (11), one obtains
C i = 2 π ( i + 1 ) ! b 1 i .
Substituting (14) and (15) into the last expression, one obtains
C i = 2 π ( i + 1 ) ! j = 1 i 1 A j B i ( b 1 , b 2 , , b 2 i ) T ,
for 1 i [ m 2 ] , and
C i = 2 π ( i + 1 ) ! j = 1 i 1 A j B i ( b 1 , b 2 , , b m ) T ,
for i > [ m 2 ] .
On the basis of (12), (16), and (17), it is easy to see that the Jacobian matrix
( C 0 , C 1 ) ( b 0 , b 1 , , b m ) = 2 π 0 0 2 π ,
is of full rank for m = 1 , and we show via the Maple program (see Algorithm 1) that the Jacobian matrices
C 0 , C 1 , , C [ 3 m + 2 4 ] ( b 0 , b 1 , , b m ) ,
for 2 m 100 , are of full rank, if a = 1 , 1 2 , 1 , or 2.
By the arguments on the rank of the Jacobian matrix of the coefficients in the expansion of the first order Melnikov function with respect to the coefficients b = ( b 0 , b 1 , , b m ) of f m , one achieves that the C j ’s satisfy the conditions of Corollary 1 or Theorem 3 with n = [ 3 m + 2 4 ] . Consequently, the Liénard differential system 5 has [ 3 m + 2 4 ] limit cycles for suitable choices of its coefficients.
It completes the proof of the theorem. □
Proof of Theorem 2. 
Consider the near-Hamiltonian system (1) with the Hamiltonian (4). For simplicity, set
F 0 ( x , y ) :   = P 0 x + Q 0 y = i + j = 0 m 1 d i j x i y j .
By Lemma 1, we have
C 0 = 2 π d 00 .
The condition Δ 0 0 in Lemma 2 indicates F 0 ( 0 , 0 ) = d 00 = 0 . Under the condition Δ 0 0 , the equality 10 with i = 1 holds by taking
P 1 = F 0 ( x , y ) | d 00 = 0 x ( 1 2 x + a x 2 ) and Q 1 = 1 y ( F 0 ( x , y ) F 0 ( x , 0 ) ) ,
that are in a neighborhood of zero
P 1 ( x , y ) = i = 0 a i 1 x i and Q 1 ( x , y ) = i + j = 0 m 2 b i j 1 x i y j ,
where b i j 1 = d i , j + 1 , and
a i 1 = k = 0 i d k + 1 , 0 S i k , 0 i m 2 , k = 0 m 2 d k + 1 , 0 S i k , l m 1 .
In a neighborhood of zero, let
P l 1 :   = i = 0 a i l 1 x i and Q l 1 :   = i + j = 0 m 2 l + 2 b i j l 1 x i y j
for l 2 . It follows that
F l 1 :   = P l 1 x + Q l 1 y = i + j = 0 m 2 l + 1 d i j l 1 x i y j + i = m 2 l + 2 d i 0 l 1 x i ,
where
d i j l 1 = ( i + 1 ) a i + 1 l 1 + ( j + 1 ) b i , j + 1 l 1 , j = 0 , 0 i m 2 l + 1 , ( i + 1 ) a i + 1 l 1 , j = 0 , i m 2 l + 2 , ( j + 1 ) b i , j + 1 l 1 , j 1 , 1 i + j m 2 l + 1 .
As is stated above, under the condition Δ 0 l 1 , the equality (10) holds by ordering in a neighborhood of x = 0
P l = i = 0 a i l x i , and Q l = i + j = 0 m 2 l b i j l x i y j ,
where
a i l = k = 0 i d k + 1 , 0 l 1 S i k and b i j l = d i , j + 1 l 1 ,
with S i k given by (13).
By induction, one has
d 00 1 = d 10 S 1 , m = 2 , B 1 ( d 10 , d 20 ) T + d 02 , m 3 ,
and for 2 l m + [ m + 1 2 ] 2 ,
d 00 l = i = 1 l 1 ( 2 i 1 ) ! ! j = 1 l i A j ( d 1 , 2 i , d 2 , 2 i , , d 2 l 2 i , 2 i ) T + ( 2 l 1 ) ! ! d 0 , 2 l + j = 1 l A j ( d 10 , d 20 , , d 2 l , 0 ) T , 2 l m 1 , j = 1 l 1 A j B l ( d 10 , d 20 , , d m 1 , 0 ) T , 2 l m ,
where A 1 = B 1 = ( S 1 , S 0 ) , A j ’s are the ( 2 j 1 ) × ( 2 j ) matrices
A j = 2 S 2 2 S 1 2 S 0 0 0 3 S 3 3 S 2 3 S 1 3 S 0 0 ( 2 j 1 ) S 2 j 1 ( 2 j 1 ) S 2 j 2 ( 2 j 1 ) S 2 j 3 ( 2 j 1 ) S 2 j 4 ( 2 j 1 ) S 0 ,
for 2 j l , and B l ’s are the ( 2 l 2 ) × ( m 1 ) matrices
B l = 2 S 2 2 S 1 2 S 0 0 0 3 S 3 3 S 2 3 S 1 3 S 0 0 ( m 2 ) S m 2 ( m 2 ) S m 3 ( m 2 ) S m 4 ( m 2 ) S m 5 ( m 2 ) S 0 ( m 1 ) S m 1 ( m 1 ) S m 2 ( m 1 ) S m 3 ( m 1 ) S m 4 ( m 1 ) S 1 ( 2 l 1 ) S 2 l 1 ( 2 l 1 ) S 2 l 2 ( 2 l 1 ) S 2 l 3 ( 2 l 1 ) S 2 l 4 ( 2 l 1 ) S 2 l m + 1 ,
and d i j = 0 if i + j m .
According to Lemma 2 and the expressions (9) and (11), one has for l = 1 , 2 , , m + [ m + 1 2 ] 2 ,
C l = 2 π ( l + 1 ) ! d 00 l .
According to the Formulas (19)–(22), we find that C i ’s depend solely on the parameters d i , 2 j in the divergence of the perturbations P 0 and Q 0 . Let E m :   = { d i , 2 j : i + 2 j m 1 } with [ m + 1 2 ] [ m + 2 2 ] elements. It follows from (19) and (20)–(22) that the Jacobian matrix
( C 0 , C 1 ) ( d 00 , d 10 ) = 2 π 0 0 2 π ,
is of full rank for m = 2 . By using the Maple program (see Algorithm 2), we have that the Jacobian matrices
C 0 , C 1 , , C m + [ m + 1 2 ] 2 E m ,
are of full rank for 3 m 16, if a = 1 , 1 2 , 1 , or 2.
