You are on page 1of 22

Mechanical Systems and Signal Processing 150 (2021) 107229

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


journal homepage: www.elsevier.com/locate/ymssp

Theoretical and numerical analysis of regular one-side


oscillations in a single pendulum system driven by a
magnetic field
Adam Wijata 1, Krystian Polczyński 2, Jan Awrejcewicz 3,⇑
Lodz University of Technology, Department of Automation, Biomechanics, and Mechatronics, 1/15 Stefanowskiego Str., 90-924 Łódź, Poland

a r t i c l e i n f o a b s t r a c t

Article history: This paper presents a theoretical and numerical analysis of one-side oscillation in a single
Received 9 March 2020 magnetic pendulum. The system is composed of a physical pendulum with a neodymium
Received in revised form 19 June 2020 magnet fixed to the end of the rod. The pendulum is driven by a pulsating, repulsive mag-
Accepted 14 August 2020
netic field generated by an electric coil placed underneath. The pendulum pivot is damped
by an elastic element. The excitation current signal has a pulsating rectangular waveform
with a controlled frequency and duty cycle. In general, magnetic interaction weakens with
Keywords:
increasing distance between the magnet and the coil, making excitation in this system not
Magnetic pendulum
Double-well potential
only time-dependent but also position-dependent. The specific type of solution, referred to
Kick-excited system as ‘‘one-side oscillation” is analysed in terms of a frequency and a duty cycle of the current
Periodic solution signal. By one-side oscillations, we mean oscillations of the pendulum characterized by the
same sign of angular displacement, without passing through the lowest and the highest
vertical equilibrium positions. The analysis is based mainly on the assumption that outside
some ‘‘active zone” the influence of magnetic interaction on the pendulum dynamics is
negligible, and the system can be discretized into two states: with and without magnetic
force. The results were confirmed by experimental data showing the different periodicity
of one-side oscillations. Limitations of the proposed analysis were found in the case of
some variants of the analysed solution type. An overview of the system dynamics is pre-
sented in the form of bifurcation diagrams obtained numerically and verified by experi-
mental estimates, which show the existence of chaotic behaviour and multiperiodicity
for various values of the frequency of the current signal.
Ó 2020 Elsevier Ltd. All rights reserved.

1. Introduction

In the last two decades, there has been considerable interest in the capabilities and applications of mechatronic systems
combining two different physical branches, mechanics and electromagnetism. Such systems are found in a wide range of
applications such as stepper motors [1], linear electromagnetic motors [2], perpetual motion desktop toys [3], magnetic

⇑ Corresponding author.
E-mail addresses: adam.wijata@dokt.p.lodz.pl (A. Wijata), krystian.polczynski@dokt.p.lodz.pl (K. Polczyński), jan.awrejcewicz@p.lodz.pl (J. Awrejce-
wicz).
1
https://orcid.org/0000-0003-3042-7112
2
https://orcid.org/0000-0002-1177-6109
3
https://orcid.org/0000-0003-0387-921X

https://doi.org/10.1016/j.ymssp.2020.107229
0888-3270/Ó 2020 Elsevier Ltd. All rights reserved.
2 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

gears [4], magnetic bearings [5] and eddy current dampers [6]. One of the simplest examples is the magnetic pendulum, the
most common form of which is a pendulum with a magnet or a ferromagnetic ball attached to its free end, which interacts
magnetically with a magnet or an electrical coil placed underneath it. Studying and understanding the behaviour of such
simple systems is crucial for the effective design and control of more complex mechanical systems with many degrees of
freedom. However, most of the research literature has focused on explaining and resolving problems related to industrial
applications, and provides little insight into the fundamentals of interactions between mechanics and magnetism [7–9]. Even
a simple spherical magnetic pendulum with three permanent magnets underneath possesses complex dynamical features,
including basins of attraction with fractal boundaries [10]. The literature devoted to studying these systems usually provides
little more than a suitable mathematical model of the observed phenomena [11–15]. Moreover, autonomous systems are
rarely encountered in industrial applications. It is therefore useful to consider a magnetic pendulum with an electrical coil
powered by an alternating current, which serves as a source of excitation in the system.
The aim of this study is to investigate dynamics of the magnetic pendulum. Not only to provide mathematical description
of it and document solutions, but also to explain its origins and mechanics. In our attempt to this problem we determine the
conditions needed to obtain a specific form of motion named ‘‘one-side oscillations”. This form of motion seems to be the
simplest form of a periodic motion in the discussed system. To facilitate the research a laboratory stand was developed.
It consists of a magnetic pendulum system with an electrical coil, which is periodically switched on and off. Based on the
derived mathematical model and numerical simulations, coil switching points in phase space for one-side oscillations were
computed. Finally, the constant frequency and duty cycle of the switching signal were determined. To the best knowledge of
the authors, this is the first study to investigate the conditions needed for a specific required motion of a magnetic pendu-
lum. In particular, no previous investigations have considered regular one-side oscillations in a single pendulum system dri-
ven by a magnetic field.
The first experimental investigation of a mathematical pendulum with a small ferromagnetic ball at the end of the rod
was reported by Bethenod in [16,17], who observed the sustained and undamped oscillations of a pendulum subjected to
an alternating high-frequency magnetic field. The magnetic field was generated by a coil orientated in parallel to the pen-
dulum rod. Bethenod proposed a theoretical solution of the linearised system dynamics taking into account two coupled dif-
ferential equations of the pendulum motion and the electrical circuit. However, the solution was incomplete because it
explained neither the mechanism nor the conditions under which the oscillations were sustained rather than damped by
the magnetic field [18]. Later, Rocard [17,19] and Knauss et al. [18] theoretically analysed Bethenod’s pendulum using power
series approximations. Their results were valid only for very small amplitudes of pendulum oscillation. For larger ampli-
tudes, the problem becomes non-linear, due to the influence of higher-order corrections to the value of the current and
because of deviations from the linear dependence of the self-inductance of the circuit. This problem was taken up by Min-
orsky [20], who artificially reduced the problem to the Mathieu differential equation with a moving parametric point. This
intuitive assumption based on the Mathieu function gave a qualitatively correct result, and the limitations of the amplitude
of the oscillatory process were described. Conditions for the occurrence of self-excitation, periodic oscillation with stationary
amplitude and no stable stationary oscillation were also explained.
Kesavamurthy et al. [21] inspired by Minorsky studies derived their own improved differential equation of the pendulum
motion and confronted the theoretical results with experimental data. The accuracy of the proposed theory was commen-
surate with the approximations and assumptions involved. Further, a similar setup of the magnetic pendulum has been
investigated theoretically, numerically, and experimentally by Khomeriki [11]. Magnetic interaction in his studies was
obtained from the energy law. Despite that, after some linearization of the governing equation, the motion of the pendulum
has been governed by the Mathieu equation with a damping factor. Analysis led to the conclusion that parametric instability
implies chaotic behaviour. The limits for the existence of parametric resonance and its dependence on system parameters
were defined, but were validated only for small pendulum angles. The limits of the parametric resonance for larger angles
were estimated by numerical computations.
A few researchers studied so-called parametric electrodynamical machines which mathematical description is similar to
the magnetic pendulum. Kaplan [22] elaborated the general approach to describing the motion of devices such as VHF res-
onator, levitator and rotary parametric motor which are classified as parametric electromechanical machines. The tendency
of the systems to sustain their own vibrations has been observed. Smith et al. [23] analysed analytically and experimentally
the behaviour of the brushless parametric rotary reluctance machines. A single 2nd-order differential equation with periodic
coefficients has been employed to describe the motion of the systems. Approximate solutions obtained analytically repre-
sented the major effects within studied machines. Moreover, it has been shown that the efficiency of these machines will
be less than 50%. The mechanism of torque production in parametric reluctance motor has been investigated by Russel
et al. [24]. Furthermore, the condition necessary for the parametric machine to function has been explained. Analysis has
shown that the torque generated by the parametric motor has been much less than by a classical synchronous reluctance
motor. Performed studies indicated that the attenuation of the upper sideband currents and amplification of the lower side-
bands currents should improve the efficiency of the parametric motor. A tuned circuit electromechanical oscillator has been
analysed by Blakley [25]. The developed oscillator can be used as a linear electric motor to oscillate small objects over long
distances. The dynamics of the system has been studied numerically and analytically. The analytical approach was based on
the approximate phase-plane analysis. The system motion has been divided into four distinct regions and then each of them
has been expressed by an autonomous differential equation.
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 3

