Anda di halaman 1dari 46

UNIVERSITY OF BRISTOL

DEPARTMENT OF
ENGINEERING MATHEMATICS

HOW DO CHILDREN SWING?

Dayal C. Strub (Engineering Mathematics)

Project thesis submitted in support of the degree of


Master of Engineering

Supervisor: Prof. A.R. Champneys, Engineering Mathematics


June 2009
How Do Children Swing?
A Study of the Dynamics of Playground Swings

Project thesis submitted in support of the degree of


Master of Engineering

Dayal C. Strub
Supervisor: Prof. A.R. Champneys

June 2009

i
Abstract

The dynamics of the playground swing is studied paying particular attention to the initiation of motion
from rest. The case considered is that of a child pumping the swing from a standing position, which has
been the cause of controversy over which of the two canonical models, namely the parametric oscillator
and the compound pendulum models found in the literature, best represented the actual motion of a
swing. This issues is tackled by considering the two models and by analysing them using the same
techniques such that they may be compared properly.
The models are first considered in their usual non-autonomous versions. The parametric oscillator model
is shown to exhibit coexistence and the instability regions in the double pendulum, due to the parametric
terms, are shown to display instability pockets when the forcing is sinusoidal.
These models are also compared with simple experiments which support the compound pendulum model
and the fact that most of the forcing is direct as opposed to parametric, as is also found when analysing
the linearised, truncated double pendulum model in which the direct and parametric terms are found to
be of first and second order in the amplitude of forcing.
An autonomous description of the canonical models is also considered, as both the experiments and simple
observation of a child on a swing show that the forcing is actually a function of the swings angle. A
simple analysis of these versions of the models is conducted, giving amplitude maps for the two models.
These maps show a geometric and an arithmetic increase of the amplitude for the parametric and double
pendulum model respectively, as expected.
Using again the results from the experiment a simple triple pendulum model is also introduced to account
for the non-rigid supports found in almost all swings. This model is then briefly compared with the
experimental results via numerical simulations.

ii
Acknowledgements

By definition, this thesis comes as the conclusion of my four years of undergraduate studies. It therefore
(hopefully) combines all the knowledge and skills that I have learned throughout these years, which I owe
to the people that I have been lucky enough to meet. Without their help, insight and patience, this work
would not have been possible.
Most importantly, I want to express my gratitude to Prof. Alan R. Champneys for his help throughout
the project, which ranged from insight into the models and analysis to technical help with the writing
and preparation of this thesis, and was very much appreciated.
I would also like to thank everyone that has had to put up with me during the preparation of this thesis,
most importantly family and friends.

iii
Contents

List of Figures v

1 Introduction 6
1.1 Review of the Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Simple Experiments of Swinging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Organisation of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 General Theory of Parametric Resonance 11


2.1 Hill Equations and Floquet Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 The Phenomena of Coexistence and Instability Pockets . . . . . . . . . . . . . . . . . . . . 14
2.3 The Effects of Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Contrast with Directly Forced Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3 Development of the Two Canonical Models 17


3.1 Parametric Oscillator Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Double Pendulum Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

4 Analysis 21
4.1 Parametric Oscillator Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.1.1 Instability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.1.2 Analysis using the Method of Averaging . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2 Double Pendulum Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.1 Instability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2.2 Analysis of Full System using Averaging . . . . . . . . . . . . . . . . . . . . . . . . 26

5 Autonomous Description and Analysis 28


5.1 Parametric Oscillator Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2 Double Pendulum Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

6 Further Modelling and Simulations 33


6.1 Derivation and Simulations of the Triple Pendulum Model . . . . . . . . . . . . . . . . . . 33

7 Conclusions and Outlook 36

A Analytical Methods for Parametric Resonance 38


A.1 Methods to Find the Resonance Tongues . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
A.2 Method to Check for the Presence of Coexistence . . . . . . . . . . . . . . . . . . . . . . . 39

B The Method of Averaging 42

References 44

iv
List of Figures

1.1 Flashlight trace produced by child, starting at point A. Figure taken from [28]. . . . . . . 7
1.2 Left: Schematic side view showing position of lights, angles and moving coordinate system.
Right: Flashlight traces and time series of (dotted), (dashed) and (solid). . . . . . . 8
1.3 Left: Time series of (dotted), (dashed) and lCM (solid). Right: Time series of
(dotted), CMtan (dashed) and CMrad (solid). Note: CM coordinates and lCM have been
scaled and adjusted for ease of comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.1 Possible combinations of Floquet multipliers for a conservative Hill equation. Left: Unsta-
ble, with tr (M ) > 2. Right: Stable, with tr (M ) < 2. Figure taken from [2]. . . . . . . . . 12
2.2 Left: Ince-Strutt diagram for the canonical Matheiu equation. Unstable regions coloured
according to magnitude of the greater Floquet multiplier, with darker colours denoting
greater values. Right: Stability diagram for the damped Matheiu equation (2.3.1), con-
sisting of multiple Ince-Strutt diagrams for different values of . . . . . . . . . . . . . . . . 14
2.3 Ince-Strutt diagrams showing coexistence, left, and instability pockets, right. . . . . . . . 15
2.4 Left: Amplitude of the directly forced pendulum as a function of frequency for various
amounts of damping. Right: Angular displacement as a function of time for direct (solid)
and parametric (dashed) forcing, with 2 = 1, F = 1. . . . . . . . . . . . . . . . . . . . . . 16

3.1 Representation of a playground swing as a simple pendulum of variable length and as a


double pendulum, to the left and right respectively. . . . . . . . . . . . . . . . . . . . . . . 17

4.1 Left: Ince-Strutt diagram for the truncated parametric oscillator (4.1.2), using transition
curves. Right: Ince-Strutt diagram for the full parametric oscillator (4.1.1). . . . . . . . . 22
4.2 Left: Sketch of slow flow phase portraits for the parametric model, Eq. (4.1.2), shown over
the Ince-Strutt diagram. U and S denoting stable and unstable regions. Right: numerically
computed phase portraits for Eq. (4.1.2), showing a stable and an unstable case. . . . . . 23
4.3 Ince-Strutt diagrams for the homogeneous part of the double pendulum model (4.2.1), and
the truncated version (4.2.3), with = 0.3 and = 0.1, to the left and right respectively. 25

5.1 Amplitude maps for autonomous descriptions. Left: Parametric model, with = 0.2.
Right: Double pendulum model with direct forcing (dashed) and combined direct-parametric
forcing (dash-dot), where = 0.65, = 0.13 and = 0.26. . . . . . . . . . . . . . . . . . . 29

6.1 Representation of a playground swing as a triple pendulum. . . . . . . . . . . . . . . . . . 34


6.2 Left: Numerically produced flashlight traces. Right: Time series of (dotted), (dashed)
and (solid) from numerical simulation and experiment, above and below respectively.
Values used for simulations: 1 = 0.43, 2 = 0.21, 1 = 0.53, 2 = 0.31, A = /9 . . . . . 35

v
Chapter 1

Introduction

The aim of this study is to better understand the dynamics of the playground swing, in particular to
investigate the physics of how a child inputs energy into each oscillation without any external contact.
The hope being that through a study and comparison of the various existing models, their strengths
and weaknesses shall be found, therefore shedding some light on the controversy over which model best
represents the actual motion. Some interesting but non-exhaustive comparisons, mainly between the
parametric and the driven (double pendulum) model, already exist [7, 35, 36]. Finally, using the under-
standing obtained from study of the existing models and from the simple experiments performed, some
improvements shall be considered.
The first step is clearly that of introducing the models and generally the literature, which shall be done
in this chapter. This shall be followed by a discussion of experimental evidence and a summary of the
layout of the thesis.

