1. Introduction
The determination of periodic solution of a nonlinear oscillator and its frequency of vibration is one of the most important fields in nonlinear physical problems. The computational methods that have been developed for analytically finding the periodic solutions involve perturbation methods [
1,
2,
3,
4], the variational iteration method [
5,
6,
7,
8], the harmonic balance method [
9,
10,
11], and the homotopy perturbation method [
6,
12,
13,
14,
15], to name a few.
Originally, the variational iteration method [
5] was applied to the initial value problem. Khuri and Sayfy [
16] extended the variational iteration method to the boundary value problem by using two Lagrange multipliers in the corrected functional. Recently, Anjum and He [
17] developed a dual Lagrange multiplier approach for the dynamics of mechanical vibrations, obtaining accurate frequency formulas and periodic solutions. The Hamiltonian-based frequency formulation [
18] is a modification of He’s frequency formulation [
19]. Because the Hamiltonian-based frequency–amplitude formulation takes account of the energy of the nonlinear vibration system to establish a Hamilton principle, its approximate solution is valid for the whole solution domain without limitations such as those in the traditional perturbation method. Many scholars have developed the dual Lagrange multiplier approach and Hamiltonian-based frequency–amplitude formulation to solve the nonlinear oscillators and the related problems, such as by using the dual Lagrange multiplier approach to the dynamics of the mechanical systems [
17], to investigate the nonlinear vibration process of a conservative oscillator by utilizing the Hamiltonian-based frequency–amplitude formulation [
18], by using the dual Lagrange multiplier to analyze large deformation contact problems [
20], by employing the modified mortar contact algorithm to deal with the modelling of contact problems [
21], by applying the Lagrange multiplier to compute the single-phase fluid flow problems in fracture dominated porous media [
22], to explore a complex nonlinear vibration system by using the simplified Hamiltonian-based frequency–amplitude formulation [
23], on the basis of Hamiltonian-based frequency–amplitude formulation and He’s new frequency formula, the dynamics of multi-walled carbon nanotube actuators near graphite sheets being addressed [
24], and to deal with the large amplitude vibration of nonlinear engineering structures by employing the Hamiltonian-based frequency formulation [
25].
The harmonic balance method (HBM) is a computationally efficient method for approximating the periodic solutions of nonlinear oscillators, which is first by substituting a trial solution in terms of the Fourier series into the governing equations. Then, the governing equations are expanded, and the terms associated with each harmonic element of
and
, are balanced. In general, this process could be very complicated when the nonlinearity is strong and the order of
m is increased, which may require help from the symbolic operation to generate the correct system of nonlinear algebraic equations to solve the unknown expansion coefficients. Liu et al. [
11] have improved this procedure by a collocation method on the discretized time points within a period, which, however, may generate the non-physical solution, because in order to keep the same spirit of HBM, they approximate the nonlinear term by truncating the contributions from other higher modes.
Traditionally, the HBM derived a sequence of nonlinear algebraic equations for seeking the analytic solutions, which makes it hard to attain the higher-order analytical solutions for a strongly nonlinear oscillator. To simplify the HBM, some modified harmonic balance methods to solve various nonlinear ordinary differential equations (ODEs) can be referred to in [
10,
26,
27,
28,
29,
30].
How to determine the frequency and the period of vibration for the cubically nonlinear jerk differential equations is a difficult issue in the periodic problems, of which some analytical methods were developed [
19,
31,
32]. In general, they are limited to a lower-order approximation. A laborious work to treat the lengthy nonlinear terms hinders the derivation of higher-order approximation. Mickens [
33] showed several methods to find the analytic solutions of second-order nonlinear ODEs and the systems of first-order ODEs; they are, in general, effective for the weakly nonlinear systems.
Most of the nonlinear ODEs for modeling nonlinear oscillations are of second order; a few dynamical systems can be described by the third-order nonlinear ODEs, like the oscillations in a nonlinear vacuum tube circuit [
34] and thermal–mechanical oscillator in fluids [
35]. In [
36], the residue harmonic balance approach was used to simulate the limit cycles of nonlinear jerk equations; Ramos [
37] developed an order reduction method for seeking the periodic solutions of nonlinear third-order ordinary differential equations. El-Dib [
38] addressed the non-periodic-type analytic solutions of some cubic nonlinear jerk oscillator with the non-perturbative method.