A novel approach to modelling a pendulum system which provided new information about its self-oscillations was pre-
sented by Skubov et al. [12]. The system was modified slightly, by replacing the ferromagnetic ball with a wire closed loop.
Their theoretical analysis was based on an asymptotical solution of Lagrange-Maxwell equations. The results demonstrated
that the motion of the system converges to one of two equilibrium positions, or to limit cycles situated in their vicinity. Fur-
thermore, these limit cycles were conformed to self-induced oscillation, exhibiting frequencies smaller than the frequency of
magnetic excitation. A different approach was proposed by Damgov et al. [26]. The system was termed ‘‘kick-excited” due to
the type of excitation, which was dependent on time as well as on the position of the pendulum. Under the inhomogeneous
action of a periodic force, the pendulum displayed ‘‘discrete/quantized” amplitudes of oscillation, which were obtained ana-
lytically and numerically. Theoretical analysis was based on the energy balance of the system and supported by numerically
obtained two-dimensional maps. Cintra et al. [27,28] studied experimentally and numerically six different spatial magnetic
pendulums under external harmonic magnetic field. They modelled the systems as Duffing oscillators. The frequency of the
driving force employed to the systems was far from the natural frequency of the oscillators. Under that certain conditions,
the phenomenon of the argumental oscillations have been observed, i.e. stable regimes of pendulum motions have been
noted.
Amplitude quantization was also observed in Duboshinski’s pendulum [17,29–31]. In contrast to Bethenod’s pendulum,
Duboshinsky changed the orientation of the coil placed under the pendulum. When the pendulum deviation angle was zero,
the axis of the coil was perpendicular to the pendulum rod. Several stable ‘‘nearly periodic motions” were obtained for each
fixed set of the excitation parameters, and these could be achieved from any initial pendulum deviation. Stable and unstable
values of the steady-state amplitudes for the different amplitudes of the external force took the form of nearly periodic
sequences of lobes corresponding to different values of possible amplitudes. The subharmonic frequency response in
Duboshinski’s pendulum has been investigated by Luo et al. [13]. The amplitude quantization has been identified as subhar-
monic resonance in response to nonlinear feeding. The relation between subharmonic resonance frequency and symmetry of
the driving force was found.
Later researchers slightly altered the Bethenod pendulum, by multiplying the number of magnets and/or coils, as well as
their localization on the pendulum arm. Some also added mechanical sources of excitation to the system. One study inves-
tigated a pendulum with two magnets fixed at its ends, driven by one or two coils placed underneath and/or above it [32,33].
An experiment with positive and negative feedback control was carried out, and the equation of motion presented in a form
similar to the Van der Pol equation. The parametric oscillations and multi-well potential were also studied experimentally.
The dynamics of a magnetic pendulum driven by a motor were studied in detail by Tran et al. [34] and Nana et al. [35,36].
The first paper was mostly focused on finding the coefficients of the magnetic pendulum as it interacted with another mag-
net fixed underneath it and driven by a specially constructed motor. The numerical calculations showed that the coefficients
that were found for one excitation frequency may underestimate the uncertainty in the system excited by other frequencies.
The experimental and simulated bifurcation diagrams displayed periodic (mostly period-1 and period-2) and chaotic regions.
Nana et al. [35,36] employed two fixed magnets and placed them symmetrically on either side of the pendulum magnet.
Their analytical, numerical, and experimental studies revealed a wider range of regular oscillations than in [34] from
period-1 to period-8.
Instead of rotational motors, harmonic parametric excitation was introduced in works [37–39]. Parametric excitation was
applied to the point of pendulum suspension. Siahmakoun et al. [37] analysed a magnetic pendulum with horizontal para-
metric excitation and fixed one magnet under its arm. They noticed typical non-linear features, such as amplitude jumps,
hysteresis, and bistable states, which were also reported in [32,33,35]. Mann [38] went one step further, and used two mag-
nets to construct a Duffing-type pendulum oscillator. Mann analysed the energy criterion for potential well escapes. The
maximum energy of the steady-state oscillation required to overcome the adjacent potential barrier was predicted by ana-
lytical and numerical calculations. Furthermore, the escape induced the transition of periodic motion to the chaotic regime.
Inspired by Mann’s work, Kwuimy et al. [39] decided to investigate the same system but with the setup inclined, which
caused a significant potential symmetry breaking. The effect of tilted parametric excitation and parametric damping was
investigated using the Melnikov method to derive the analytical condition for chaotic motion. The stability of the system
(the domain of regular motion) increased when the tilting excitation approached the vertical position.
Another approach to modelling magnetic pendulum systems using modified Duffing oscillators was reported by Moon
et al. [40] and further developed by Donnagáin et al. [41]. In both works, a ferromagnetic beam played the role of a pendu-
lum, which was buckled between two magnets that underwent forced oscillations. Moon et al. simplified the equation of
motion by taking a single degree of freedom Galerkin approximation and explained why it yields reasonable results. They
also discussed the existence of a strange attractor. The purpose of the second paper was to analyse how magnetization affects
the dynamics of the system. The Duffing equation incorporated a Preisach nonlinearity that involved the ‘‘memory” of the
nonlinearity and, in effect, acted as an additional damping term.
Systems consisting of a single magnetic pendulum have been also considered as devices for generating electrical energy
[42,43]. The external forces that induce the pendulum oscillations can be converted into electrical energy inside coils by elec-
tromagnetic induction. These systems can work under both small and large amplitudes of external driving forces, for exam-
ple from a natural environment.
A number of studies have focused on magnetic pendulum systems with more than one degree-of-freedom. The bifurca-
tion dynamics of a chain of two coupled magnetic pendulums under a pulsating repulsive magnetic field were investigated in
references [44,45]. Control of two axially-coupled magnetic pendulums was explored in [46]. The numerical and theoretical
4 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

analysis was devoted mostly to synchronization phenomena. The motion of a double-linked magnetic pendulum was studied
in detail in [47]. A complex model of magnetic interaction was employed to simulate bifurcation diagrams and trajectories
that were in good agreement with experimental results. Systems of triple magnetic pendulums were studied in papers
[7,48]. The first paper exploited the magnetic field to dampen vibrations in the system, whereas in the second paper chaotic
motion were reported. Numerous papers have also been aimed at energy harvesting in multi-degree-of-freedom systems,
composed of a magnetic pendulum in the vicinity of electric coils [49–54]. The advantages and disadvantages of such sys-
tems in comparison to other harvesting devices were discussed, as well as their applications.
This paper presents the construction (Section 2) and system modelling (Section 3) of a magnetic pendulum. Section 4 pro-
vides an overview of its bifurcation dynamics. Its one-side oscillations are analysed theoretically, numerically, and experi-
mentally in Section 5. The results are discussed in details in Section 6.

2. Experimental setup

Fig. 1 shows the experimental rig, with the pendulum marked as (1). The arm of the pendulum is built of an aluminium
alloy part and a textolit (composite material) rod. The free end of the pendulum is equipped with a strong neodymium mag-
net (2) 22 mm in diameter and 10 mm in height. The other end of the pendulum is fixed to the brass shaft and marked as (3).
The shaft is supported by two small rolling bearings. One end of the shaft is connected to a rubber elastic element (4). The
elastic element has a rectangular cuboid form 40:3 mm in length, 8:5 mm in height, and 6:5 mm in width. It is also fixed at
the other end to a FUTEK TFF325-FSH04055 stationary reaction torque sensor, marked as (5). An electric coil (6) is placed
under the pendulum arm and their axes coincide when the pendulum is in the lower equilibrium position. The properties
of the coil are as follows: inductance 22 mH, resistance 10:6 X, diameter of the copper wire 0:5 mm, external diameter of
the coil 40 mm, coil bore diameter 17 mm, coil height 31 mm. The air gap between the neodymium magnet and the electric
coil is fixed at 1:6 mm. A polymer disk (7) with 79:6 mm diameter is attached to the shaft and it is a part of the magnetic
interaction measurement system (for more details see Appendix A). The materials were selected with a view to reducing
unwanted magnetic influence on the magnet and coil, so are mostly non-magnetic. The frame (8) is made of aluminium pro-
files joined together by brass bolts. The stand may be easily reconstructed as a stand for a chain of pendulums (maximum
three), joined axially as described in [44]. The angular position of the pendulum is registered by an HEDS-9040#J00 optical

incremental encoder (9) and the code wheel (10) with 1000 increments per rotation, giving a resolution of 0:36 . The electric
coil is attached to a KORAD KA3005D DC power supplier working as a current source. The current was set at a steady value of
1 A and was controlled by a self-made electronic unit based on transistors. The unit can open or close the current flow
according to the rectangular signal delivered from the SIGLENT SDG1025 signal generator. The direction of the current in
the coil provides repelling interaction between the neodymium magnet and the coil. The angular displacement of the pen-
dulum obtained from HEDS sensor and the excitation signal produced by SIGLENT generator have been acquired by a NI USB-
6341 device, and then the data was processed using the LabVIEW program on a PC class computer.