1.1 Review of the Literature


The universality of the playground swing, together with its interesting but nonetheless elementary dy-
namics, led Tea & Falk [28] to propose the swing as a pedagogical aid for the teaching of mechanics. This
property was made even more evident when some later papers, [6, 7, 27], showed that demonstrations are
easily set up, and other papers, [26,33], showed that there exist abundant similar examples of parametric
oscillators. One such example is O Botafumeiro, the Santiago de Compostela thurible [26, 33].
The article by Tea & Falk [28] considered the initiation of motion, and the so-called pumping of swings.
The term pumping is here used to describe the input of energy into the system, via the repeated
change in the swingers position and/or orientation relative to the swing. Many articles then fol-
lowed [68, 10, 12, 13, 20, 23, 27, 3436], as the authors felt that the dynamics of the swing had either
not been done justice, or had been completely misunderstood. This led to a variety of models and anal-
yses. These however only really present one major difference; between the models with parametric and
direct forcing, (although there are also differences between models with autonomous and non-autonomous
forcing).
Originally the case of a standing child pumping a swing was considered, probably as it was thought to
have simpler dynamics than the seated case, and was modelled as a parametric oscillator [27,28], which is
still widely believe to be the model which best represents the swings dynamics. A parametric oscillator
is a system in which one of the parameters of the system varies in time causing the system to oscillate [3],
as shall be seen in more detail later. In the the simplest form of such a model for the swing, the child is
considered to be a point mass, located at his centre of mass, which varies the length of the pendulum by
standing and crouching, in order to pump the swing. If one is able to recall how to pump a swing, one

6
Fig. 1.1: Flashlight trace produced by child, starting at point A. Figure taken from [28].

immediately realises that this model has little in common with actual the motion carried out by children
on a swing. In particular in sitting as well as in standing positions there is clearly a great deal of leaning
backwards and forwards. This discrepancy must be due, either to little attention being paid to the actual
motion carried out by children, or to a desire for simplicity, but most probably to a combination of the
two.
In the article by Tea & Falk [28], there is a photograph which is useful in understanding how the model
came into existence. This photograph shows a rough path of the motion of the centre of mass of a child,
and is reproduced in Fig. 1.1. From this figure, one can see that the parametric oscillator model is an
oversimplification, but at the same time that there would appear a parametric component to the motion.
More discussion of experimental evidence will follow in 1.2.
The parametric oscillator model was first challenged by Gore [12], as it is unable to account for the
initiation of motion, as the force of gravity is seen to be central when the swing is at rest, therefore
giving no force component with which to create a torque and initiate motion. Some authors dismiss this
problem, by appealing to the fact that perfect rest cannot be achieved in practice, such that parametric
amplification is always possible [10, 34]. Taking this problem into account however, and through a more
careful observation of the motion performed by children on swings [7], led to the compound pendulum
model. In this model, the child oscillates back and forth relative to the swings supports, rather than
moving up and down them. Thus, this is no longer purely parametric forcing, and it shall later be shown
to be better described as direct forcing. Many variations of this model have been considered, namely
nonrigid swing supports leading to a triple constrained pendulum [12, 20], rigid supports giving a double
constrained pendulum [20], and a rigid body child [7], as opposed to a point mass.
The case of the seated child has also been considered [8, 35, 36], and the prominent model is that of a
pendulum, where the usual point mass is replaced by a rigid dumb-bell with three masses and connected
at the central mass.
This report shall focus on the case of a standing child pumping a swing, which is the most popular one
in the literature, and has caused the most debate regarding its dynamics. It is believed however, that in
practice and in the compound pendulum models, the seated and the standing cases have a lot in common,
as both involve a great deal of leaning backwards and forewards.
The central aim of this thesis is to examine and compare previous models in some detail; most crucially
to put them all on the same mathematical footing. In so doing, as hopefully a hierarchy of models of
increasing fidelity is studied, an answer will be sought to the basic question: How do children swing?

7
0
1
0
1
0
1
0
1
0 1, pivot
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0 l1
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0 11
1 00
0
1 00 2, hand
11
0 11
1 00
0
1
0
1
0 00
11
1
0 00
11
1
0 00
11
1
0
1
0
l1 4, shoulder
0
1
2
0
1
0
1
0
1

,,
0
1 l3
00 1
11 0
00 3,1
11 0
00 1
11 0seat
0
1 0

0
rad 1
0
1
0
1
tan

time
Fig. 1.2: Left: Schematic side view showing position of lights, angles and moving coordinate system.
Right: Flashlight traces and time series of (dotted), (dashed) and (solid).

1.2 Simple Experiments of Swinging

Most of the studies to date, with the major exception of the study by Post et al. [23], have not been
empirical studies. This experimental study closely followed the modelling work of Case & Swanson [8]
and focused on the seated case.
The results from the study confirm the results of Case & Swanson [8] in a number of ways. Most
importantly the study shows that direct forcing plays a major role and that parametric forcing only
plays a subordinate role. Also, as the amplitude of the swing increases, the contributions of direct and
parametric forcing were found to decrease and increase respectively, whilst still maintaining the same
roles, which is also found in the models [7,8]. The experimental results also validate the use of sinusoidal
forcing, as the average harmonicity was found to be 0.97, where a value of one denotes perfectly sinusoidal
motion and the definition can be found in the article by Post et al. [23]. Even though, the authors do
mention that they specifically asked the participants to perform their movements as smoothly and as
regularly as possible [23].
The experiment also produced some unexpected results that are not predicted by the models. Firstly,
the tangential and radial distances of the persons centre of mass from the swings supports were found
to reach their maximum values while the swing moved from the maximum negative amplitude to its
lowest point. Whereas, the generally accepted direct forcing occurs at the maximum negative amplitude,
therefore implying that these distances reach their maximum values at this point. The other interesting
result from the study was that the coordination pattern of the person on the swing was largely determined
by the swings motion, even though this result is probably most relevant for the seated case, in which
coordination between the persons various limbs occurs.
This study is not intended to be an experimental one, however brief phenomenological experiments were
conducted in order to obtain empirical data to validate the standing case models. In essence the setup was
a combination of the one used by Post et al. [23], even though much simpler, and the method employed
by Tea & Folk [28] to take the trace-photo, see Fig. 1.1. Four lights were placed on the swing (at the
pivot and the seat) and the person (on the hands and at shoulder height), see Fig. 1.2, which immediately
brings a triple pendulum to mind. This possible model will be considered in Chapter 6.

8
,CMt ,CMr
,,lCM

0 0

time time

Fig. 1.3: Left: Time series of (dotted), (dashed) and lCM (solid). Right: Time series of (dotted),
CMtan (dashed) and CMrad (solid). Note: CM coordinates and lCM have been scaled and adjusted for
ease of comparison.

It should be noted that the position of the second light on the person should be the centre of mass.
However, the actual centre of mass was not found in this study, as the focus was on finding the phase
difference between the various components, and generally to obtain a qualitative understanding. This left
the choice of placing the light anywhere along the sagittal plane, which is the imaginary plane from head
to toe through the body [23]. The final choice was therefore that of maximum possible distance between
the lights on the swings seat and the person, with the hope that this might reduce the percentage error
when finding the angles. The shoulder was therefore the best choice as the head was discarded for fear
of other, unrelated movements.
The experiment then consisted of taking a video of a person pumping a swing in the dark, such that only
the lights were visible, making them more easy to distinguish. This was considered to be the simplest
possible method of data collection, as there is plenty of software available for video analysis, which one
can use to find the time series of the lights. The software used in this study was Matlab together with
some of its image processing toolboxes.
Quite simply what was done was to input the video into Matlab and find the position of the lights in
each frame and therefore over time. However, three major difficulties were encountered in finding the
points. Two of the difficulties in finding the position of the lights were due to the presence of noise
from background lights, and the disappearance of the lights behind the swings frame. The last, non-
trivial difficulty was in distinguishing between the points, which was solved by using the knowledge of
the geometry of the swing and the positioning of the lights. Specifically, the light at the pivot was fixed
and the highest in all frames, and the one on the seat was always the lowest, leaving the two lights on the
person somewhere in the middle, with the one on the hand always closest to the pivot, which is clearly
seen in the trace in Fig. 1.2.
Following Post et al. [23], a local (moving) Cartesian coordinate system with one axis fixed to the swings
support (radial direction) and the other orthogonal to it (tangential direction), was used to follow the
movement of the centre of mass of the person. The two movements that one would expect a person to
make in order to pump a swing, namely leaning and crouching, are usually never assumed to both occur
in a model. However, the results obtained from the experiment show that both methods are used, as can
be see in Fig. 1.3. Now, changing to the moving coordinate system gives a better idea of the distinction
between direct and parametric forcing, the results can be seen in Fig. 1.3.

1.3 Organisation of the Thesis


The following provides a brief summary of how the chapters in this thesis are organised.
The first step will be to introduce some of the relevant theory for the dynamical systems being considered,

9
specifically the non-autonomous ones. This shall be done in Chapter 2, where Floquet theory, the general
theory of periodic first-order systems of linear ordinary differential equations, and Hill-type equations
will be introduced. The main focus, in introducing these equations, shall be on their stability properties
and certain phenomena displayed by the parameter regions. This introduction shall be kept brief as the
theory can be found in most text books, such as the references in Chapter 2. This Chapter shall be
referred to often as this theory will be used throughout the report.
Chapter 3 is the one in which variants of existing models will be derived from first principles from the
mechanics of a swing. The models chosen are the parametric oscillator model and the double pendulum
model, as these are the simplest versions of the two canonical models. Simplicity being sought after, as
it is always important in mathematical modelling, but also with the notion of the playground swing as a
pedagogical aid in mind.
In Chapter 4 these models will be analysed. This will be done in a slightly more rigorous manner that
the often ad hoc physical approach adopted in previous articles [68, 10, 12, 13, 20, 23, 27, 3436]. In this
analysis, the forcing will be a function of time, leading to non-autonomous systems. This choice is due
to the non-autonomous systems having similar dynamics to the autonomous ones, as shall be shown in
this study, but being somewhat simpler to study.
Autonomous systems will be considered in Chapter 5, where the analysis shall follow one of the physical
approaches found in the literature [35, 36], in order to obtain simple approximate maps showing the
increase in amplitude.
Finally, the triple pendulum model, which has never been properly considered and is effectively an
improvement to the double pendulum one, shall be introduced in Chapter 6. The dynamics of this
models shall briefly be considered using both simulations and actual results from the experiments.
Some conclusions regarding the dynamics of swing will be given in Chapter 7, together ideas for possible
extensions to the study.
Also, a brief introduction to the methods used for Hill-type equations and the method of averaging, which
is used in several places in Chapter 4, are given in the Appendix A and Appendix B respectively.