Besides mechanics and engineering tools, jerks can be exercised in the study of various electromagnetic systems [
39], and also investigated to recognize the dynamics of the Earth’s fluid [
40]. The jerk equation helps scientists and engineers establish systems that run smoothly and efficiently, ameliorating user experience and system longevity, e.g., vehicles, elevators, and robotics. The jerk equation is important in the research of motion because it delineates the rate of change of acceleration with respect to time. This idea is especially significant in fields like physics, engineering, and robotics for realizing the smooth motion control, the system performance, and the trajectory planning.
As pointed out by Mickens [
33], an important advantage of the harmonic balance method (HBM) is that it can be applied to the nonlinear oscillatory problem for which the nonlinear term is not small. When the HBM is properly used, it gives excellent approximation to the periodic solution. Unfortunately, the use of HBM leads to very complicated nonlinear algebraic equations that have to be solved for the analytic periodic solution with order greater than two. To avoid the solution of nonlinear algebraic equations and for saving much computational cost without a lengthy derivation of these nonlinear equations, we are going to propose a powerful linearized harmonic balance method (LHBM) in the paper. It would be apparent that the LHBM outperforms the HBM and its modification appeared in the literature. This study, by combining the decomposition–linearization technique [
41,
42] with the HBM for seeking the periodic solutions of second-order nonlinear oscillators and third-order nonlinear jerk oscillators, is fully a novel work endowed with high originality.
The background of LHBM is a decomposition–linearization technique, which is executed on the nonlinear ODE by changing it to a linear ODE around a referenced solution. Previously, the decomposition–linearization technique was combined with the Lindstedt–Poincaré method in [
41] to determine the second-order nonlinear free vibration of nonlinear oscillator. Then, the decomposition–linearization technique was combined with the homotopy perturbation method in [
42] to treat the nonlinear differential/integral equations and nonlinear jerk equations. The LHBM is drastically different from the above two methods; it does not need to introduce any perturbation parameter in the derivation of the frequency and the periodic solution of nonlinear oscillator.
2. Second-Order Nonlinear Oscillators
We consider a second-order nonlinear oscillator:
where
and
are nonlinear functions of
.
For different systems, there exist different fundamental solutions. For instance,
is a suitable initial guess for the system that satisfies the initial conditions
and
; however, it has an unknown period
to be sought.
Not all second-order ODEs in Equation (
1) permit the periodic motion or limit cycle [
43]. In the paper, we consider a special class of Equation (
1) of which the periodic solution exists, because our method is based on the harmonic balance method. For the Liénard equation, there are well-established conditions, as described by Theorem 3.2.1 in [
43], for the existence of a unique periodic solution. In general, He and Garcia [
44] derived necessary and sufficient conditions for
having a periodic solution satisfying
,
and
; they are the existence of a function
in
, such that
and
.
Equation (
1) involves many second-order nonlinear oscillators as special cases. We name a few:
Those ODEs have a lot of applications in engineering and science, being the most widely used models in the study of nonlinear oscillations. The second-order nonlinear oscillators are dynamical systems that exhibit complex and often chaotic behavior under periodic forcing [
45]. Equation (
1) is the general nonlinear oscillator, which can depict various issues, such as the cubic Duffing equation [
46], the cubic–quintic Duffing equation [
47], the vibration of a conical beam [
48], the Mathews and Lakshmanan oscillator [
49], and the microelectromechanical system [
50].
To simplify the new analytic method, we decompose Equation (
1) as follows:
where
is a constant weight factor. Starting from the initial guess
, one solves the following linearized ODE [
42]:
to seek a higher-order analytic periodic solution. This technique is termed a decomposition–linearization method, advocated by Liu et al. [
42] as a basis to treat the analytic solution of nonlinear ODE.
When Equation (
8) is a periodic system, we can take
such that it becomes
In general, the fundamental frequency
is an unknown constant. Below, we take one example to show the process to solve Equation (
10) approximately.
We apply the linearization technique developed in [
41] to
This equation has been used to simulate the vibration of the human eardrum [
33]. The vibration model is constructed in an asymmetric form, because the radial fibers undergo a vibration of moderate amplitude toward the outside as compared to the vibration toward the inside. Since the potential function
is unsymmetric, this oscillator is an unsymmetric oscillator, one sort of the Helmholtz oscillator [
51]. The chaotic behavior of ship rolling motion in beam sea has been studied by Liu [
52], of which a typical equation to explore the instability of ship capsize is the following quadratic nonlinear oscillator:
In studies of ship motion, the analysis of large-amplitude nonlinear rolling motion is important for understanding capsize dynamics.