3. Physical and mathematical models

A physical model of the investigated system is shown in Fig. 2. Like a physical pendulum, it is characterized by a moment
of inertia J ½kgm2 , mass m ½kg, and the centre of mass location s ½m. The motion of the pendulum is constrained to only one
rotation, around the x axis. The angle of rotation is denoted by u ½rad. External torques acting on the pendulum are associ-

Fig. 1. The experimental stand of the magnetic pendulum: (1) pendulum; (2) neodymium magnet; (3) brass shaft; (4) elastic rubber element; (5)
tensometric torque sensor; (6) electric coil; (7) polymer disk; (8) aluminium frame; (9) optical encoder; (10) code wheel.
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 5

Fig. 2. Physical model of the studied magnetic pendulum.

ated with the friction in the ball bearing M F ðu


_ Þ ½Nm, elastic element deformation M K ðu; u
_ Þ ½Nm and magnetic interaction
Q m ðu; t Þ ½Nm.
Based on the physical model, a mathematical model is derived from Newton’s second law with the form
Ju
€ ¼ mgs sin u  MF ðu
_ Þ  M K ðu; u
_ Þ þ Q m ðu; tÞ ð1Þ
where g ½m=s2  stands for gravitational acceleration equal to 9:81 m=s2 . The presented model is non-autonomous from a
mechanical point of view [55,56], because it contains forcing torque Q m ðu; tÞ which explicitly depends on independent time
variable t.
As it was repeatedly reported by many researchers [57,58], a typical journal bearing friction characteristic follows the so-
called Stribeck curve (see Fig. 3). Starting from zero velocity due to the Stribeck effect friction torque decreases from static to
kinetic (Coulomb) friction in a continuous manner. Then a friction torque reaches its minimum value and starts to grow with
velocity, what is often called the viscous friction. For the purpose of this study, we decided to neglect the potential dynamic
character of friction and described it with a static Stribeck curve. Over the years several formulas have been proposed to
describe the Stribeck curve. The most common one uses the exponential function [59] and is expressed by
  
u
_2
M F ðu
_ Þ ¼ M c þ ðM s  Mc Þ exp tanh vu
_ þ cu
_ ð2Þ
vs
where M c ½Nm is the magnitude of Coulomb (kinetic) friction, v s ½rad=s stands for Stribeck velocity and defines how quickly
friction value decreases from static to kinetic friction. M s ½Nm is the static friction value and c ½Nms=rad stands for viscous
friction coefficient. Because we approximated the signum function with a hyperbolic tangent function with the regulariza-
tion parameter v, the M F ðu_ Þ curve does not reach exactly M s value for small angular velocity. The parameters of Eq. (2) were
identified numerically, based on the recorded free motion of the pendulum (see Table 2 and Appendix B).
We assumed that the reaction torque of an elastic joint is a linear function and has the form

Fig. 3. Typical journal bearing friction torque characteristic M F ðu


_ Þ known as the Stribeck curve.
6 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

MK ðu; u
_ Þ ¼ ke u þ c e u
_ ð3Þ

where ke ½Nm=rad is the stiffness of the joint and ce ½Nms=rad is a viscous damping coefficient. Measurements of torque
dependence on angular deformation were performed. The experimental data and the torque were compared accordingly
to Eq. (3), as shown in Fig. 4. One can observe that the applied linear formula does not describe the recorded experimental
data correctly. It seems that some nonlinearity and/or hysteresis effects may appear inside the elastic element. Since some
discrepancy between the results is visible but the overall character is similar, we decided to use the linear model of the elas-
tic element in our further investigations.
Magnetic interaction torque Q m ðu; t Þ depends on static characteristic M m ðuÞ ½Nm and the switching signal iðtÞ½:

Q m ðu; t Þ ¼ iðtÞ  M m ðuÞ ð4Þ


What is an unusual condition because the excitation torque depends on time as well as the angular position of the system.
Actually, this torque influences the system dynamics significantly, only in a limited range of the angular position u. Outside
this area, its contribution is negligible. We called this range the ‘‘active zone” [26,31] and denoted by uA (see Fig. 5). We
defined the boundary of the active zone with the equation
V s ðuA Þ
¼ 0:999 ð5Þ
V s ðuA Þ þ V m ðuA Þ

where V s ðuÞ ½J stands for the potential energy of the system yielded from the gravitational field and the elastic element char-
acteristic. V m ðuÞ ½J is the potential energy of the magnetic interaction.
As far as magnetic interaction is concerned, the diameters of the magnet and the coil in our stand are probably too large to
simplify the interaction between them to a pair of magnetic poles, since this approach failed to fit our experimental data
[44]. Instead, we decided to use a mathematical description based on experimental measurements. In order to find an appro-
priate mathematical formulation, we looked at the potential energy of the system, where magnetic interaction forms a peak,
dividing the initial one-well gravitational potential into a double-well potential (as shown in Fig. 5a). Of the different peak-
shaped functions (Lorentzian [60], Pearson VII [61], Pseudo-Voigt [62,63]), the Gaussian [60] function was chosen to describe
the potential energy of the magnetic interaction
 
u2
V m ðuÞ ¼ a exp  ð6Þ
b

where coefficients a ½Nm and b ½rad are constant for the fixed current amplitude. This idea of modelling is different from
that presented in [38–41], where the potential energy of the system was expressed by a polynomial function. Our approach

Fig. 4. (a) Experimentally recorded characteristics of the elastic element and (b) torque calculated based on Eq. (3).

Fig. 5. The one-well and double-well potential energies of the system (a); the experimental measurement and approximation of the torque Mm for 1 A of
the coil current (b). The highlighted range between angles uA indicates the so-called ‘‘active zone”.
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 7

differs in that, it enables the mechanical and magnetic potential energy to be distinguished. According to the mechanical
definition of the potential, one can calculate the torque M m ½Nm acting on the pendulum for steady coil current as
 
@V m 2a u2
Mm ðuÞ ¼  ¼ exp  u ð7Þ
@u b b
Clearly, besides V m ðuÞ the system is also subjected to the potential energy accumulated in the elastic element and the
gravitational. The sum of these two potential energies reads
1
V s ðuÞ ¼ mgsð1  cos uÞ þ ke u2 ð8Þ
2
Fig. 5b shows the experimental measurement and the approximation (Eq. (7)) of the torque M m for 1 A of coil current.
Switching signal iðtÞ (from Eq. (4)) is modelled in two ways: continues and discrete.

3.1. Continuous modelling

The electrical coil in our stand is powered with a pulsating current signal. In general, the current change in electrical cir-
cuits with inductance is described with a differential equation. Nevertheless, we decided to model it with a static function of
time, since the coil time constant is much smaller than the period of the pendulum oscillation. Moreover, the coil is powered
by an electric power supply which controls the current level.
Pulsating current signal in our stand has a form of a rectangular wave. The general idea behind the mathematical descrip-
tion of the rectangular wave is based on calculating the signum of the sine wave, particularly
1
iðtÞ ¼ I0 ð1 þ signðsin ðxi ðt þ t 0 ÞÞ þ i0 ÞÞ ð9Þ
2
In this formulation xi ½rad=s controls the pulsation frequency, i0 ½ allows changing the signal duty cycle, t 0 ½s sets the
initial phase of the sine wave and I0 ½ is the pulsation amplitude. Fig. 6 shows the idea behind Eq. (9) and defines the
parameters i0 and t0 . For numerical simulations, a smooth and continuous function is prepared by replacing the signum func-
tion with a hyperbolic tangent approximation
1
iðtÞ ¼ I0 ð1 þ tanh ðe sin ðxi ðt þ t 0 ÞÞ þ i0 ÞÞ ð10Þ
2
where e [–] is a regularization parameter. To express the formula (10) by the frequency f ½Hz and the duty cycle w ½% of the
rectangular signal, the following substitutions are made:
    
1 2pw 1 2pw
xi ¼ 2pf ; i0 ¼  sin p ; t0 ¼ p ð11Þ
2 100 2pf 100
The initial phase of the sine wave xt0 is calculated in a way to preserve the initial phase of the resulting rectangular sig-
nal. The rectangular wave always starts in the middle of the rising edge.
Since in our studies we employed Eq. (10) to describe switching on/off of the coil, the pulsating amplitude I0 is set to 1.

3.2. Discrete modelling

The electrical coil circuit time constant is over 400 times smaller than the pendulum oscillation period. This fact enables
us to simplify the description of the system dynamics, by assuming that the coil current changes its value instantaneously. In
other words, two states of the system can be distinguished: (1) when the coil is switched off and (2) when it is switched on.
Such a system is described with two different ODEs, one for each state. The first models the free motion of the pendulum in a
single-well potential, when the coil is switched off iðt Þ ¼ 0 (state-1)

Fig. 6. The idea behind the model of the coil current rectangular wave.
8 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Ju
€ þ mgs sin u þ M F ðu
_ Þ þ MK ðu; u
_Þ¼0 ð12Þ
The second describes free oscillations of the pendulum in a double-well potential when the coil is switched on iðt Þ ¼ 1
(state-2)

Ju
€ þ mgs sin u þ M F ðu
_ Þ þ MK ðu; u
_ Þ  M m ðuÞ ¼ 0 ð13Þ
Both states are depicted in terms of their potential energy in Fig. 7.
Switching between Eqs. (12) and (13) happens according to the period T ¼ 1f and the duty cycle w of the given switching
signal (see Fig. 8a):

stateð1Þ ! ð2Þ for t ¼ kT; k 2 N;

w
stateð2Þ ! ð1Þ for t ¼ T þ kT; k 2 N: ð14Þ
100%
With regard to the continuous model, the initial state of the system is state-2. From a dynamical point of view, state-1 and
state-2 are different in terms of fixed points. The fixed points represent the equilibrium solutions of the system [55,56,64].
When the coil is switched off (state-1), the system has one stable fixed point ðu; u_ Þ ¼ ð0; 0Þ, whereas in the second state this
fixed point becomes unstable (saddle) and two other stable fixed points appear. Hence, one can think of the dynamics of the
system in terms of a periodically changing fixed points arrangement, as shown in Fig. 8b. Basins of attraction for two stable
fixed points in the state-2 are pictured in Fig. 8c.
The discrete model consists of two autonomous states, but the whole system is non-autonomous because switching
between two states depends on an independent variable – time t.