10
Chapter 2

General Theory of Parametric


Resonance

Seeing as a large part of the study concerns parametic resonance and in particular in the context of non-
autonomous models, this chapter will introduce Hill-type equations and some of the theory necessary
for their study. It should be noted that all the models in this study involve pendula of different kinds
and are therefore inherently nonlinear. The full nonlinear systems can be studied in a number of ways
and has been done for the inverted pendulum with oscillating support by van Noort [30]. One possible
fist approach is to consider the quasi-nonlinear case by taking the next term after the linear one in the
Taylor approximation of the nonlinear terms, and then to assume this extra term to be sufficiently small
such that the method of averaging can be used [24]. However as the main concern of this study is the
initiation of motion from rest, the models will be considered in their linear approximations about the null
solution, for which Floquet theory is relevant.
For a more complete introduction to Floquet theory and Hill type equations, one can consult one of the
many texts available, such as Jordan & Smith [14], Verhulst [31], Arnold [2] and Rand [24].

2.1 Hill Equations and Floquet Theory


Hills equation is the name given to the class of homogeneous, linear, second-order differential equations
with real, periodic coefficients [18]. The importance of these equations to this study lies in the fact that
the equations of motion for the various non-autonomous models of the playground swing will all be Hill-
type or similar equations. The reason for some of the equations only being similar and not Hill equations,
is due to them containing inhomogenuities from direct forcing, as well as the parametric terms.
The canonical form of the well known Hill equation is

x + [a + p (t)] x = 0, x R, p (t) = p (t + T ) , (2.1.1)

where T is the minimum period of the periodic function p (t), a is a positive constant and means d/dt.
Most of the equations considered in this study will have sinusoidal forcing. Replacing the periodic function
in Hills equation by the sinusoidal one p (t) = b cos t, results in the equation referred to as the Mathieu
equation
x + [a + b cos t] x = 0, x R. (2.1.2)

11
Fig. 2.1: Possible combinations of Floquet multipliers for a conservative Hill equation. Left: Unstable,
with tr (M ) > 2. Right: Stable, with tr (M ) < 2. Figure taken from [2].

Such a sinusoidal forcing term can be considered to be a first approximation in a Fourier series, to a more
general periodic function p (t).
Another Hill-type equation which shall be seen throughout the study is Inces equation, which in its
canonical form can be written

[1 + a cos t] x + b sin t x + [c + d cos t] x = 0, x R, (2.1.3)

for various real constants a, b, c and d.


Comparing Inces and Mathieus equations, one notices that the latter is a special case of the former.
All the above equations may be transformed into first order systems, by introducing a new variable
which replaces the first derivative. The reason for making this transformation is that there exists a
fairly complete theory for first order linear ordinary differential equations with periodic, time dependent
coefficients. This is commonly referred to as Floquet theory. These first order differential equations are
of the form
x = A (t) x, x Rn (2.1.4)

with A (t) a continuous, T-periodic n n matrix.


It should be noted that, even though the coefficients are periodic, the solutions to systems such as (2.1.4)
are not necessarily so. Both unbounded and bounded solutions are possible, and even though it is not
always possible to find the exact solution, Floquet theory gives one some information.

Theorem 2.1.1 [Verhulst [32]] Floquets Theorem


Each fundamental matrix of (2.1.4) can be written as the product of two n n matrices

(t) = P (t) eBt , (2.1.5)

with P (t) a T-periodic n n matrix and B a constant n n matrix.

It is clear that both the existence of periodic solutions and the stability of the null solution depend upon
the eigenvalues of the matrix B. These eigenvalues are related to those of the Mondromy matrix, which is
the transformation matrix of the phase space over the period T . The relation being that the eigenvalues
of the matrix B are, upon choice, equal to certain exponents i , known as characteristic exponents of
the Monodromy matrix, where i = exp{i T } are the eigenvalues of this matrix, known as the Floquet
multipliers [31]. The Monodromy matrix is, by definition, given by

M = 1 (t) (t + T ) , (2.1.6)

where is the fundamental matrix, which is defined to be the n n identity matrix for t = t0 . The
Monodromy matrix, M , therefore reduces to (t0 + T ), simplifying further still with the choice t0 = 0.

12
The stability conditions are then found by considering the Wronskian, W (t) = det ( (t)), and using
Liouvilles Theorem [2], in the form
Z t 
W (t) = W (t0 ) exp tr (A (s)) ds . (2.1.7)
t0

Upon inspection, one notices that Mathieus equation, when written as a first-order system, has tr (A (t)) =
0. This is also the case for certain choices of the parameters in Inces equation. In this case, by (2.1.7),
det (M ) = W (2) = 1, where it is assumed that t0 = 0.
The eigenvalues must, by definition satisfy det (M I) = 0. Combining this with the condition on the
determinant of the Monodromy matrix from Liovilles Theorem gives

2 tr (M ) + 1 = 0. (2.1.8)

The precise computation of the Monodromy matrix and its trace is not analytically possible for most
systems. However, one can still make certain deductions without finding tr (M ). The first step is to see
that the eigenvalues, are given by

1
q
2
1,2 = tr (M ) tr (M ) 4 (2.1.9)
2

There are therefore three possible cases. These are |tr (M ) | < 2, |tr (M ) | > 2 and |tr (M ) | = 2.
The first case, gives complex conjugate Floquet multipliers which lie on the unit circle, see Fig. 2.1. All
solutions in the parameter region |tr (M ) | < 2 are therefore bounded, and this region is known as the
stable parameter region.
The second case, namely |tr (M ) | > 2, gives rise to real, distinct Floquet multipliers, both of which have
the same sign. The fact that the determinant of a matrix may be expressed as the product of its eigenval-
ues [14], together with the condition on the Monodromy matrix, tells one that 1 2 = 1. Therefore, one
of the multipliers must exceed unity in absolute value. This multiplier represents an exponential growth,
and therefore corresponds to an instability, causing this parameter region to be unstable.
The last case, |tr (M ) | = 2, may be spilt into two cases. The first tr (M ) = 2, which gives double Floquet
multipliers at = 1, and the second case tr (M ) = 2, which instead gives double, negative multipliers
at = 1. In these cases, the two multipliers are equal and there is one solution of period 2 and 4 for
the two cases respectively, and the other solution is unbounded [14]. These cases therefore give harmonic
and subharmonic stability boundaries in parameter space, and are known as transition curves.
If one represents these regions in the (a, b) parameter space, the resulting figure is known as an Ince-Strutt
diagram [9], and the instability regions are sometimes referred to as Floquet tongues. The diagram for the
canonical Mathieu equation can be seen in Fig. 2.2. This Figure was produce in Matlab using numerical
Floquet Theory, which involves integrating the system numerically, and checking the size of the Floquet
multipliers. The multipliers must then be on or within the unit circle in the complex plane for stability
to occur. The unstable regions in Fig. 2.2 show the magnitude of the greater eigenvalue, thus giving an
idea of the extent of the instability.
The curious choice of considering negative values of a is due to the physical significance which it assumes
when considering the pendulum of length l with a vertically oscillating support. In this case, a = g/ l 2 ,


where is the forcing frequency. One can therefore think of negative a as representing negative g, and
therefore a pendulum defying gravity, or more specifically upside down. The stability region for negative
a is therefore, interestingly enough, the stability region for such an upside down pendulum when subject
to vertical oscillations [1, 9].

13
5 5

4 4

= 0.5
3 3

= 0.4
a

a
= 0.3
2 2

= 0.2
= 0.1
1 1
=0

0 0
1 0 1 2 3 4 5 1 0 1 2 3 4 5
b b
Fig. 2.2: Left: Ince-Strutt diagram for the canonical Matheiu equation. Unstable regions coloured
according to magnitude of the greater Floquet multiplier, with darker colours denoting greater values.
Right: Stability diagram for the damped Matheiu equation (2.3.1), consisting of multiple Ince-Strutt
diagrams for different values of .