To find the analytic solution of Equation (
11), we apply the Lindstedt–Poincaré method [
33]; the starting point is
where
is to be determined. Let
We can derive a sequence of linear ODEs:
The analytic solution obtained by the Lindstedt–Poincaré method (LPM) is given in [
33]:
where
The details of the linearized Lindstedt–Poincaré method (LLPM) were derived in [
41]. Given an initial guess
of the function
in Equation (
11), we have
After inserting
into Equation (
20), we need to solve
in terms of
, it becomes a nonhomogeneous Mathieu-type ODE:
where
For the Mathieu-type ODE, to find analytic periodic approximations, one can refer to [
53].
Notice that and cannot be defined if . In the whole paper, we take . For the simplest harmonic oscillator , and .
By using
the analytic solution obtained by using the LLPM was derived in [
41], given by
where
in which
We compare the computed results with different
in
Table 1 to that computed by the fourth-order Runge–Kutta method (RK4). The maximum error (ME) for Equation (
25) obtained by LLPM is smaller than the ME obtained by LPM for all
as shown in
Table 1. This case reveals that the linearized technique in Equation (
21) is useful and powerful for seeking the more accurate periodic solution.
3. A New Linearized Harmonic Balance Method
To motivate the development of a modified harmonic balance method (HBM), let us consider a cubic nonlinear oscillator:
where
B is a constant.
According to the HBM, a possible second-order approximation of Equations (
28) and (
29) is the following periodic solution:
and there exist three unknown constants
,
and
to be determined. Inserting Equation (
30) into Equations (
28) and (
29) yields
By balancing the lower-order terms of
and
and neglecting the higher-order terms in Equation (
32), we can derive
Equations (33)–(35) constitute three coupled nonlinear algebraic equations to determine three unknown constants
,
and
. The procedure of HBM is quite lengthy and needs more further assumptions to derive the following solution [
33]:
whose accuracy is not good, in the order of
when
.
To overcome the drawbacks of the harmonic balance method (HBM) as mentioned in the above, and motivated by the linearized Lindstedt–Poincaré method (LLPM) in
Section 2, we develop a linearized version of HBM, namely the linearized harmonic balance method (LHBM). In
Section 4, we will apply the LHBM to solve Equations (
28) and (
29); the procedure becomes simpler and the accuracy is also raised.
We seek the analytic solution of Equation (
22) by using the HBM:
where
are unknown coefficients subject to
for the requirement to satisfy the given initial conditions
and
. The number
signifies the order of approximation. In general, we take
for the third-order analytic solution. Because HBM is applied to the linearized Equation (
22), rather than the nonlinear Equation (
12), we call the present new technique a linearized harmonic balance method (LHBM) as a modification of HBM.
3.1. A Helmholtz Oscillator
Now we seek the periodic solution of Equation (
11) by using the LHBM. Inserting Equation (
37) for
into Equation (
22) and taking the balance of harmonic terms
, we can derive
Indeed, Equation (
38) and the above equations constitute a system of linear equations for
. In the LHBM, we solve a few linear equations to determine
and
, rather than the nonlinear algebraic equations in HBM. This is a great advantage of LHBM over HBM for an easier treatment of the high-order periodic solution of nonlinear ODE.
Given an initial guess of
, we employ Equations (
38)–(
43) to determine
and
. After inserting
and
in Equation (
23) to Equations (
39)–(
43) and lifting Equation (
40) to the first one, we can derive
Equation (47) is required when
.
Equation (
44) is used to compute
, while other equations, including Equation (
38), are used to determine the Fourier coefficients
. They can be directly solved by using the Gaussian elimination method. For
, we have
To describe the iteration technique, we denote and at the j-th step values of the frequency and Fourier coefficients. For , the iteration process is summarized as follows.
- (i)
Given , , , and ,
- (ii)
Do
, solving
- (iii)
Computing
derived from Equation (
44),
- (iv)
If
then stop; otherwise, go to (ii). We cannot take
, because
appears in the denominator of
and
in Equation (
23).
We fix , and the initial guess of is given by ; the convergence is very fast, within six iterations, under the convergence criterion . The convergence behavior is not sensitive to the initial guess of , and we list the number of iterations (NI) for different values of . For , we have NI = 7; for , we have NI = 8; for , we have NI = 6; for , we have NI = 7; and for , we have NI = 7.