3.3. Parameter identification

The excitation signal (see Eq. (10)) and magnetic interaction (see Eq. (7)) parameters were taken directly from fitting the
equations to the measurements (see Fig. 5b and Appendix A) and they are shown in Table 1.
The parameters of the pendulum system were found by comparing the simulation results to experimentally recorded
time series of the free motion of the pendulum. Identification took place in phases. First, the pendulum supported on the
ball bearing was identified. Then, the elastic element was added to the experiment and to the model. The parameters were
computed using Wolfram MathematicaÒ software and its NonlinearModelFit function, which is basically a function for finding
a local minimum in the optimization process [65]. Initial values of the parameters for the fitting process have been obtained
experimentally (see Fig. 4 and Appendix B). Identified system parameters are shown in Table 2.
Fig. 9 presents a comparison of the experimentally recorded and simulated time series with identified parameters.

Fig. 7. Discrete system states represented by potential energy plots: (a) state-1 (coil switched off), (b) state-2 (coil switched on).

Fig. 8. Switching signal between discrete system states (a), t1 – the coil is switched off, t2 – the coil is switched on; a polar plot of fixed points bifurcation
(b), solid curves stand for stable solutions, whereas the dashed curve stands for an unstable solution; basins of attraction are plotted for state-2 (c), black
dots - stable fixed points and white dot – unstable.
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 9

For the same excitation configuration, the simulation results of both models (continuous and discrete) are very close and
almost indistinguishable. The continuous model is computed more quickly in numerical simulations, whereas the discrete
model facilitates system behaviour analysis.

4. Bifurcation analysis

In order to find the dynamical behaviour of the system with regard to changes in frequency, bifurcation analysis was per-
formed in which frequency played the role of a control parameter. First, bifurcation dynamics were estimated based on
experimental results. The coli current was set to 1 A and the excitation signal had a fixed duty cycle w ¼ 30%, whereas
its frequency increased (or decreased) linearly over time, in what may be described as a sweep signal. Exemplary time his-
tories of the signal are shown in Figs. 10a and 11a. Points visible in the experimentally estimated bifurcation diagrams cor-
respond to the position of the pendulum in the time instance of a rising edge of the excitation signal (see Figs. 10b and 11b).
Next, we recreated the experiment by means of a numerical simulation. The results of the simulations are shown in Figs. 10c
and 11c, for increasing and decreasing frequencies, respectively. Finally, we computed the bifurcation diagrams, as displayed
in Figs. 10d and 11d. The algorithm used for numerical computation of the diagram follows a stable solution, i.e. initial con-
ditions for simulations for consecutive frequencies are taken from the results of the previous simulations. In each simulation,
1200 periods of the excitation signal were computed. First, 1000 periods were treated as a transient motion. The points from
last 200 periods were recorded and plotted in the diagrams. In all cases, the range of the control parameter was ð0:5  6ÞHz.

Table 1
Parameters of magnetic interaction and excitation
signal.

Magnetic interaction Value


a 4:253  102 Nm
b 1:818  102 rad
Excitation signal
I0 1 
e 200 
f variable Hz
w variable %

Table 2
Identified system parameters.

Mechanical system Value


J 6:787  104 kgm2
mgs 5:800  102 Nm
ke 1:742  102 Nm=rad
ce 1:282  104 Nms=rad
Friction
Mc 2:223  104 Nm
Ms 4:436  104 Nm
vs 5:374  101 rad=s
c 7:369  105 Nms=rad
v 5:759 

Fig. 9. Time series computed numerically and recorded experimentally for continuous (a) and discrete (b) models.
10 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Fig. 10. Bifurcation dynamics for increasing control parameter f : a) the frequency sweep signal, b) experimentally estimated diagram, c) numerically
estimated diagram, d) the numerically computed bifurcation diagram.

In the experimental and simulation diagrams, the estimated frequency changed at a rate of 25  104 Hz s
, whereas the bifur-
cation diagrams were computed with a frequency increment of 0:005 Hz.
Bifurcation diagrams inform about solution periodicity in reference to the excitation signal. One point in the diagram for a
given frequency means, that the solution has the same period as the excitation (e.g see f = 3.5 Hz). Two points indicate that
the solution period is two times greater than the period of excitation, which is called a period-2 solution (e.g see f = 2.25 Hz).
Analogically, other periodicities are denoted by ‘‘period-n‘‘. On the other hand, high points count (like for f = 1.4 Hz) is likely
an indication of chaotic motion.
The system exhibits different kinds of motion, chaotic and regular. The regular motion of the system itself possesses var-
ious features of periodicity (period-1, period-2, period-4, period-6, period-10). The numerically computed diagram estimates
reveal almost all the dynamic features observed experimentally. The experimentally computed points of bifurcation are also
captured correctly with respect to the experiment. Bifurcation analysis reveals the existence of coexisting attractors. For
example, in Figs. 10b and 11b, for a frequency of 2:25 Hz the motion of the system is period-2 for both increasing and
decreasing frequencies, but the solutions are different. The type of attractor depends on the initial conditions. Some poten-
tially chaotic zones visible in the estimated diagrams may appear because of long transitional motion, which was not omit-
ted. However, this issue does not affect the simulation diagrams in Figs. 10d and 11d. For example, the regular period-2
solution for f ¼ 2:85 Hz predicted in Fig. 11d was verified experimentally and is shown as a phase portrait in Fig. 12b.
Almost all the dynamical features observed experimentally are present in the numerical bifurcation. Those missing were
found after changing the initial conditions of the simulation (see the frequency range ð4  4:5Þ Hz in Fig. 10d and
ð5  5:5Þ Hz in Fig. 11d). Exemplary cross-sections of the bifurcation diagrams obtained numerically and experimentally
are shown in Fig. 12. Since our stand can measure only the angular position of the pendulum, we need to differentiate
the recorded displacement time series numerically in order to plot the phase trajectories.
The Poincaré sections were also computed during signal post-processing. This is why their locations may be slightly inac-
curate. This effect can be observed e.g. in Fig. 12d, where two of the Poincaré points should be before the rapid decrease in
velocity (which occurs due to magnetic field activation). These Poincaré points were taken for the time instance when the
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 11

Fig. 11. Bifurcation dynamics for decreasing control parameter f : experimentally (b) and numerically (c) obtained bifurcation diagrams estimates for a
frequency sweep signal (a) and numerically computed bifurcation diagram (d).

Fig. 12. Numerically (a, c) and experimentally (b, d) obtained phase trajectories and Poincaré sections for duty cycle w ¼ 30%.

coil was switched from ‘‘off” to ‘‘on”. In supplementary data available on-line, videos can be found showing experimental
runs relating to Fig. 12.
Phase plots are a common tool for visualizing periodic orbits. An established practice for non-autonomous systems is the
incorporation of Poincaré section points into a phase plot, to indicate the periodicity of the solution. Sections were made for
every time instance when the coil was switched from ‘‘off” to ‘‘on”. Since in our system the excitation is not typical, because
it depends on the angular position as well as on time, we wanted to provide more information by using phase plots. When
the coil is turned on, a potential energy barrier is created, which can be treated as a kind of obstacle to the pendulum. For
systems with impacts, an obstacle is often indicated in the phase plot and by an analogy we have indicated the active zone
uA with a light-green colour. Because the barrier is not always present in our system, the system states are differentiated
with a phase trajectory line colour: blue when the coil is switched off and red when it is switched on. In addition, over the
active zone the basins of attraction of the stable fixed points are illustrated with a darker colour, which depicts the barrier
shape. Such improved phase plots are shown in Fig. 13 and will be used later in our study.
12 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Fig. 13. Periodic solutions for a duty cycle fixed to 30% and excitation frequency equal to 1:16 Hz (a, b) and 5:85 Hz (c, d): phase plots – (a), (c) and system
potential energy plots – (b), (d).

Three basic system behaviour scenarios can be deduced in the active zone:

(i) when the coil is switched off, the solution goes freely across the active zone (see Fig. 13a and b);
(ii) when the coil is switched on and the system has enough kinetic energy, it can go through the potential energy barrier
created by the magnetic field (see Fig. 13c and d);
(iii) when the coil is switched on and the system does not have enough kinetic energy, it bounces off the potential energy
barrier (see Fig. 13a and b).

Another important observation is that any significant amount of energy can be added to the system only when the coil is
switched on and the pendulum is within the active zone.