2.2 The Phenomena of Coexistence and Instability Pockets


The resonance tongues considered earlier are a common feature of all Hill-type equations. Two phenomena
that are related to the resonance tongues are the so-called phenomena of coexistence and of instability
pockets.
The first phenomenon involves the disappearance of resonance tongues, as seen in Fig. 2.3 which is
missing a resonance tongue at 2 equal to one. The disappearance occurs due to the coincidence of the
transition curves which define the resonance tongues. This phenomenon is visible in Inces equation for
certain choices of parameters [18, 21]. Mathieus equation on the other hand, which is a special case of
Inces equation, happens not to exhibit coexistence due to it having less parameters [21, 24]. One way
of understanding this phenomenon is by considering the Hill determinant, see A.1. The phenomenon
arises when one of the off-diagonal terms in the determinant vanishes, effectively dividing the determinant
into two parts, one of which is the same as the determinant for the other transition curve of the tongue
in question. Then coexistence occurs and an infinite number of possible tongues disappear. A complete
explanation of the phenomenon and the method to check for its occurrence are given in A.2.
Instability pockets on the other hand occur when the transition curves meet transversely, creating an
instability pocket [4, 5], see Fig. 2.3. This does not occur in the canonical Hill equations met previously,
but can be seen if one considers the near Matheiu equation

k
X

x + [a + bpc (t)] x = 0, pc (t) = cos t + cj cos (jt) , (2.2.1)
j=2

which is a Hill equation for which the harmonics up to the kth one have not vanished. In this case, one
gets any number up to k1 instability pockets, for appropriate, small values of {cj } [5]. This phenomenon
is found in the non-autonomous description of the double pendulum model when the forcing is taken to
be sinusoidal. In this case when expanding the trigonometric functions of the forcing term, one gets a
sequences with all harmonic terms present, similar to the one in (2.2.1). A complete understanding of
this phenomenon requires a geometric study of the equations involving Hills map, which assigns to each
parameter point (a, b) the corresponding Poincare matrix [4, 5]. A study of this kind is beyond the scope
of this study.

14
2
0.8

0.6

0.4

0.2

0 0

2 2
0 1/4 1 9/4 0 0.18 0.71 1.61 2.86 4.46

Fig. 2.3: Ince-Strutt diagrams showing coexistence, left, and instability pockets, right.

2.3 The Effects of Damping


The non-autonomous equations of motion that shall later be derived for the playground swing will not
include any of the dissipation actually present. However the effects of friction in the swing, specifically
at the support, and of damping due to air resistance, play an important role in the playground swing.
In general, considering dissipations leads to a process that can no longer be described by rigid body
dynamics, but one in which one should consider the motion of the air and the internal thermal states.
The case of the playground swing is simple enough that one can include an additional term to the
equations of motion to take account of these effects [16]. The system may therefore be regarded as being
acted upon by a force of friction, which depends on its velocity.
General dissipation forces may be included in the systems by using a Rayleigh dissipation function when
forming the Lagrangian [11]. When all that is required is a dissipative force proportional to the velocity,
this can also be added directly to the Euler-Lagrange equations of motion. If one then wants to reproduce
the physical motion of the swing, a value for the dimensionless damping coefficient is chosen from empirical
data.
Instead of considering each model separately, and adding damping, the general effects of adding a term
proportional to the first derivative in a Mathieu and an Ince equation will now be considered.
The canonical Mathieu equation, upon addition of a frictional term becomes

x + x + [a + b cos t] x = 0, x R, (2.3.1)

where is the damping coefficient.


The effect of damping on the Matheiu equation is to lift the resonant tongues from the a axis. The
amount that the tongues lift, is inversely proportional to the width of the undamped tongue [9]. This
can easily be seen by applying the method of averaging to the damped equation, where the damping
coefficient is scaled by the same small coefficient that multiplies b [24].
As the case of damped equations does not constitute the main part of the study, the approach taken
here, will instead be to use the magnitude of the Floquet multipliers to gain an insight into the relative
instabilities within a Floquet tongue. From the Ince-Strutt diagram for the undamped Matheiu equation,
Fig. 2.2, which includes the magnitude of the Floquet multiplier, one gets an idea of how damping will
affect the stability. Clearly the regions with greater Floquet multipliers are more unstable and will stay
that way for greater values of the damping coefficient. This is exactly what is seen in Fig. 2.2, where six
different values of are shown on the same Stability diagram.
The effect of damping on Inces equation, (2.1.3), is somewhat harder to study. The zero thickness of

15
7
= 0.15
amplitude 6

x
4

= 0.3
3

= 0.45
2

1
0.5 0.75
2
1 1.25 1.5
time
Fig. 2.4: Left: Amplitude of the directly forced pendulum as a function of frequency for various amounts
of damping. Right: Angular displacement as a function of time for direct (solid) and parametric (dashed)
forcing, with 2 = 1, F = 1.

the Floquet tongue makes numerical integration impossible, and the singular nature of coexistence makes
the simple addition of damping to Inces equation fruitless. However, it has been shown by Recktenwald
that the effects are the same as those for the Mathieu equation [25].

2.4 Contrast with Directly Forced Systems

The dynamics of directly forced pendulum is well known, see for example Baker & Blackburn [3], and shall
only be introduced briefly since some of the models contain direct forcing and as a means of comparison
with the parametric pendulum.
The linear equation for a directly forced, damped pendulum may be written

x + x + 2 x = F cos t, x R, (2.4.1)

where the time has been nondimensionalised and 2 = 2 /02 is the ratio of the forcing to the natural
frequency, see Chapter 3 for a discussion of the nondimensinalisation.
The amplitude of the system increases as 2 approaches one, which is the resonant state, as can be seen
in Fig. 2.4. This increase is without bounds if the damping is not present, or in the notation of (2.4.1)
when is zero [3].
The most interesting difference to note between the parametric and the direct forcing is the way in which
the amplitude of the two systems increases. For this purpose, one may consider (2.4.1) without damping
and the Mathieu equation in the form

x + 2 + F cos t x = 0,
 
x R. (2.4.2)

The amplitude of the direct system increases linearly, whereas the amplitude of the parametric system
increases exponentially [6], as can be seen in Fig. 2.4. This can be understood by analysing the solutions
of the two systems, in particular for the parametric system, this result is given by Floquets Theorem,
2.1.1, where the solutions are exponential functions.

16
Chapter 3

Development of the Two Canonical


Models

In this chapter, the two basic models that have been used in the literature to describe the motion of
playground swings shall be considered. The equations of motion shall be derived, using the Lagrangian
formalism, and then simplified using nondimensional analysis and and other relevant transformations.
In essence this chapter, by considering the two canonical models, contains a derivation of all the models
considered to date for swinging in the standing position, these may be found in the references [68, 10,
12, 13, 20, 23, 27, 28, 3436].

3.1 Parametric Oscillator Model

In this model, the child is represented by a point mass located at his centre of mass. Furthermore the
mass of the swing, the friction in the supports and the air resistance are all ignored. These simplifica-
tions give the well known mathematical pendulum. The parametric oscillation then comes from the child
varying the length of the pendulum by standing and squatting at given times.
Taking a fixed function of time to represent the motion of the centre of mass of the child leads to a
simplified non-autonomous description of the playground swing. The fact that this is an approximation
may be understood by recalling ones own experience on swings, and realising that one is influenced by
the motion of the swing, and moves accordingly whilst pumping. A non-autonomous description on the

0
1 0
1
0
1 0
1
0
1
0 0
1
1
0
1
0
1
0
1
0
1 0
1
0
1 0
1
0
1
0 0
1
1
0
0
1
1
0
0
1
1 0
1
0
1
0
1
0
1
0
1
0
1

0
1
0
1 l 0
1
0
1
0 0
1
1
0
0
1
1
0
0
1
l (t) 1
0 0
1
1
0
1 l1 0
1
0
1
0
1 0
1 11
00
0
1 0
1 00
11
00
11 m2
0
1
0 0
1
1
0
0
1
1
0
0
1
1
0 0
1
1
0
1
0
1
0
1
0
1 0
1
0
1
0
1 0
1
0
1 l2
0
1 0
1
0
1
0 0
1
m11
00 1
0
0
1
00
11 1
0 00
11 0
1
1 00
11 0
1
0
1
0
1 m11
00
1 0
1
0
1
0
1 0
1
0
1 0
1
0
1 0
1
0
1 0
1

Fig. 3.1: Representation of a playground swing as a simple pendulum of variable length and as a double
pendulum, to the left and right respectively.

17
other hand is closer to a system which is forced by a motor, which produces a forcing that is merely
dependent upon time. The reason for choosing to consider the non-autonomous system is therefore only
due to the analysis being slightly simpler.
Often the function of time used for the motion of the centre of mass of the child is the cosine function,
which leads to an Ince or Matheiu equation, as shall be shown below.
There is another simplification that shall be used in the modelling through out this report. The momen-
tary forcing that produces the changes, in this case of the length, also provides a change in the effective
gravitational field of the pendulum [3]. The gravitational field would become g (t) = g0 l, where g0 is the
original gravitational field. These changes shall be ignored in what follows, and a constant gravitational
field will be used instead.
Now to derive the equation of motion for this system, the Lagrangian formalism is well suited as this
removes the need to consider the forces involved. Using a polar description as the generalised coordinates,
see Fig. 3.1, one can easily see that the kinetic energy of the system is given by

1 2 
T = m l + l2 2 ,
2

where a dot implies differentiation with respect to time. If the reference level, at which the potential
energy is zero, is then fixed at the swings support, the potential energy is found to be

V = mgl cos .