Table 2, for different
, compares the ME within one period to the exact solution obtained by the RK4 to integrate Equation (
11); at the same time, we compare
obtained from Equations (
19), (
26) and (
50).
Table 2 also lists the number of iterations (NI) carried out for Equation (
50). The value of
is chosen to minimize ME.
For Equation (
11), the exact value of
is given by
where
Upon comparing to
Table 1, the accuracy of LHBM is very good, competitive with that obtained by the LLPM [
41].
To further enhance the accuracy, we can also develop a two-stage LHBM, where we first raise
to the second-order solution as that derived by the LHBM:
where
,
and
are determined iteratively by
Then, we can derive the following linearized equation with respect to
in Equation (
53):
In the LHBM, we take
and derive the following linear system:
where
By means of Equation (
56), it is easy to generate
for each value of
, which is updated by
The improvement of accuracy obtained by the two-stage LHBM is shown in
Table 3, which is about one order of that in
Table 2.
In Equation (
53), if we take
and
, the two-stage LHBM is recovered to the single-stage LHBM. Indeed, the values of
are close to
,
and
. For instance,
,
, and
are obtained by the two-stage LHBM for the case of
. So the two-stage LHBM and the single stage LHBM are close to each other within the order
; the improvement of the accuracy of the periodic solution is about one order, whereas the value of the frequency
is altered a little. Unless one wants to obtain a highly precise solution, the single stage LHBM is good enough for the general purpose of the nonlinear oscillation problem to obtain a sufficiently accurate periodic solution.
3.2. A Nonlinear Damping Oscillator
Consider [
33]:
This equation considers a particle of unit mass moving in a viscous medium with a nonlinear damping term.
By starting from the initial solution
, the linearization of Equation (
60) with respect to
is
and thus,
In terms of
, we need to solve
where
The analytic solution of Equation (
60) by using the LLPM was [
41]
where
On the other hand, the third-order Lindstedt–Poincaré solution of Equation (
60) was given on page 65 of [
33]:
where
By using the LHBM, we insert Equation (
37) into Equation (
63), and take the balance of harmonic terms
to derive
where
After inserting
in Equation (
64) to Equations (
69)–(
73) and lifting Equation (
70) to the first one, we can derive
Equation (78) is required when
. When Equation (
75) is used to compute
, other equations and Equation (
38) are used to find the Fourier coefficients
.
For , the iteration process is given as follows.
- (i)
Given , , , and ,
- (ii)
Do
, solving
- (iii)
Computing
deduced from Equation (
75),
- (iv)
If then stop; otherwise, go to (ii). It is noticed that we cannot take ; otherwise, the iteration cannot be performed, since appears in the denominator.
The iteration is carried out until
.
Table 4 lists ME obtained by LHBM. Also, we compare
obtained by LPM, LLPM and LHBM with different
. Upon comparing to Table 2 in [
41], the accuracy of LHBM is very good, competitive with that obtained by LLPM.
3.3. A Kick Oscillator
The most simple non-smooth second-order ODE is
where
and
. If
, it is a harmonic oscillator with the natural frequency
. If
, it is a simple state-dependent kick oscillator [
54].
We take
with
The linearization of Equation (
83) with respect to
is
where
Taking
in the LHBM, we can solve the Fourier coefficients by applying the Gaussian elimination method to
where
.
For , the iteration process is given as follows.
- (i)
Given , , and ,
- (ii)
Do
, solving
- (iii)
- (iv)
If then stop; otherwise, go to (ii).
If we take
and
, we compare the periodic solutions in
Figure 1; they are close with ME =
. The accuracy obtained by LHBM is shown in
Table 5. The best value of
is
.
4. Duffing Oscillator
In this section, we will develop a more accurate analytic solution for the Duffing oscillator by using the linearized harmonic balance method (LHBM). Consider [
33]:
The Duffing equation describes the motion of a mass attached to a stretched wire, whose restoring force consists of a linear spring and a nonlinear spring. Many applications of the Duffing oscillator equation in physics can be seen in [
46].
We solve
where
As shown in
Figure 2, the solution obtained from Equation (
89) is very accurate, with ME smaller than
.