5. One-side oscillations

5.1. Scenario I of the period-1 solution

Bifurcation analysis showed that the excitation signal frequency has a significant influence on the system dynamics, and
that the solutions were not unique for a given frequency and duty cycle. The analysis not only shed light on the system beha-
viour, but also raised new questions. An obvious question seems to be: which signal parameters do we need for a given
solution? We tried to answer this question for a period-1 solution observed for frequencies ranging between 3 Hz and
4 Hz (see Figs. 10 and 11).
We considered the basic scenario of the solution which is shown in Fig. 14. The cycle starts at the point ðu0 ; x0 Þ, when the
system is in the first state (the coil is switched off – blue curve). When the trajectory is inside the active zone, the systems

Fig. 14. The analysed scenario I of the period-1 solution: potential energy plot (a) and phase plot (b).
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 13

state switches to the second state (the coil switches ON – red curve) at the point ðuk ; xk Þ. We call this point ‘‘the kick”
because the system is kicked into the higher energy state. Next, the trajectory bounces off the potential energy barrier
and goes back to the initial point ðu0 ; x0 Þ, where the system returns to the first state. Since we have chosen ðu0 ; x0 Þ as
the reversal point, velocity x0 ¼ 0.
For the cycle to continue, the state-1 system goes to ðuK ; xK Þ point after a time t 1 and should return to the initial point
ðu0 ; x0 Þ after a time t2 in state-2:
fu; u
_ g1 ¼ f 1 ðu0 ; x0 ; t1 Þ ¼ fuk ; xk g;

fu; u
_ g2 ¼ f 2 ðuk ; xk ; t2 Þ ¼ fu0 ; x0 g ð15Þ
Possible kick point location in phase space ðu; u
_ Þ is restricted by the solution scenario and the system characteristic (see
Fig. 15a):

Limitation 1. the kick must be within the active zone, uk  uA (solution scenario limitation);
Limitation 2. since it was assumed that the system switches from state-2 to state-1 outside the active zone, the minimal
value of u0 is uA (limitation curve is the phase trajectory of the state-1 system from initial condition ðuA ; 0Þ);
Limitation 3. after the kick, the system should not escape from the potential energy well (this means that, from a dynam-
ical point of view, a stable manifold of ð0; 0Þ saddle in state-2 system is the limit).

Sets of fu; u
_ ; u0 g for both states of the system were computed (see Fig. 15b). Their common part determines all possible
kick points ðuk ; xk Þ (see Fig. 15a) and corresponding u0 .
For every kick point, times t1 and t 2 were computed (see Fig. 14). Knowing how long each state lasts during the whole
cycle of the solution, one can simply calculate the excitation signal parameters:
1 t2
f ¼ ½Hz; w ¼ ½%: ð16Þ
t1 þ t2 t1 þ t2
Excitation signal parameters calculated accordingly to Eqs. (16) are plotted in Fig. 16a with a blue curve. As mentioned,
the influence of the magnetic field outside the active zone is negligible, which suggests that the system can switch from
state-2 to state-1 anywhere outside the active zone, without affecting the periodical trajectory. Duty cycle w can be reduced
to a value, when the system switches just after leaving the active zone, or it can be increased to make the switch just before
entering the active zone. Adequately calculated boundaries for a duty cycle parameter are shown in Fig. 16a. As a result, the
set of admissible excitation signal parameters for so-called one-side oscillations is obtained. By one-side oscillations, we
mean oscillations of the pendulum characterized by the same sign of angular displacement, without passing through the
lowest and the highest vertical equilibrium positions. From the left, the set ends sharply at a frequency around 2:85 Hz,
whereas from the right side the set ends with a single point, which is approached smoothly. The left side sharp end results
from the system characteristics, namely from the potential energy peak value. Since the motion amplitude increases with
decreasing frequency of excitation (see Fig. 16c), the pendulum has enough energy to go across the potential energy barrier
and escape the well. On the other hand, the assumed scenario limitations are the cause of the smooth right side end of the
set. Specifically, the analysed scenario assumes that the trajectory reversal point ðu0 ; 0Þ is outside the active zone, so the
minimal motion amplitude approaches uA (see Fig. 16c). Solutions with an amplitude smaller than uA are discussed in
Section 6.
We validated the set of configuration parameters presented in Fig. 16a using experimental and numerical methods. The
numerical results shown in Fig. 17a prove the kick point locations calculated earlier. Experimental measurements were per-
formed for the following configurations of the excitation signal ff ; wg ¼ f3:5; 80g; f4:0; 80g; f4:46; 71g. The results are pre-
sented in Fig. 17b, together with the simulation results. However, we were unable to obtain one-well oscillations for
frequencies lower than 3:063 Hz experimentally. The probable reason for this is the vertical compliance of the support,
which is not included in our mathematical models. It turns out that for motion with high amplitude and velocity, after

Fig. 15. a) The location of kick points in the limiting area (boundary curves 1, 2, 3 are defined in the text above); b) sets of fu; u
_ ; u0 g for both states of the
system (state-1 – blue, state-2 – red). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)
14 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Fig. 16. Excitation signal parameters for the analysed solution type: the calculated set of admissible excitation signal parameters (a), kick point location
change (b), and motion amplitude u0 change over frequency (c).

Fig. 17. Simulation (a) and experimental (b) validation for the following fixed parameters ff ½Hz; w ½%g: (1) ! f4:46; 71g, (2) ! f4:0; 80g, (3) ! f3:5; 80g,
(4) ! f3:1; 30g.

the kick the support system (which includes a brass shaft, bearings housing and roll bearings) deflects, making it possible for
a pendulum to go over the peak potential energy value. Videos showing experiments related to Fig. 17 can be found in the
supplementary data available on-line.
It should also be mentioned that our system has many coexisting solutions (as can be observed e.g in the bifurcation dia-
grams Figs. 10 and 11). The presented numerical results were computed from initial conditions close to the predicted orbits,
and during experimental verifications we also tried to guide the system towards the desired trajectory.

5.2. Multiperiodic regular one-well oscillations

In our discussion of the motion process in scenario I (Section 5.1), we saw that switching on the coil when the pendulum
is outside the active zone does not affect the phase trajectory. This assumption suggests that it is possible to multiply the
periodicity of motion and maintain the same phase trajectory. We used the extremal values for the frequency and duty cycle
obtained from scenario I to derive the frequencies and duty cycles of the multiperiodic solutions based on the previous anal-
ysis. Fig. 18 shows the time history plot of three excitation signals iðtÞ. The amplitude of the signals has been artificially chan-
ged for better visualization of their different duty cycles. The red curve corresponds to iðtÞ with the minimum duty cycle wmin
in scenario I, whereas the green curve corresponds to the current signal with the maximum duty cycle wmax . The blue curve
corresponds to the signal with the duty cycle wnT , which gives the multiperiodic solution according to the phase trajectory of
scenario I. The duty cycle wnT of the multiperiodic solution has to fulfil the following conditions:
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 15

Fig. 18. Time histories of the excitation signals with regard to extremal values of duty cycle obtained for scenario I (the red curve represents the minimum
value of the duty cycle, the green curve stands for the maximum value, and the blue curve represents a signal for some multiperiodic solution). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Period T of period-1 motion is a multiple of a period of multiperiodic motion XnT , hence

T ¼ nXnT for n 2 Nþ ff1g ð17Þ

The relations between the times of signals wmin , wmax , and wnT are as follows:

tminON  sON

tmaxOFF  sOFF ð18Þ


Now we can write equations which help us to transit from times to duty cycles:
sON ¼ wnT XnT
sOFF ¼ XnT  sON ¼ XnT ð1  wnT Þ
tminON ¼ wmin T

tmaxON ¼ wmax T

tmaxOFF ¼ T  t maxON ¼ T ð1  wmax Þ ð19Þ


Taking into account Eqs. (19) and substituting them to Eqs. (18), the following conditions for the duty cycle of the mul-
tiperiodic solution are obtained:
wnT
nwmin

wnT  1  nð1  wmax Þ ð20Þ


Numerical computations show that for the studied problem the conditions given by Eqs. (20) can be satisfied only for
n ¼ 2 and n ¼ 3.

Fig. 19. Areas of w  f configuration for period-2 and period-3 motions obtained from analysis based on scenario I (a); experimental and simulation phase
plots with Poincaré sections with regard to marked cross locators: period-2 (b) and period-3 where {f ½Hz; w ½%g: (1) ! f6:5; 70g, (2) ! f7:8; 80g,
(3)! f9:42; 78g, (4) ! f10:31; 80g.
16 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

The areas regarding period-2 and period-3 in terms of w  f configuration are shown in Fig. 19a. The two pairs of fre-
quency and duty cycle have been chosen from each of the presented areas and verified with the experiment and simulation
(see Fig. 19b, c).