Combining the two in the usual manner gives the Lagrangian of the system, and the equations of motion
are found using the Euler-Lagrange equations. Only one of these equations is required, namely the one
for , as the length will be used as a forcing term. The length will therefore be a function of time,
specifically l (t) = l (1 + a (t)), where l is the average length, and a (t) is a periodic function of time,
and is a small constant. This turns the system into what is typically refered to as a one and a half
degree-of-freedom system. The equation of motion is therefore found to be

l + 2l + g sin = 0, (3.1.1)

where the mass times the length has been factored out.
The next step is to nondimensionalise in order to simplify the equation such that it becomes more
transparent and the relative magnitude of each term may be seen [17]. In order to do so the dimensionless
time = t shall be used, where parametric forcing frequency. The reason being that the resulting
equations of motion are of a similar form to the canonical Ince equation. The length r ( ) = l ( ) /l is
also introduced and the equation of motion becomes

r + 2r + 2 sin = 0, (3.1.2)

where denotes differentiation with respect to and 2 = 02 / 2 is the ratio of the natural and the
forcing frequency, and the natural frequency is given by 02 = g/l.
It should be noted that the equation of motion contains the first derivative of the angle . This term is
present due to the length being a function of time, and is often neglected when deriving the equation of
motion of a swing by simply replacing the length in the equation for a simple pendulum by a function of
time. This term leads affects the stability of the null solution, as shall be shown in Chapter 4.
Now, there is a transformation that can be made in order to remove this first derivative [6], and put
the equation of motion into the standard form [18], which in the linear case is a Hill equation. The

18
transformation is given by
 
1 d 2r
= exp A ( ) , where A ( ) = .
2 d r

It is important to notice that the nondimensional length, r is never zero, since is chosen to be smaller
than one and that the transformation does not alter the stability. Taking the constant of integration to
be unity, the transformation reduces to
= r. (3.1.3)

The equation of motion therefore becomes


 
r
+ 2 sin

= 0. (3.1.4)
r r

This last form of the equation of motion is the one that will be used in the analysis of the system, mainly
in keeping with what has been done in the literature and in order to show certain features which have
been overlooked in the past.
In order to simplify the equation of motion for analysis, small angle motion is considered. This lineari-
sation about the null solution of our system, gives information about the initiation of motion, which is
of great interest. When linearising one uses the fact that the small angle limit, as approaches zero,
represents the same limit of approaching zero, in the original and transformed equations. The linearised
version of the transformed equation is therefore

rz + 2 r z = 0 + O z 3 ,
 
(3.1.5)

where the independent variable has been replaced by z. This is seen to be an Ince equation if one considers
the forcing function, r ( ) = 1 + cos , namely

(1 + cos ) z + 2 + cos z = 0 + O z 3 .
 
(3.1.6)

The equation is then usually simplified further by considering small amplitude forcing such that the
equation can be transformed into a Mathieu equation by only considering linear terms in

z + 2 + 1 2 cos z = 0 + O 2 .
   
(3.1.7)

These equations shall be analysed in Chapter 4, however there are some important features that should
be noticed. Firstly, before the truncation, Eq. (3.1.6) is an Ince equation, and the truncation for small
leading to a Mathieu equation is only a first approximation. Secondly, even in the Matheiu equation
the parametric term is different to what is seen in the literature, and leads to the disappearance of the
2 = 1 tongue. The reason for this is again the extra first derivative term that is usually neglected.

3.2 Double Pendulum Model


The concept of a compound pendulum model for the standing case of a child pumping a swing was
introduced by Gore [12, 13]; effectively as a triple pendulum, even though no explicit mention of this
was made. Later on McMullan [20], considered a double pendulum model. The compound pendulum
model was originally proposed to explain the initiation of motion from rest, which is not possible in the
parametric model. It was later argued that this model is actually closer to the motion performed by
children on swings for all amplitudes [7, 8].
This model concentrates on the leaning back and forth of the child in order to cause movement in the

19
swing. This leads, in the simplest case, to a compound double pendulum given a few assumptions. The
child is again taken to be a point mass located at his centre of mass, which simplifies the system by
removing the necessity of considering moments of inertia. The swings seat is then taken to be a pivot
and considered to be a point mass. This configuration can be clearly seen in Fig. 3.1, where m1 is the
seats mass, and m2 is the childs mass. The lengths therefore represent the swing length, l1 , and the
distance of the childs centre of mass from the seat, l2 .
This model is the well known double pendulum, which consists of two parts. These parts shall be referred
to as the inner and outer arm, depending on their distance from the pivot, when the system is at rest with
both parts hanging vertically downwards. It should be noted that in this study the angle of the outer
arm, , is measured from the inner arm, rather than the vertical as is more common in the literature.
In terms of equations of motion, any choice of angles would be fine, in the case of the swing however,
the double pendulum is actually constrained, by the child holding on, which means that the outer arm
oscillates relative to the inner one. This makes the choice used in this study, and by Case [7], the ideal
one when prescribing the forcing function of the system, which is the angle. The generalised coordinated
are taken to be the two angles, which leads to the following expression for the kinetic energy
 
1 1
M l1 + m2 l2 m2 l1 l2 cos 2 + m2 l22 m2 l1 l2 cos + m2 l22 2 ,
2 2
  
T =
2 2

where M = m1 + m2 is the combined mass.


The potential energy is found by fixing the reference level at the swings support

V = M gl1 cos + m2 gl2 cos ( + ) .

The simplest form of equation of motion is obtained by assuming that the angle is a forcing term, that is
an explicitly known function of time, for example by setting (t) = A (1 cos (t)). The Euler-Lagrange
equation for is therefore

M l12 + m2 l22 2m2 l1 l2 cos + 2m2 l1 l2 sin + M gl1 sin + m2 gl2 sin ( + )

(3.2.1)
+ m2 l22 m2 l1 l2 cos + m2 l1 l2 sin 2 = 0,


This equation must now be nondimentisionalised. Similarly to what was done for the parametric model,
the forcing frequency will be used when nondimensionalising the time, giving the transformation = t.
The equation of motion therefore becomes

(1 + 2 cos ) + 2 sin + 2 sin 2 sin ( + )


(3.2.2)
+ ( cos ) + sin 2 = 0,

where 2 = 02 / 2 , and = m2 l2 /M l1 and = m2 l22 /M l12 are the dimensionless parameters.


Similarly to what was done for the parametric model, the equation of motion can be linearised

(1 + 2 cos ) z +2 sin z + 2 (1 cos ) z


(3.2.3)
= ( cos ) sin 2 + 2 sin + O z 2 ,


where the independent variable is z.


Recalling that is the forcing function, one clearly notices the parametric terms and also the direct terms,
which are the inhomogenuities of the equation. The equation is still not of the Hill-type for the choice
of mentioned previously, what can be done however is to Taylor expand all the functions of , which is
exactly what shall be done for the analysis.

20
Chapter 4

Analysis

This Chapter contains an analysis of the models derived in the previous Chapter. Seeing as the full
equations of motion are not readily integrable in terms of elementary functions, the equations considered
shall be the linear ones for small or moderate amplitude motion. Even when considering the linearised
case, the equations of motion do not necessarily become integrable, they do however simplify enough
for the analysis. Also linearisation about the null solution of our systems, the swing hanging vertically
downwards, gives information regarding the initiation of motion, and is therefore particularly important.
In analysing the models, the instabilities that arise due to parametric resonance, and cause the increase
in amplitude of the swing, will be considered. The instability regions in parameter space will then be
found, and Ince-Strutt diagrams will be created. This will be followed by other relevant analysis, for
example of the direct resonance, depending on which model is being considered.

4.1 Parametric Oscillator Analysis


In keeping with tradition, the analysing of the parametric model shall involve the transformed equation,
Eq. (3.1.4). It should be understood however that the two equations are equivalent. The linearised
version of the transformed equation of motion is considered, namely

rz + 2 r z = 0 + O z 3 .
 
(4.1.1)

4.1.1 Instability Analysis

Before proceeding with the analysis, one needs to check the Liouville condition (2.1.7) for the equation
to find the solutions that define the transition curves. This is easily done by rearranging Eq. (3.1.5) into
a first order system and checking the condition. This check does not require a specific function for r (t),
as the trace is zero, and the transition curves are defined by the harmonic and subharmonic solutions,
namely the solutions with period 2 and 4.
Following what is done in the literature [2, 6, 9], the forcing function is taken to be r ( ) = 1 + cos and
the small parametric amplitude equations, which are first order in , shall be considered first.
Dividing the equations of motion by r, and expanding in a Taylor series in , keeping only the terms to
first order gives
z + 2 + 1 2 cos t z = 0 + O 2 ,
   
(4.1.2)

where the nondimensional time, , has been replaced by t for convenience.