Combing Corollary 1 or Theorem 3 and the rank of the Jacobian matrices, one achieves that the near-Hamiltonian system 5 has m + [ m + 1 2 ] 2 limit cycles for suitable choices of the parameters.
It completes the proof of the theorem. □
We note that, in the previous discussions, for any m 2 , theoretically, it is possible to obtain the rank of the Jacobian matrices with a specific a via Algorithms 1 and 2, only if a computer has a sufficiently strong CPU and the memory ensuring its running for enough time. If we choose m = 100 , then Algorithm 1 takes 3955 s to run. However, Algorithm 2 runs only to m = 16 due to the limitation of the memory of our computer (16 GB).
Algorithm 1 The rank of the Jacobian matrices (18)
restart: with (Linear Algebra):
num: = proc (a,m)
local S, F, A, F1, B, BMy, M, c, i;
       S: = proc (n);
       local i, sum;
              sum = 0;
              for i from trunc (1/2 ∗ n + 1/2) to n do
                     sum: = sum + (i)! ∗ ((n−i)! ∗ (2i−n)!)−1 ∗ (−a)n−1 ∗ 2 2i−n
              od
               sum
       end proc
       F: = proc (i,j)→(j+1) ∗ S(i+2−j);
       A: = proc (j)
       local AM;
              if j = 0 then
                     AM: Matrix (1, 1, 1)
              else
                     if j = 1 then
                            AM: = Matrix ([4, 3])
                     else
                            AM: = Matrix (2 ∗ j−2, 2 ∗ j, F)
                     end if;
              end if;
              AM
       end proc;
       F1: = proc (i, j) → S (i + 2 − j);
       B: = proc (i)
       local BM;
              if i = 1 then
                     BM: = Matrix ([2, 1])
              else
                     if i ≤ trunc (1/2 ∗ m) then
                            BM: = Matrix (2 ∗ i − 2, 2 ∗ i, F1)
                     else
                            BM: = Matrix (2 ∗ i − 2, m, F1)
                     end if;
              end if;
              BM
       end proc;
       BMy: = proc (i)
       local AM, j;
              AM: = Matrix (1, 1, 1);
              for j from 0 to i − 1 do
                     AM: = MatrixMatrixMultiply(AM, A (j))
              od;
              MatrixMatrixMultiply (AM, B (i))
       end proc;
       M:= proc (i)
              if i ≤ trunc (1/2 ∗ m) then
                     << 0 >|< BMy(i) >|< Matrix(1, m−2 ∗ i, 0)> >
              else
                     << 0 >|< BMy(i) >>
              end if;
       end proc;
       c: = <<1>| Matrix (1, m, 0) > ;
       for i to trunc (3/4 ∗ m + 1/2) do
              c: = <c,M (i)>
       od;
       Rank (c)
end proc;
Algorithm 2 The rank of the Jacobian matrices (23)                                                                                                          
restart: with (Linear Algebra):
rank: = proc (a,m)
local  k, PQ, l, i, divPQ, C, diPQ, cof, G, G1, E, e, ser;
    k: = trunc (1/2 * m + 1/2) * trunc (1/2 * m + 1);
     PQ: = 0;
    for l from 0 to m -1 do
        for i from 0 to l do
            PQ: = PQ + c[i, l-i] * x i * y l i
        end do;
    end do;
    divPQ[0]: = PQ;
    C[0]: = subs ([x = 0, y = 0], divPQ[0]);
    ser: = proc (l)
        convert (series (1/(1-2 * x + a * x 2 ), x = 0, l), polynom)
    end proc;
    diPQ: = proc (l, F)
    local P, Q;
        P: = sort (collect ((subs (y = 0, F)-subs ([x = 0, y = 0], F)) * x 1 * ser (2 * k-2 * l + 2), x));
        Q: = sort (expand (simplify ((F-subs (y = 0, F)) * y 1 )), [x, y]);
        sort (diff (P, x) + diff (Q, y), [x, y])
    end proc;
    for i to k do
        divPQ[i]: = diPQ (i, divPQ[i -1]);
        C[i]: = (factorial (i + 1)) 1 * subs ([x = 0, y = 0], divPQ[i])
    end do;
    cof: = proc (i)
    local f, l, j;
        f: = 0;
        for l from 0 to i -1 do
            for j from 0 to trunc (1/2 * l) do
                f: = f + c[l-2 * j, 2 * j] * x l 2 j * y 2 j
            end do;
        end do;
        coeffs (f, [x, y])
    end proc;
    G: = 0;
    for i from 0 to m + trunc (1/2 * (m -1)) -1 do
        G: = G + C[i] * x i
    end do;
    G1: = coeffs (G, x);
    E, e: = GenerateMatrix (G1, [cof (m)]);
    Rank (E)
end proc;