6. Results discussion

Similar magnetic pendulum systems to that considered here have been discussed in the literature [11–13,26,30–33].
However, the crucial difference is in the coil excitation signal, which is pulsating in our system whereas in the systems
described in previous studies it alternates. This is why we cannot directly compare our results to the findings of other
authors. One analogy which can be drawn is the presence of chaotic and regular regimes and multiperiodic orbits, which
seems to be a characteristic feature of kick-excited systems [26]. The set of switching signal parameters reported in
Fig. 16 corresponds to period-1 one-side oscillations of our magnetic pendulum. In Fig. 16, one can see that for a 30% duty
cycle one-side oscillations are possible in the frequency range of ð2:9 Hz; 3:62 HzÞ. However, in the bifurcation diagram pre-
sented in Fig. 10d, the period-1 solution is present up to 4:8 Hz. The solution outside the range predicted by us differs from
the assumed scenario, since either the coil switches off inside the active zone or the whole trajectory may even be within the
active zone. The scenario for this solution is shown in Fig. 20, as well as the scenario for an analogical solution with the coil
switched on for the majority of the cycle. The point where the system transitions form state-1 to state-2 is denoted as
ðulk ; xlk Þ. We tried to determine the relationship between point ðuk ; xk Þ and ðulk ; xlk Þ by finding the intersection points
of phase space trajectories which have been computed for the same initial conditions ðuk ; xk Þ, whereas Eqs. (12) and (13)
were integrated numerically backwards and forwards in time, respectively.
The computations were performed for initial conditions points ðuk ; xk Þ over the area restricted by curves ðx ¼ 0Þ;
ðu ¼ uA Þ and part of the stable manifold (curve (3) in Fig. 15). As a result, the search area is divided into two sectors, which
are marked in light blue and light red in Fig. 21. Solutions exist for the blue coloured set (trajectory curves intersect) but for
the red colour there is no solution for the assumed scenarios. Kick points found for scenario I lie on the borderline between
the blue and red sectors. The physical interpretation for this is that in scenario I the energy dissipated during each period of
oscillations is perfectly balanced by the energy added to the system at the kick point. If the kick is delayed (blue sector in
Fig. 21), the system receives too much energy and an excessive amount must be subtracted from the system at the point
ðulk ; xlk Þ. An exemplary relation between the switching points is shown by the solid coloured curves 1 and 10 representing
the values of ðuk ; xk Þ and ðulk ; xlk Þ, respectively. Of course, in the red sector there is too little energy added to the system and
the cycle cannot continue. The point where the borderline meets ðx ¼ 0Þ line is actually the stable fixed point ðuS ; 0Þ for the
state-2 system.

Fig. 20. The potential courses of the pendulum for scenario II (a) and scenario III (b).

Fig. 21. Initial conditions ðuk ; xk Þ indicate the intersection of trajectories (blue area) and lack of intersection (red area). The curves marked as 1 and 10 stand
for trial values of ðuk ; xk Þ and corresponding values ðulk ; xlk Þ, whereas the stable equilibrium position is denoted by the point ðuS ; 0Þ. (For interpretation of
the references to colour in this figure legend, the reader is referred to the web version of this article.)
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 17

Fig. 22. Phase plots of sharp (a) and smooth (b) transitions between trajectories for scenario II; sharp (c) and smooth (d) transitions between trajectories for
scenario III (the solid/dashed red curve represents state-2 and the solid/dashed blue curve represents state-1). (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.)

The point where the two phase space trajectories intersect corresponds to the point ðulk ; xlk Þ for scenario II (see Fig. 22a
and b) when the angular velocity xlk is positive and the second intersection point with negative velocity corresponds to the
point ðulk ; xlk Þ from scenario III (see Fig. 22c and d). In Fig. 22 the solid/dashed red curve represents the same trajectory
obtained for state-2 in all four plots, whereas the solid/dashed blue curve represents different trajectories of state-1. Accord-
ing to our previous assumptions, outside the active zone phase the portraits for both states of the system are very similar and
almost identical. On the other hand, they differ significantly inside the active zone. This is why the trajectories intersect at a
sharp angle (see Fig. 22a and c) when the intersection point is closer to the middle of the active zone, while the intersection
angle becomes narrow if the point is closer to the active zone boundary uA . Even though scenarios II and III may be possible
from an energy balance perspective, because of the sharp transition from one state to the other it is not always possible to
obtain a solution in the continuous model of the system.

Fig. 23. Differences between the continuous and discrete model of the system in terms of a state switching signal (a) and phase trajectories for scenario I
(b), scenario II (c), and scenario III (d); label 1 stands for the continuous model and label 2 is associated with the discrete model.
18 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Fig. 24. Evolution of the basins of attraction in terms of the coil current.

The actual limitations of the analysis are difficult to estimate. To do so, one would have to analyse the time evolution of
the transition from one state to the other. Such an analysis would be crucial for investigating a system with a gradual tran-
sition between states, e.g. a system with a coil powered by a sinusoidal instead of a rectangular wave.
In general, we tried to find trajectories for the real system (described by the continuous model) by means of discrete
model analysis. Unlike the discrete model where there are only two states, in the continuous model (and in the real system)
because of the gradual change of the current signal intermediate states can be distinguished. The difference is depicted in
Fig. 23a, where the point marked at continuous curve 1 indicates our interpretation of the kick point in the continuous
model.
In Fig. 23b–d, the influence of the gradual current change on the system trajectories for different scenarios is emphasized.
It can be observed that trajectories corresponding to the continuous model (curve 1) are smoother around the kick point. In
Fig. 23c, one can clearly see the state change even before the kick indication for the continuous model solution. The gradual
transition from state 1 to state 2 is also depicted in Fig. 24, which presents the evolution of the basins of attraction in terms of
coil current level.
As can be noticed from Fig. 24, the two-well potential is well developed for current amplitude around 0:5 A. Although the
current is only 50% of the nominal value, the active zone range is around 95% of the nominal value.

7. Summary and conclusions

This paper has described a magnetic pendulum system in which magnetic interaction between the magnet placed at the
end of the pendulum and an electrical coil placed underneath serves as the source of the excitation. The coil is powered by a
pulsating rectangular wave current signal. Since the magnetic interaction weakens with the distance between the magnet
and the coil, excitation in the system depends on time as well as on the position of the pendulum. A physical model of
the system was assumed and its parameters were identified by means of experiment. Two approaches to mathematical mod-
elling were taken: the first with a continuous approximation of the rectangular current signal, and the second with the sys-
tem discretised into two states, i.e. with the coil switched on or off.
A specific type of solution was analysed, when the pendulum oscillates to one-side of the coil, i.e. with the same sign of
angular displacement. The analysis was conducted with a discretised mathematical model of the system and based on the
assumption that besides the so-called ‘‘active zone” the magnetic force in the system can be neglected. We identified kick
points numerically in terms of angular position and velocity, where the coil should be switched on in order to sustain
one-side oscillations. Next, the kick points were used to determine the excitation signal parameters in terms of a frequency
and a duty cycle. These parameters were verified experimentally on our laboratory stand, as well as numerically with a con-
tinuous model of the system. The results were also extended to multiperiodic one-side oscillations, in the case of our system
up to period-3 solutions, which were also verified experimentally. We reported two further scenarios for one-side oscilla-
tions, which cannot be determined in a continuous system on the basis of discrete system analysis.
The original contributions of this work can be summarized as follows:

1. Unlike other studies [38–41] in which the potential energy is modelled with a polynomial function, we modelled the
potential energy magnetic term with a Gaussian function. This makes it possible to differentiate the mechanical and mag-
netic potential energy terms. The obtained magnetic torque model, despite its mathematical simplicity, provides good
agreement with the experimental data.
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 19

2. Analysis of the origin of one-side oscillations reveals that, because of localized forcing, the same oscillations can be
obtained for different duty cycles and the same frequency of the excitation signal.
3. Our findings on multiperiodic solutions show that this feature in kick-excited systems may arise from localized forcing.
This aspect of the research suggests that in such systems the periodicity of the regular solution can be multiplied without
disturbing the trajectory.

In future studies, we will investigate other types of periodic trajectories in a magnetic pendulum system. We also plan to
improve our approach by including a gradual change of the current signal in our analysis. Further research will seek to con-
firm and develop our initial findings by means of analytical methods, probably requiring the construction of a simplified
mathematical model of a kick-excited system.

CRediT authorship contribution statement

Adam Wijata: Conceptualization, Methodology, Software, Formal analysis, Writing - original draft, Visualization. Krys-
tian Polczyński: Conceptualization, Methodology, Software, Formal analysis, Writing - original draft, Visualization, Investi-
gation. Jan Awrejcewicz: Supervision, Writing - review & editing, Funding acquisition.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Acknowledgement

This work has been supported by the Polish National Science Centre under the grant OPUS 14 No. 2017/27/B/ST8/01330.
This article is an extended version of a paper originally presented at ICEDyn 2019 congress.

Appendix A

The measurement method of a steady magnetic moment M m ½Nm is shown in Fig. A.1a. The polymer disk is joined with
the tensometric beam by the string. When the pendulum is angled due to repulsive magnetic field generated by the coil, we
start to move the tensometric beam slowly downward. The string force F ½N and angle u are recorded. For static equilibrium
position:

Mm ¼ mgs sin u þ FR ðA:1Þ


The measurements were conducted for different values of the coil current and the results are reported in Fig. A.1b.