This is clearly seen to be a Mathieu equation. The first thing to notice is that the equation has tongues

21
0.3

0.8

0.6
0.2 U S U S U S U S

0.4

0.1

0.2

0 0

2 2
0 1/4 1 9/4 0 1/4 1 9/4

Fig. 4.1: Left: Ince-Strutt diagram for the truncated parametric oscillator (4.1.2), using transition curves.
Right: Ince-Strutt diagram for the full parametric oscillator (4.1.1).

that emanate from the points 2 = n2 /4, n Z on the axis in parameter space. Next, by simple
inspection of Eq. (4.1.2) one sees that for 2 = 1, the parametric term disappears leaving the equation for
simple harmonic motion, such that there is no resonance tongue. This disappearance is interesting as this
tongue also disappears for the full equation, Eq. (4.1.1), as a result of coexistence. Now that the general
features of the Floquet tongues have been established, using regular perturbations or any other method
introduced in Appendix A.1, one can find the transition curves as regular expansions in . Forgetting for
a moment that the actual equation itself is only of first order, these equations are given up to third as:
1
0 2 + O 4 ,


2





1 3 33 2 381 3


+ O 4 ,



4 8 128 2048

2 = (4.1.3)

1, exactly,







9 + 25 2 125 3 + O 4  .



4 256 2048

These calculations have been performed in Maple and it has been checked that the leading order terms
agree with those in the literature (eg. [14, 24]). The Ince-Strutt diagram for the system, produced using
these equations, is shown in Fig. 4.1.1.
However, the transformed equation for the parametric swing can be analysed without resorting to approx-
imations for small parametric oscillations. This is clearly seen if one replaces the nondimensional length,
r, by the chosen parametric forcing function in Eq. (4.1.1), in which case one gets an Ince equation,
namely
(1 + cos t) z + 2 + cos t z = 0 + O z 3 .
 
(4.1.4)

The reason therefore, for considering the truncated equation, Eq. (4.1.2), was partly due to the fact this
is the one studied in the literature and partly due to the fact that often when analysing the system, with
the method of averaging for example, one must make this approximation.
As the equations is now an Ince equation, and also due to the interesting disappearance of the 2 = 1
tongue for the small parametric oscillation case, one must check for coexistence. The easiest way to check
for coexistence is to use the conditions, Eq. (A.2.6), given in A.2. Substituting the coefficients of the
Ince equation, which when written in their canonical form are a = , b = 0, c = 2 and d = , into the

22

0.2

v 1 3
2 = +

z
4 8 0

u 0.2

S S 0.2
z
0 0.2

v v
U 0.4

0.2

z
u u 0

0.2

1/4 2
0.4

0.8 0.4 0
z 0.4 0.8

Fig. 4.2: Left: Sketch of slow flow phase portraits for the parametric model, Eq. (4.1.2), shown over the
Ince-Strutt diagram. U and S denoting stable and unstable regions. Right: numerically computed phase
portraits for Eq. (4.1.2), showing a stable and an unstable case.

condition for coexistence gives



Q (m) = (1 m) (1 + m) , m Z,
2
(4.1.5)

P (m) = 4m2 4m 3 ,

m Z.
8

Coexistence, therefore clearly occurs as Q (1) = 0. Recalling that Q appear in the determinants for
the even tongues, one notices that the only two even tongues that survive are those emanating from
2 = 0, 1. For the sake of comparison, the equations for the same tongues as those considered earlier for
the truncated equation are found, with the aid of Maple, to third order in :



0, exactly,



1 3 15 2 45 3


4

4 8 + 128 2048 + O ,









= 1, exactly,
2
(4.1.6)


9 135 2 45 3


+ O 4 ,




4 256 2048








4, exactly.


The general Ince-Strutt diagram, which clearly shows the disappearance of the tongue at 2 = 1, can be
seen in Fig. 4.1.1. The transition curves, which have been placed over the figure, are clearly seen to be
approximations that are no longer valid for large values of .
4.1.2 Analysis using the Method of Averaging

The parametric pendulum model is a simple one, in that it only contains parametric terms. The previous
instability analysis therefore gives a good understanding of the stability of the null solution, making
further analysis using the method of averaging unnecessary, but also ideal to get an understanding of the
effect of parametric forcing on the slow flow equations. The double pendulum model will be analysed

23
using the averaging method, making this analysis useful in order to have similar results for the two
models.
This analysis shall only contain a few steps of the procedure, which can be found in Appendix. B, and
the results.
The first step is to put Eq. (4.1.2) in the correct form for averaging, which is

z + 2 z = 2 1 cos t z + O 2 .
   
(4.1.7)

The next step is to replace 2 by n2 /4 + where n is an integer, to study the system about the Floquet
tongues, as is done by Tondl et al. [29]. Finally, one has the choice of which transformation, and therefore
slow flow equations to consider. The choice in this case, as always, being the linear combination of cos t
and sin t, which gives rise to the following slow flow equations
 
3
u = v,
8
  (4.1.8)
3
v = + u.
8

The dynamics of the slow flow equation is straight forward, as can be seen from the equations, or from the
phase portraits shown in Fig. 4.2. This figure also contain two phase portraits of Eq. (4.1.7), computed
numerically, for the stable and unstable case.

4.2 Double Pendulum Analysis


The linearised equation of motion of the double pendulum model was found to be

(1 + 2 cos ) z +2 sin z + 2 (1 cos ) z


(4.2.1)
= ( cos ) sin 2 + 2 sin + O z 2 .


This equation has both parametric and direct forcing terms, in the form of time dependent coefficients
of the generalised coordinate and inhomogenuities respectively.
The forcing function is taken to be (t) = A (1 cos t), where the nondimensional time variable has been
replaced by t for convenience and A is twice the amplitude, and realistically would not take values much
greater than /18, see 1.2.
For this choice of forcing, the equations of motion are not of the desired Hill-type. What is done is therefore
to assume that the amplitude of the forcing, A, is small and then to Talyor expand all functions of . This
leads to an inhomogeneous generalised Ince-like equation, similar to those studied by Recktenwald [25],
specifically

X
X
X
X
ai cos (it) z + bi sin (it) z + ci cos (it) z = di cos (it) , (4.2.2)
i=0 i=0 i=0 i=0

where ai , bi , ci and di with i Z are polynomials in A, which shall henceforth be replaced by , and at
times also of , and 2 . The equation is not a generalised Ince equation both due to the inhomogenuity
and the fact that a0 is not unity.
As an initial study, one may consider the equation up to second order in , as was done also by Case &
Swanson [7,8]. This choice of truncation is a particularly important one, as it is the smallest order which

24
2

A
A

0 0

2 2
0 0.18 0.71 1.61 2.86 4.46 00.18 0.71 1.61 2.86 4.46 6.43

Fig. 4.3: Ince-Strutt diagrams for the homogeneous part of the double pendulum model (4.2.1), and the
truncated version (4.2.3), with = 0.3 and = 0.1, to the left and right respectively.

maintains both the direct and the parametric forcing found in Eq. (4.2.2)
   
3 2 2
1 + 2 (4 cos t cos (2t)) z + 2 [2 sin t sin (2t)] z
2 2
   
3
+ 2 1 1 + 2 2 (4 cos t cos (2t)) z = + 2 cos t + 2 2 + O 3 .
 
4 4
(4.2.3)
From this truncated equation of motion, one clearly sees that the direct forcing is of first order in ,
whereas the parametric forcing is of second order, which is a result already known to Case [7].

4.2.1 Instability Analysis

The equations of motion for the double pendulum model, Eq. (4.2.1), contain both direct and parametric
forcing terms, such that an instability analysis similar to the one conducted for the parametric model is
no longer possible. Instead the full system can be analysed using the method of averaging, which shall
be done in the next Section. One can still consider the stability regions and find their features by briefly
forgetting the inhomogeneous part of the equation, such that only the parametric terms are considered.
Firstly, one must locate the points in parameter space from which the Floquet tongues emanate. These
points lie on the axis with equal to zero, therefore by substituting = 0 into Eq. (4.2.1) and requiring
that the resulting equation is periodic, one finds that

n2 (1 + 2)
2 = , n Z, (4.2.4)
4 1

where the Liouville condition has been checked and the tongues are given by the solutions of period
2 and 4. Now, the Taylor expansion of the equation of motion, Eq. (4.2.2), suggests the presence of
instability pockets, as the full equation of motion is effectively a near Ince equation with all harmonics
present, see 2.2. The number of pockets is limited by the number of harmonics, leading one to expect
there to be an infinite number of instability pockets. These multiple instability pockets are clearly seen
in Fig. 4.3.
In the truncated equation of motion, Eq. (4.2.3), on the other hand, only two harmonics are present,
allowing for only one instability pocket, as seen in Fig. 4.3. It should be noted that the Ince-Strutt
diagram for the truncated equation of motion has been computed for large values of , for which the
equation is no longer valid, which partly explains the differences with the Ince-Strutt diagram for the full
equation of motion.