4. Conclusions

Tian and Han [19] provided a new idea in 2017, which obtains the expressions of the high order coefficients in the asymptotic expansion of the first order Melnikov function (Abelian integrals) near a homoclinic loop under some additional conditions, to obtain more limit cycles near a (double) homoclinic loop. The new idea is to introduce an elementary center. Wei et al. [12] in 2021 further developed their results based on their good idea, and obtained more limit cycles in an ( m + 1 ) th degree generalized Liénard differential system with the Hamiltonian one of degree 4. Inspired by the new idea, we apply it to Hopf bifurcation near an elementary center in this paper, aiming at finding more limit cycles of specific systems with perturbations of degree large m.
This work is to find small-amplitude limit cycles generated from an elementary center of Hamiltonian systems with the 4-degree Hamiltonian H = 1 2 y 2 + 1 2 x 2 2 3 x 3 + a 4 x 4 ( a 0 ) , which have four phase portraits, under two types of polynomial perturbations of degree m. The lower bounds of the maximum number of limit cycles are given by Theorems 1 and 2. Obviously, the same number of limit cycles, in the perturbed systems with given m, always appears near an elementary center regardless of phase portraits of the unperturbed Hamiltonian system. To obtain the lower bounds, two Maple Programs are showed in Section 3, and make the lower bounds come true via the first order Melnikov function.