Fig. A1. (a) Measurement method of the magnetic moment M m and (b) measured magnetic torque values.
20 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

Fig. B1. (a) Experimental and numerical phase plots of variables u€ and u during free vibrations of the pendulum (without the elastic coupling) used for
estimation of the mgs=J ratio; b) experimentally estimated friction characteristic (blue points) with numerically identified friction model (red curve).

Appendix B

In general, the system parameters were identified numerically, but they were first estimated from experimental measure-
ments. Analysis of the free vibrations of the pendulum provides approximate values for the mgs
J
ratio. The motion of free oscil-
lations (where the damping torques are neglected) is expressed as
mgs
u
€ ¼ sin u ðB:1Þ
J
The relation between the angular acceleration u € and the angular displacement u stands for a sinusoidal function, where
the mgs
J
ratio plays the role of the amplitude of the sine wave. By making a phase plot of variables u
€ and u based on exper-
imental data and fitting the model governed by Eq. (B.1) to the data (see Fig. B.1a), we obtained the first estimate of mgs
J
, which
was equal to 85:5 rad=s2 .
To visualise friction characteristic for ball bearings in our set-up, the estimate of a friction characteristic was derived
numerically from the experimentally recorder angular displacement of a free motion (without elastic coupling). Using esti-
mated mgs
J
value and the equation

mgs 1
u
€ ¼ sin u  M ðu
_Þ ðB:2Þ
J J F
we prepared a parametric plot of M F ðu_ Þ estimate (blue points in Fig. B.1b). As it can be seen in comparison with Fig. 3, all
typical features of the Stribeck curve can be recognised from points distribution i.e. static friction, Stribeck effect and viscous
friction.

Appendix C. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.ymssp.2020.107229.

References

[1] R. Ke˛piński, J. Awrejcewicz, D. Lewandowski, J. Gajek, Experimental investigations of stability in a hybrid stepper motor, Adv. Intell. Syst. Comput. 317
(2015) 81–90, https://doi.org/10.1007/978-3-319-10990-9_8.
[2] J. Gajek, J. Awrejcewicz, Mathematical models and nonlinear dynamics of a linear electromagnetic motor, Nonlinear Dyn. 94 (2018) 377–396, https://
doi.org/10.1007/s11071-018-4365-0.
[3] E. Bonisoli, C. Delprete, Nonlinear and linearised behaviour of the Levitron Ò, Meccanica 51 (2016) 763–784, https://doi.org/10.1007/s11012-015-
0238-5.
[4] S.S. Nielsen, R.K. Holm, P.O. Rasmussen, Conveyor system with a highly integrated permanent magnet gear and motor, IEEE Trans. Ind. Appl. 56 (2020)
2550–2559, https://doi.org/10.1109/TIA.2020.2977877.
[5] D.D. Kozanecka, Theoretical and experimental investigations of dynamics of the flexible rotor with an additional active magnetic bearing, in: R. Sehgal
(Ed.), Perform. Eval. Bear., InTech, Rijeka, 2012, pp. 163–192. https://doi.org/10.5772/51113.
[6] H.Y. Zhang, Z.Q. Chen, X.G. Hua, Z.W. Huang, H.W. Niu, Design and dynamic characterization of a large-scale eddy current damper with enhanced
performance for vibration control, Mech. Syst. Signal Process. 145 (2020), https://doi.org/10.1016/j.ymssp.2020.106879 106879.
[7] K. Tsubono, A. Araya, K. Kawabe, S. Moriwaki, N. Mio, Triple-pendulum vibration isolation system for a laser interferometer, Rev. Sci. Instrum. 64
(1993) 2237–2240, https://doi.org/10.1063/1.1143967.
[8] F. Sun, K. Oka, Feasibility analysis of two iron balls’ simultaneous suspension using flux path control mechanism, J. Syst. Des. Dyn. 5 (2011) 1155–1166,
https://doi.org/10.1299/jsdd.5.1155.
[9] N. Ida, Design and control of a magnetic pendulum actuator, Proc. Int. Conf. Optim. Electr. Electron. Equipment, OPTIM, 2012, pp. 439–443. https://doi.
org/10.1109/OPTIM.2012.6231898
A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229 21