25
4.2.2 Analysis of Full System using Averaging
The method of averaging, specifically second order averaging, allows for the analysis of the complete
equation of motion, and in so doing also allows one to find the transition curves.
One starts by rearranging Eq. (4.2.3) such that its in the correct form for averaging, see Appendix B. This
is done by removing the coefficient from the second derivative of the generalised coordinate by dividing
through and Taylor expanding in . Keeping terms up to second order in , gives

(1 )  2
z + 2 + 2 cos t
 
z =

2
   
3 2 1 (4.2.5)
(2 sin t sin (2t)) z + cos t + cos (2t) z
4 4
+O 3 ,


where = 1 + 2, has been introduced to simplify the equations.


Following Tondl et al. [29], one considers a small parameter space about the tongues, namely

(1 ) n2
2 = + 1 + 2 2 , n Z. (4.2.6)
4

As was seen previously the location of the tongue tips depends on both and , however this dependence
is not studied, and Eq. (4.2.6) is rearranged and substituted into Eq. (4.2.5), in order to eliminate ,
resulting in

n2
z = F1 + 2 F2 + O 3 ,

z +
4
n2
F1 = + (1 + cos t) 1 z,
1 4
3n2
  
1 1
F2 = (1 + cos t) 2 + 1 + cos t cos (2t) z (sin t sin (2t)) z .
1 1 16 4
(4.2.7)
This equation is now in the correct form to perform averaging, which is done choosing the linear com-
bination of cos t and sin t representation, as in Appendix B. The calculations are trivial, as seen in
Appendix B, but computationally demanding, and where therefore done using Maple and not included.
The slow flow equations were found for both the tongue emanating from n = 1 and n = 2 in Eq. (4.2.6),
as they are expected to give different different results, since direct forcing is present at 1 : 1 resonance.
The result for n = 1 is

u = 1 v + 2 (2 1 ) v + O 3 ,


v = 1 u 2 [(2 + 2 ) u 1 3 ] + O 3 ,


where,
64 2 32 134 + 35 + 67 (4.2.8)
1 = ,
64 ( 1)
2
64 32 122 + 29 + 61
2 = ,
64 ( 1)
2
1 272 264 264 + 256
3 = .
64 ( 1)

26
The result for n = 2 is instead

2
2 12 4 v + O 3 ,
 
u = 1 v +
2 2
6 i 2
  
h 1
2 + 12 + 5 u + 6 1 + + O 3 ,
 
v = 1 u +
2 4 2 4
where, (4.2.9)
16 2 8 20 + 5 + 7
4 = ,
16 ( 1)
16 2 8 32 + 11 + 13
5 = ,
16 ( 1)

6 = .
1

By simple inspection of the slow flow equations for the two cases, one immediately notices certain dif-
ferences. Most importantly the shift of the equilibrium point, due to the direct forcing is clearly seen at
O () in the slow flow equations about the tongue at n = 2, as expected.
The 1 : 1 resonance equations, Eq. (4.2.9), shall now be analysed, however the same analysis can be
performed for all tongues. Firstly one can find the equilibrium point of the slow flow and use it to shift
the coordinates such that the equilibrium point coincides with the null solution. The equilibrium is found,
by requiring no flow, to be
 
1 6 (1 + (1 + 41 ))
(u0 , v0 ) = , 0 . (4.2.10)
4 1 + (2 + 12 + 5 )

This implies that the desired shift is (u, v) (w + u0 , v), simplifying the equations of motion to
"  #
0 1 + 2 12 4
z + O 3 , z = (w, v)T ,

z = (4.2.11)
2 1 2 + 12 + 5

0

from which one can easily study the stability of the equilibrium point, by considering the eigenvalues of
the system. Also, one can find the transition curves by seeking the constant solutions of the slow flow
equations, which by definition give the periodic solutions of the original quasi-Ince equation. Firstly, one
looks at the equation to first order in , for which one gets that the eigenvalues satisfy, 21 + 12 = 0. To
obtain constant solutions, one must set 1 = 0, and for all other values, the equilibrium point is a centre,
as the eigenvalues are purely imaginary. This result is to be expected as the parametric terms are of
order 2 and the averaging results for the parametric pendulum showed that centres are present outside
the tongues, see Fig. 4.2. Now, substituting 1 = 0 into the slow flow equations and considering them to
order 2 , one find that the eigenvalues satisfy, 22 + (2 4 ) (2 + 5 ) = 0. This condition finally gives
the transition curves
1 + 2 1 + 2 ,
4
2 = (4.2.12)
(1 ) 1 2 5 .

Differently to what has been seen for the Floquet tongues until now, the coefficients in the two transition
curves are not the same, possibly due to the higher harmonics present and the instability pockets.
As expected, within the tongue, the equilibrium point of the slow flow equations is a saddle point, as was
also seen for the parametric model, Fig. 4.2.

27
Chapter 5

Autonomous Description and


Analysis

The two canonical models for the playground swing shall now be considered to be autonomous systems.
This is a more realistic description of the playground swing, as was mentioned earlier, since the child is
influenced by the motion of the swing [36]. Generally in the autonomous description, the pumping of the
swing is taken to consist of instantaneous movements, leading to Heaviside-type forcing. This however
can only be justified as a means to simplify the analysis, as the pumping of a swing consists of smoother,
almost sinusoidal movements, see 1.2.
The autonomous systems may be analysed in a number of ways, the one generally preferred in the
literature involving the study of the increase in energy and of the work done by the child [10, 28, 34].
Another, effectively energy based, approach is the one used by Wirkus et al. [35, 36]. This method
involves studying the Euler Lagrange equations, and finding the increase in amplitude of the swing, by
appealing to the conservation of angular momentum and at times also of energy.
The two approaches and the results are similar, however the approach taken by Tea & Falk [28] and
others is felt to give the best physical understanding of the dynamics as it considers the changes in the
energy of the system, which in physical terms are the cause of the increase in amplitude. Having said
this, the method chosen for this study is the one used by Wirkus et al. [35, 36], as it is felt to be slightly
more rigorous. This method, together with the correct physical understanding, is therefore believed to
be the better one of the two methods.
The analysis of the parametric model follows that of Wirkus et al. [35], however that of the double
pendulum model was not considered in their article and is slightly more complicated.

5.1 Parametric Oscillator Model

The motion performed by the child to pump the swing involves the child standing at the lowest point of
the motion and crouching when the swing reaches its maximum amplitude. The changes in length of the
pendulum that represents the swing and child are assumed to take place in a short time t.
One starts with the Euler-Lagrange equations derived in Chapter 3

d 2 
l + gl sin = 0. (5.1.1)
dt

The first effect on the swing to consider is the one from the child standing at the bottom of the motion,
where || , with a small constant. The time period in which this change occurs is given by 0 t t

28

an+1

an+1

2 2

00 /2 00 /2
an an
Fig. 5.1: Amplitude maps for autonomous descriptions. Left: Parametric model, with = 0.2. Right:
Double pendulum model with direct forcing (dashed) and combined direct-parametric forcing (dash-dot),
where = 0.65, = 0.13 and = 0.26.

and the Euler-Lagrange equation is integrated over this time period, giving
Z t
ls2 lc2 0 = gl sin dt, (5.1.2)
0

where the subscripts c and s denote crouching and squatting respectively, and and 0 denote the
velocities at the times t and 0 respectively.
Now, since | sin () | , the integral on the right-hand side is O (), and therefore tends to zero as as
0, which amounts to having no torque and therefore conservation of angular momentum, as can be
seen by considering the forces on the system. For this instantaneous change in length Eq. (5.1.2) can be
rearranged such that
 2
lc
= 0 . (5.1.3)
ls
The other change in the system occurs when the child crouches, increasing the effective length of the
system, which occurs when the amplitude of the swing is greatest. During the time in which this change
occurs, the velocity of the system is small, specifically || . Hence during an instantaneous change,
the velocity is null, and the the kinetic energy is conserved. To see whether the swings amplitude is
affected, one must integrate the Euler-Lagrange equation from time t1 to some time t t1 + t, which
gives Z t
l2 (t) (t) ls2 1 = gl sin d. (5.1.4)
t1

Next, dividing by l2 (t) and integrating from t1 to t1 + t

t1 +t  2 t1 +t t
ls 1
Z Z Z
1 1 dt = 2
gl sin d dt. (5.1.5)
t1 l (t) t1 l (t) t1

The integrals on the left and the right are of order and t and therefore vanish in the limit as and
t approach zero, leaving = 1 .
Therefore, the only change in the system is due to the child standing. To find the change in amplitude,
one must notice that the energy of the system is conserved when the childs position is unchanged, and
can be written
1
E = l2 2 gl cos . (5.1.6)
2