We write the asymptotic solution obtained by the Lindstedt–Poincaré method (LPM) given in [
33]:
where
To proceed to the higher-order analytic solution by using the LHBM, we seek the analytic solution of Equation (
89) by
where
are unknown coefficients to be determined, satisfying
due to
. Inserting Equation (
93) into Equation (
89) and taking the balance of harmonic terms, we can derive
When
, Equation (96) is needed. When
, Equation (95) can be merged into Equation (96).
We take
; by means of Equations (95), (97) and (98), we have
It is easy to generate
for each value of
, given by
For , the iteration process is given as follows.
- (i)
Given , , , and ,
- (ii)
Do
, computing
- (iii)
Computing
derived from Equation (
94) by inserting Equation (
90) for
and
,
- (iv)
If then stop; otherwise, go to (ii). Obviously, we cannot take , which would render the iteration a failure.
We take
to be the optimal value. In
Table 6 for different
, we compare ME1 obtained by LHBM within one period to the exact solution, and at the same time, we compare
obtained from Equations (
92) and (
103); the exact
is given by
Upon comparing to ME2 obtained from Equation (
91), the present periodic solutions are more accurate. Upon comparing to Table 3 in [
41], the accuracy of LHBM is very good, even better than that obtained by the LLPM.
Instead of
, we can also derive an iteration method in terms
directly. It follows from Equation (
90) that
Inserting it for
into Equations (95)–(97) generates an iteration method in terms of
:
By means of Equations (98) and (107)–(109), it is easy to compute
for each value of
given by
Inserting Equation (
106) into Equation (
94) yields
where
and
are calculated from Equations (
110)–(
112) by inserting
. The initial guess of
can be any
, say
.
It follows from Equation (
105) that
and thus, a simple formula for
is available as follows:
When the convergent value
is obtained, inserting it into the above equation, we can compute
. We found that this iteration converges very fast with a few steps.
We take
and
to be the optimal value. In
Table 7 for different
, we compare the ME obtained by the second harmonic balance method to the exact solution obtained by RK4 to integrate Equation (
88), and at the same time, we list
obtained from Equations (
104) and (
115). Comparing to
Table 6, the presented two type harmonic solutions are almost the same. Upon comparing to Table 3 in [
41], the accuracy of LHBM is very good, even better than that obtained by the LLPM.
By using
we can transform Equations (
28) and (
29) to
They are a special case of Equation (
88) with
by deleting
.
We can apply the LHBM to find the third-order periodic solution of Equations (
28) and (
29) by
We take
. Upon comparing to ME2 =
obtained from Equation (
36), the present harmonic solution with the error denoted by ME1 =
is more accurate. The improvement of accuracy is about two orders. By using the original HBM, it is hard to find the analytic periodic solutions with order greater than three, because the procedure is very complicated. In contrast, the LHBM can be easily used to find the higher-order analytic periodic solutions of nonlinear oscillators.
According to the result in [
33] the periodic solution of Equation (
88) is
The error obtained by this equation is denoted as ME of Equation (
123). As tabulated in
Table 7, the present harmonic solution with the error denoted by ME is more accurate. The improvement of accuracy is about three to four orders.
According to He’s formula [
18], we have
where
for Equation (
88). In
Table 7 we can observe that the accuracy of the frequency obtained by the Hamiltonian-based frequency–amplitude formulation is of the order
. However, owing to its simplicity, it can be applied to a more complex conservative system.
We extend Equation (
88) to
We can derive the same Equation (
89), but with
By using the HBM, it is hard to derive the frequency–amplitude formula; however, if we take
and use the following equations derived from the LHBM:
we can derive the following formula to determine the frequency:
In
Table 8, the exact value of
is compared to that computed by Equations (
124) and (
130). The improvement of the frequency obtained by Equation (
130) compared to that obtained by Equation (
124) is about five orders.
Applying the LHBM to Equation (
125) with
, we can derive the following periodic solution:
where
is given by Equation (
130).
By using the dual Lagrange multiplier approach for Equation (
125), Anjum and He [
17] can derive the following periodic solution:
where
In
Table 9, for different values of
B and a fixed
, we compare the ME obtained by Equations (
132) and (
133). The accuracy of LHBM is very good, better than that obtained by the dual Lagrange multiplier approach by about one or two orders.
Figure 3 compares the exact solution and the periodic solutions obtained by Equations (
132) and (
133) for the Duffing oscillator with
. It can be seen that the periodic solution obtained from Equation (
132) is almost coincident with the exact solution. It is apparent that the periodic solution obtained from Equation (
132) is more accurate than that obtained from Equation (
133).