[10] A.E. Motter, M. Gruiz, G. Károlyi, T. Tél, Doubly transient chaos: generic form of chaos in autonomous dissipative systems, Phys. Rev. Lett. 111 (2013),
https://doi.org/10.1103/PhysRevLett.111.194101 194101.
[11] G. Khomeriki, Parametric resonance induced chaos in magnetic damped driven pendulum, Phys. Lett. Sect. A Gen. At. Solid State Phys. 380 (2016)
2382–2385, https://doi.org/10.1016/j.physleta.2016.05.049.
[12] D.Y. Skubov, D.S. Vavilov, Dynamics of the conductivity bodies of pendulum types in alternating magnetic field, ZAMM – J. Appl. Math. Mech./Z. Angew.
Math. Mech. 94 (2014) 951–956, https://doi.org/10.1002/zamm.201300257.
[13] Y. Luo, W. Fan, C. Feng, S. Wang, Y. Wang, Subharmonic frequency response in a magnetic pendulum, Cit. Am. J. Phys. 88 (2020) 115, https://doi.org/
10.1119/10.0000038.
[14] J.A. Blackburn, S. Vik, B. Wu, H.J.T.T. Smith, Driven pendulum for studying chaos, Rev. Sci. Instrum. 60 (1989) 422–426, https://doi.org/10.1063/
1.1140394.
[15] J.A. Blackburn, G.L. Baker, A comparison of commercial chaotic pendulums, Am. J. Phys. 66 (1998) 821–830, https://doi.org/10.1119/1.18966.
[16] J. Bethenod, Sur l’entretien du mouvement d’un pendule au moyen d’un courant alternatif de fréquence élevée par rapport à sa fréquence proper,
Comptes Rendus Hebd Des Séances l’Académie Des Sci. 207 (1938) 847–849.
[17] P.S. Landa, in: Nonlinear Oscillations and Waves in Dynamical Systems, Springer, Netherlands, Dordrecht, 1996, https://doi.org/10.1007/978-94-015-
8763-1_18.
[18] H.P. Knauss, P.R. Zilsel, Magnetically maintained pendulum, Am. J. Phys. 19 (1951) 318–320, https://doi.org/10.1119/1.1932808.
[19] Y. Rocard, Dynamique générale des vibrations, Masson, Paris, 1949.
[20] N. Minorsky, Stationary solutions of certain nonlinear differential equations, J. Franklin Inst. 254 (1952) 21–42, https://doi.org/10.1016/0016-0032(52)
90003-3.
[21] N. Kesavamurthy, G.P. Rao, A study of Bethenod’s phenomenon, IEEE Trans. Circ. Theory 19 (1972) 215–218, https://doi.org/10.1109/
TCT.1972.1083437.
[22] B.Z. Kaplan, Topological considerations of parametric electromechanical devices and their parametric analysis, IEEE Trans. Magn. 12 (1976) 373–380,
https://doi.org/10.1109/TMAG.1976.1059035.
[23] B.H. Smith, Some characteristics of a class of parametric reluctance machines, Proc. Inst. Electr. Eng. 126 (1979) 162–166, https://doi.org/10.1049/
piee.1979.0036.
[24] A.P. Russell, I.E.D. Pickup, Principles of operation of the parametric reluctance motor, Electr. Mach. Power Syst. 5 (1980) 485–496, https://doi.org/
10.1080/07313568008955425.
[25] J.J. Blakley, On an analysis of a novel tuned circuit electromechanical oscillator with a long traverse of motion, Proc. IEEE 70 (1982) 310–311, https://
doi.org/10.1109/PROC.1982.12298.
[26] V. Damgov, I. Popov, ‘‘Discrete” oscillations and multiple attractors in kick-excited systems, Discret. Dyn. Nat. Soc. 4 (2000) 99–124, https://doi.org/
10.1155/S102602260000011X.
[27] D. Cintra, P. Argoul, Non-linear argumental oscillators: Stability criterion and approximate implicit analytic solution, Int. J. Non Linear. Mech. 94 (2017)
109–124, https://doi.org/10.1016/j.ijnonlinmec.2017.03.013.
[28] D. Cintra, P. Argoul, Nonlinear argumental oscillators: a few examples of modulation via spatial position, JVC/J. Vib. Control 23 (2017) 2888–2911,
https://doi.org/10.1177/1077546315623888.
[29] D.I.I. Penner, D.B. Doubochinski, M.I.I. Kozakov, A.S.S. Vermel’, Y.V.V. Galkin, Asynchronous excitation of undamped oscillations, Uspekhi Fiz. Nauk. 16
(1973) 402, https://doi.org/10.1070/PU1973v016n01ABEH005156.
[30] D.B. Doubochinski, J. Tennenbaum, R. De Wattignies, Theory and applications of the macroscopic quantization effect in nonlinearly-coupled vibrating
systems, in: 1st Euro-Mediterranean Conf. Struct. Dyn. Vibroacoustics, Marrakech, Marocco, 2013, pp. 23–26.
[31] J. Tennenbaum, Amplitude quantization as an elementary property of macroscopic vibrating systems, 21st Century Sci. Technol., 2005, pp. 50–63
[32] Y. Kraftmakher, Experiments with a magnetically controlled pendulum, Eur. J. Phys. 28 (2007) 1007–1020, https://doi.org/10.1088/0143-0807/28/5/
023.
[33] Y. Kraftmakher, Demonstrations with a magnetically controlled pendulum, Am. J. Phys. 78 (2010) 532–535, https://doi.org/10.1119/1.3276412.
[34] V. Tran, E. Brost, M. Johnston, J. Jalkio, Predicting the behavior of a chaotic pendulum with a variable interaction potential, Chaos 23 (2013), https://doi.
org/10.1063/1.4812721 033103.
[35] B. Nana, S.B.B. Yamgoué, R. Tchitnga, P. Woafo, Dynamics of a pendulum driven by a DC motor and magnetically controlled, Chaos Solitons Fract. 104
(2017) 18–27, https://doi.org/10.1016/j.chaos.2017.07.027.
[36] B. Nana, S.B. Yamgoué, R. Tchitnga, P. Woafo, Nonlinear dynamics of a sinusoidally driven lever in repulsive magnetic fields, Nonlinear Dyn. 91 (2018)
55–66, https://doi.org/10.1007/s11071-017-3839-9.
[37] A. Siahmakoun, V.A. French, J. Patterson, Nonlinear dynamics of a sinusoidally driven pendulum in a repulsive magnetic field, Am. J. Phys. 65 (1997)
393–400, https://doi.org/10.1119/1.18546.
[38] B.P. Mann, Energy criterion for potential well escapes in a bistable magnetic pendulum, J. Sound Vib. 323 (2009) 864–876, https://doi.org/10.1016/j.
jsv.2009.01.012.
[39] C.A. Kitio Kwuimy, C. Nataraj, M. Belhaq, Chaos in a magnetic pendulum subjected to tilted excitation and parametric damping, Math. Probl. Eng. 2012
(2012) 1–18, https://doi.org/10.1155/2012/546364.
[40] F.C. Moon, P.J. Holmes, A magnetoelastic strange attractor, J. Sound Vib. 65 (1979) 275–296, https://doi.org/10.1016/0022-460X(79)90520-0.
[41] M.Ó. Donnagáin, O. Rasskazov, Numerical modelling of an iron pendulum in a magnetic field, Phys. B Condens. Matter 372 (2006) 37–39, https://doi.
org/10.1016/j.physb.2005.10.098.
[42] A.N. Kadjie, P. Woafo, Effects of springs on a pendulum electromechanical energy harvester, Theor. Appl. Mech. Lett. 4 (2014), https://doi.org/10.1063/
2.1406301 063001.
[43] P. V. Malaji, M. Rajarathinam, V. Jaiswal, S.F. Ali, I.M. Howard, Energy harvesting from dynamic vibration pendulum absorber, in: A. Rama Mohan Rao,
K. Ramanjaneyulu (Eds.), Lect. Notes Civ. Eng., Springer, Singapore, 2019, pp. 467–478. https://doi.org/10.1007/978-981-13-0365-4_40.
[44] K. Polczyński, A. Wijata, J. Awrejcewicz, G. Wasilewski, Numerical and experimental study of dynamics of two pendulums under a magnetic field, Proc.
Inst. Mech. Eng. Part I J. Syst. Control Eng. 233 (2019) 441–453, https://doi.org/10.1177/0959651819828878.
[45] K. Polczyński, A. Wijata, G. Wasilewski, G. Kudra, J. Awrejcewicz, Modelling and analysis of bifurcation dynamics of two coupled pendulums with a
magnetic forcing, in: I. Kovacic, S. Lenci (Eds.), IUTAM Symp. Exploit. Nonlinear Dyn. Eng. Syst., Springer International Publishing, Cham, 2020, pp. 213–
223. https://doi.org/10.1007/978-3-030-23692-2_19.
[46] A. Fradkov, B. Andrievsky, K. Boykov, Control of the coupled double pendulums system, Mechatronics 15 (2005) 1289–1303, https://doi.org/10.1016/j.
mechatronics.2005.03.008.
[47] M. Wojna, A. Wijata, G. Wasilewski, J. Awrejcewicz, Numerical and experimental study of a double physical pendulum with magnetic interaction, J.
Sound Vib. 430 (2018) 214–230, https://doi.org/10.1016/j.jsv.2018.05.032.
[48] J.P. Berdahl, K. Vander Lugt, Magnetically driven chaotic pendulum, Am. J. Phys. 69 (2001) 1016–1019, https://doi.org/10.1119/1.1387041.
[49] R. Kumar, S. Gupta, S.F. Ali, Energy harvesting from chaos in base excited double pendulum, Mech. Syst. Signal Process. 124 (2019) 49–64, https://doi.
org/10.1016/j.ymssp.2019.01.037.
[50] N. Gure, A. Kar, E. Tacgin, A. Sisman, N.M. Tabatabaei, Hybrid energy harvesters (HEHs)—a review, in: N. Bizon, N.M. Tabatabaei, F. Blaabjerg, E. Kurt
(Eds.), Energy Harvest. Energy Effic. Technol. Methods, Appl., Springer, 2017, pp. 17–61. https://doi.org/10.1007/978-3-319-49875-1_2.
[51] P.V. Malaji, S.F. Ali, Analysis and experiment of magneto-mechanically coupled harvesters, Mech. Syst. Signal Process. 108 (2018), https://doi.org/
10.1016/j.ymssp.2018.02.025.
22 A. Wijata et al. / Mechanical Systems and Signal Processing 150 (2021) 107229

[52] P.V. Malaji, S.F. Ali, Magneto-mechanically coupled electromagnetic harvesters for broadband energy harvesting, Appl. Phys. Lett. 111 (2017) 1–6,
https://doi.org/10.1063/1.4997297.
[53] P.V. Malaji, S.F. Ali, Analysis of energy harvesting from multiple pendulums with and without mechanical coupling, Eur. Phys. J. Spec. Top. 224 (2015)
2823–2838, https://doi.org/10.1140/epjst/e2015-02591-7.
[54] P.V. Malaji, S.F. Ali, Broadband energy harvesting with mechanically coupled harvesters, Sensors Actuators A Phys. 255 (2017) 1–9, https://doi.org/
10.1016/J.SNA.2016.12.003.
[55] L. Perko, in: Differential Equations and Dynamical Systems, Springer, New York, New York, NY, 2001, https://doi.org/10.1007/978-1-4613-0003-8.
[56] J. Awrejcewicz, in: Ordinary Differential Equations and Mechanical Systems, Springer International Publishing, 2014, https://doi.org/10.1007/978-3-
319-07659-1.
[57] D.E. Sander, H. Allmaier, H.H. Priebsch, M. Witt, A. Skiadas, Simulation of journal bearing friction in severe mixed lubrication – validation and effect of
surface smoothing due to running-in, Tribol. Int. 96 (2016) 173–183, https://doi.org/10.1016/j.triboint.2015.12.024.
[58] J.L. do Vale, C.H. da Silva, Kinetic friction coefficient modeling and uncertainty measurement evaluation for a journal bearing test apparatus, Meas. J.
Int. Meas. Confed. 154 (2020) 107470, https://doi.org/10.1016/j.measurement.2020.107470.
[59] F. Marques, P. Flores, J.C. Pimenta Claro, H.M. Lankarani, A survey and comparison of several friction force models for dynamic analysis of multibody
mechanical systems, Nonlinear Dyn. 86 (2016) 1407–1443, https://doi.org/10.1007/s11071-016-2999-3.
[60] C. Forbes, M. Evans, N. Hastings, B. Peacock, Statistical Distributions, fourth ed., John Wiley & Sons, New Jersey, 2010. https://doi.org/10.1002/
9780470627242.
[61] J.A. Greenwood, H.O. Hartley, Guide to Tables in Mathematical Statistics, Princeton University Press, Princeton, New Jearsy, 1962.
[62] T. Ida, M. Ando, H. Toraya, Extended pseudo-Voigt function for approximating the Voigt profile, J. Appl. Crystallogr. 33 (2000) 1311–1316, https://doi.
org/10.1107/S0021889800010219.
[63] Y. Liu, J. Lin, G. Huang, Y. Guo, C. Duan, Simple empirical analytical approximation to the Voigt profile, J. Opt. Soc. Am. B 18 (2001) 666, https://doi.org/
10.1364/josab.18.000666.
[64] S. Strogatz, M. Friedman, A.J. Mallinckrodt, S. McKay, Nonlinear dynamics and chaos: with applications to physics, biology, chemistry, and engineering,
Comput. Phys. 8 (1994) 532, https://doi.org/10.1063/1.4823332.
[65] WolframAlpha, NonlinearModelFit—Wolfram Language Documentation, https://Reference.Wolfram.Com/Language/Ref/NonlinearModelFit.Html.
(2020) (accessed 01 June 2020).

You might also like