Author Contributions

Conceptualization, L.W. and Y.T.; Formal analysis, L.W. and Y.T.; Investigation, L.W. and Y.T.; Methodology, L.W., Y.T. and Y.X.; Software, L.W.; Validation, L.W. and Y.T.; Writing—original draft, L.W. and Y.T.; Writing—review and editing, L.W., Y.T. and Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

The first author is partially supported by NNSF of China Grant No. 11901144. The second author is partially supported by NNSF of China Grant No. 11871042 and by Shanghai Rising-Star Program of China Grant No. 18QA1403300. The third author is partially supported by NSF of Zhejiang Province Grant No. LY20A010002, and by NNSF of China Grant No. 11671114.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to the referees for their suggestions and comments, which improve our paper in presentation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hilbert, D. Mathematical problems. Bull. Am. Math. Soc. 1902, 8, 437–479. [Google Scholar] [CrossRef] [Green Version]
  2. Arnold, V.I. Loss of stability of self-induced oscillations near resonance, and versal deformations of equivariant vector fields. Funkcional. Anal. Priložen. 1977, 11, 95. [Google Scholar]
  3. Li, J. Hilbert’s 16th problem and bifurcations of planar polynomial vector fields. Int. J. Bifurc. Chaos Appl. Sci. Eng. 2003, 13, 47–106. [Google Scholar] [CrossRef] [Green Version]
  4. Smale, S. Mathematical problems for the next century. Math. Intell. 1998, 20, 7–15. [Google Scholar] [CrossRef]
  5. Lins, A.; de Melo, W.; Pugh, C.C. On Liénard’s equation. In Proceedings of the Geometry and Topology (Proc. III Latin Amer. School of Math., Inst. Mat. Pura Aplicada CNPq, Rio de Janeiro, 1976); Lecture Notes in Mathematics. Springer: Berlin, Germany, 1977; 597, pp. 335–357. [Google Scholar]
  6. De Maesschalck, P.; Dumortier, F. Classical Liénard equations of degree n≥6 can have [ n 1 2 ] + 2 limit cycles. J. Differ. Equ. 2011, 250, 2162–2176. [Google Scholar] [CrossRef]
  7. Dumortier, F.; Li, C. On the uniqueness of limit cycles surrounding one or more singularities for Liénard equations. Nonlinearity 1996, 9, 1489–1500. [Google Scholar] [CrossRef]
  8. Gavrilov, L.; Iliev, I.D. Cubic perturbations of elliptic Hamiltonian vector fields of degree three. J. Differ. Equ. 2016, 260, 3963–3990. [Google Scholar] [CrossRef]
  9. Lloyd, N.G.; Lynch, S. Small-amplitude limit cycles of certain Liénard systems. Proc. R. Soc. Lond. Ser. A 1988, 418, 199–208. [Google Scholar]
  10. Sheng, L.; Han, M. Bifurcation of limit cycles from a compound loop with five saddles. J. Appl. Anal. Comput. 2019, 9, 2482–2495. [Google Scholar] [CrossRef]
  11. Tian, Y.; Han, M.; Xu, F. Bifurcations of small limit cycles in Liénard systems with cubic restoring terms. J. Differ. Equ. 2019, 267, 1561–1580. [Google Scholar] [CrossRef]
  12. Wei, L.; Zhang, Q.; Zhang, X. On limit cycles near two centers and a double homoclinic loop in Liénard differential system. J. Differ. Equ. 2021, 300, 226–251. [Google Scholar] [CrossRef]
  13. 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]
  14. Christopher, C.; Lynch, S. Small-amplitude limit cycle bifurcations for Liénard systems with quadratic or cubic damping or restoring forces. Nonlinearity 1999, 12, 1099–1112. [Google Scholar] [CrossRef]
  15. Han, M. Liapunov constants and Hopf cyclicity of Liénard systems. Ann. Differ. Equ. 1999, 15, 113–126. [Google Scholar]
  16. Petrov, G.S. Nonoscillation of elliptic integrals. Funktsional. Anal. i Prilozhen. 1990, 24, 45–50, 96. [Google Scholar] [CrossRef]
  17. Zhao, Y.; Zhang, Z. Linear estimate of the number of zeros of abelian integrals for a kind of quartic Hamiltonians. J. Differ. Equ. 1999, 155, 73–88. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, C. Estimate of the number of zeros of Abelian integrals for an elliptic Hamiltonian with figure-of-eight loop. Nonlinearity 2003, 16, 1151–1163. [Google Scholar] [CrossRef]
  19. Tian, Y.; Han, M. Hopf and homoclinic bifurcations for near-Hamiltonian systems. J. Differ. Equ. 2017, 262, 3214–3234. [Google Scholar] [CrossRef]
  20. Han, M.; Yang, J. The Maximum Number of Zeros of Functions with Parameters and Application to Differential Equations. J. Nonlinear Model. Anal. 2021, 3, 13–34. [Google Scholar] [CrossRef]
  21. Han, M.; Yang, J.; Yu, P. Hopf bifurcations for near-Hamiltonian systems. Int. J. Bifurc. Chaos Appl. Sci. Eng. 2009, 19, 4117–4130. [Google Scholar] [CrossRef] [Green Version]
  22. Yu, P.; Han, M. Four limit cycles in quadratic near-integrable systems. J. Appl. Anal. Comput. 2011, 1, 291–298. [Google Scholar]
  23. Arcet, B.; Romanovski, V.G. On Some Reversible Cubic Systems. Mathematics 2021, 9, 1446. [Google Scholar] [CrossRef]
  24. Christopher, C. Estimating limit cycle bifurcations from centers. In Differential Equations with Symbolic Computation; Trends in Mathematics; Birkhäuser: Basel, Switzerland, 2005; pp. 23–35. [Google Scholar] [CrossRef]
  25. Benterki1, R.; Llibre, J. Periodic Solutions of a Class of Duffing DifferentialEquations. J. Nonlinear Model. Anal. 2019, 1, 167–177. [Google Scholar]
  26. 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]
  27. Han, M. On Hopf cyclicity of planar systems. J. Math. Anal. Appl. 2000, 245, 404–422. [Google Scholar] [CrossRef] [Green Version]
  28. Li, W.; Llibre, J.; Zhang, X. Melnikov functions for period annulus, nondegenerate centers, heteroclinic and homoclinic cycles. Pac. J. Math. 2004, 213, 49–77. [Google Scholar] [CrossRef] [Green Version]
  29. Wei, L.; Zhang, X. Limit cycles bifurcating from periodic orbits near a center and a homoclinic loop with a nilpotent singularity of Hamiltonian systems. Nonlinearity 2020, 33, 2723–2754. [Google Scholar] [CrossRef]
  30. Geng, W.; Tian, Y. Bifurcation of limit cycles near heteroclinic loops in near-Hamiltonian systems. Commun. Nonlinear Sci. Numer. Simul. 2021, 95, 105666. [Google Scholar] [CrossRef]
  31. Han, M. Bifurcation Theory of Limit Cycle; Science Press: Beijing, China, 2013. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wei, L.; Tian, Y.; Xu, Y. The Number of Limit Cycles Bifurcating from an Elementary Centre of Hamiltonian Differential Systems. Mathematics 2022, 10, 1483. https://doi.org/10.3390/math10091483

AMA Style

Wei L, Tian Y, Xu Y. The Number of Limit Cycles Bifurcating from an Elementary Centre of Hamiltonian Differential Systems. Mathematics. 2022; 10(9):1483. https://doi.org/10.3390/math10091483

Chicago/Turabian Style

Wei, Lijun, Yun Tian, and Yancong Xu. 2022. "The Number of Limit Cycles Bifurcating from an Elementary Centre of Hamiltonian Differential Systems" Mathematics 10, no. 9: 1483. https://doi.org/10.3390/math10091483

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