29
Considering the energy, Eq. (5.1.6), over a time interval from just after the child stands to just before
the child crouches, at the top of the swing,s motion, one gets

1 2 2 1
l gls cos (0) = ls2 02 gls cos , (5.1.7)
2 s 2

where represents the amplitude of the swing after the child stands up. This equation may be rear-
ranged to give
ls 2
1 cos = . (5.1.8)
2g
Similarly, one can find the amplitude that the child would have reached had he not stood by considering
the energy over a time interval starting with the child crouching, when = 0, and ending just before the
child stands up, when = 0. The equation being

lc 2
1 cos 0 = . (5.1.9)
2g 0

Finally dividing Eq. (5.1.7) by Eq. (5.1.8), and using the relation between the velocities, Eq. (5.1.3), one
gets
 3
1 cos lc
= , (5.1.10)
1 cos 0 ls
which may be simplified further, when the angle is sufficiently small, using the Taylor series for cosine
 3/2
lc
0 . (5.1.11)
ls

Considering successive oscillations of the swing produces a simple linear map for the amplitude, which
when considering a form for the length similar to that used in the non-autonomous description, may be
written in the standard form  3
1+
an+1 an , n Z, (5.1.12)
1
where a represents the amplitude, n is a count of the number of oscillations and the change in length is .
Firstly, one should notice the change in power due to the fact that the child pumps twice per oscillation
of the swing. Also, one notices, see Fig. 5.1, that the increase in amplitude is geometric, as was the case
also with the non-autonomous description of the parametric model.

5.2 Double Pendulum Model


Similarly to what was done for the parametric model, an autonomous description of the double pendulum
model shall be considered in which the forcing is instantaneous. The child is assumed to pump the swing
by leaning backward when the swing reaches the maximum negative angle and by leaning forward when
the swing reaches the maximum positive angle. The angle , see Fig. 3.1, therefore takes its maximum
value A while > 0 and is zero while < 0.
One again starts with the nondimensional Euler-Lagrange equations for the double pendulum model,
derived in Chapter 3, namely

d h i
(1 + 2 cos ) + ( cos ) + 02 [sin sin ( + )] = 0. (5.2.1)
dt

30
The jump in from zero to A is modelled as occurring in a time 0 t t, during which || , with
a small constant. Integrating the equation of motion from a time 0 to t t gives
h i Z t
(1 + 2 cos ) + ( cos ) (1 + ) 0 = 02 [sin sin ( + )] d, (5.2.2)
0

where the subscript 0 denotes values at time t = 0.


Next, dividing by the coefficient of and integrating again, from 0 to t, gives

t t t t
( cos ) 0 1
Z Z Z Z
0 + dt dt = 02 [sin sin ( + )] d dt,
0 ( 2 cos ) 0 ( 2 cos ) 0 ( 2 cos ) 0
(5.2.3)
where replaces 1 + .
Now, letting and t approach zero, as the pumping is assumed to be  instantaneous,
 one can neglect
2
the second and third integrals from the left, as they are O () and O (t) respectively. Both cases
being due to || , in the second integral as a direct consequence of this condition, and in the third as
an indirect consequence, using the fact that as approaches zero, approaches a constant value. Setting
the third integral to zero amounts to conservation of angular momentum, as the original integrand is the
torque.
The remaining integral can be solved using computer algebra software such as Maple, giving
"  1/2 !#t
1 1 + + 2
0 = arctan tan . (5.2.4)
2 [(1 + 2) (1 + + 2)]1/2 1 + 2 2
0

However, the angles, , in Eq. (5.2.4) are negative, but what is of interest is the increase of amplitude.
Replacing the angles by the amplitudes, denoted a, and substituting in the values of , which are assumed
to be small enough to justify using a Taylor expansion, reduces the equation to
 
A 1
a a0 = 1 . (5.2.5)
2 1 + 2

The exact same increase in amplitude occurs when the child leans forward when the swing reaches the
maximum positive angle, doubling the effect.
Considering successive oscillations of the swing gives the linear map
 
1
an+1 = an + A 1 , (5.2.6)
1 + 2

in which the angle increases by a fixed amount, as expected, see Fig. 5.1.
The increase in the amplitude map resulting from this analysis depends very much on the dimensionless
parameters of the model, and actually becomes negative for certain choices. This interesting phenomenon
is believed to be due to oversimplifications in the analysis, and leads one to question the use of this method
for more complicated models.
Wirkus et al. [35,36] also consider a combination of both the parametric and the direct pumping, arguing
that for large enough amplitudes this will give the optimum results. However this is not a method
employed by children naturally, and therefore discarded. What is interesting instead is to consider the
effect that a shift from leaning forward when the swing reaches the maximum positive angle to when the
angle is minimum, whilst still leaning back when the swing reaches the maximum negative angle. This
shift towards parametric forcing is of interests since the double pendulum has both parametric and direct
components to it, such that for large angles one expects the parametric forcing to be dominant. This
together with the fact that the forcing was seen to be almost sinusoidal in the experiments, 1.2, and

31
that Post et al. [23] actually found a slight shift towards parametric forcing, makes this an interesting
possibility.
The parametric forcing in the double pendulum is modelled as a jump in from A to zero, taking place
in a time 0 t t, during which || . Integrating the equation of motion from a time 0 to t gives
Z t
(1 + 2) (1 + 2 cos A) 0 = 02 [sin sin ( + )] dt, (5.2.7)
0

where the integral can be neglected as and t approach zero, again giving conservation of angular
momentum, simplifying the equation to

(1 + 2 cos A)
= 0 . (5.2.8)
(1 + 2)

The increase in amplitude is then found in the same way as for the parametric model, by using conservation
of energy, which in this case may be written
 
1 1
E= (1 + ) cos 2 + [ cos ] + 2 02 [cos cos ( + )] . (5.2.9)
2 2

Considering the energy over the time interval from just after the child leans forward, with = 0, to when
the swing reaches the maximum positive angle, gives
 
1 2
(1 + ) 02 [1 ] = 02 cos [1 ] . (5.2.10)
2

Similarly, considering the energy over the time from just after the child leans backwards, when the swing
reaches its maximum negative angle, to when the swing has zero amplitude, gives
 
1 2
02 [cos 0 cos (0 + A)] = (1 + ) cos A 0 02 [1 cos A] . (5.2.11)
2

These last two equations give the desired relations between the angular velocities and the angles, and
can therefore be rearranged and substituted into Eq. (5.2.8) to replace the angular velocities

1 1 + 2 cos A
1 cos = [1 cos 0 + cos (0 + A) cos A] . (5.2.12)
1 1 + 2

One can now simplify the equation slightly by using Taylor expansions, for small angles , even though
the double angle cosine limits the simplification possible to

1 1 + 2 cos A  2
2 =

0 (1 cos A) 20 sin A , (5.2.13)
1 1 + 2

where one needs to remember that 0 < 0.


When considering successive oscillations of the swing, the effect of this parametric forcing together with
that of the direct forcing can be seen by considering the amplitude map

1 1 + 2 cos A h 2
i
a2n+1 = (an + k) (1 cos A) + 2 (an + k) sin A , (5.2.14)
1 1 + 2

where k is the constant from the direct forcing, see Eq. (5.2.5). This amplitude map, together with the
one from simple direct forcing is shown in Fig. 5.1. Again, the results from this analysis are felt to be
limited by its simplicity, making the comparison between these two possible methods of pumping not
truly possible.

32
Chapter 6

Further Modelling and Simulations

In this chapter a relatively new model will be introduced, namely the triple pendulum model. This is seen
to be a refinement of the double pendulum model, and is felt to be one of the most needed improvements
out of the many possible improvements and aspects of real playground swings that can be considered.

6.1 Derivation and Simulations of the Triple Pendulum Model


The triple compound pendulum model for a playground swing was first considered by Gore [12, 13], as
mentioned previously. The reason for introducing this model was due to the inability of the parametric
model to account for the starting of motion from rest. However this model was only used as a means to
explain the initial displacement of the swing from absolute rest and was not given much attention. It
was later replaced by the double pendulum model in its various forms.
The model, as usual, does not consider frictional forces or damping. It is essentially an extension of the
double pendulum model, considered in 3.2. The only difference considered here is that of an extra
mass placed along the swings support at the point where the child holds on. With this modification, see
Fig. 6.1, one obtains a constrained triple pendulum.
For the sake of simplicity and as a first approximation, the forcing considered here is the same as the one
used for the double pendulum model, in which the child leans back and forth such that is the forcing
function, see Fig. 6.1. There are however two improvements that could be made to the forcing function.
The first would be to introduce the fact that the child pumps both by crouching and by leaning, as seen
in the experiments and Fig. 1.3, such that both and l3 are forcing functions. One could also consider
the actual force produced by the child in pulling himself up from the backwards leaning position, and
replace the forcing angle with this other forcing. This improvement would involve considering explicit
forces, and would require the introduction of constraints, for example on .
The the kinetic energy for the system, with the simple forcing function , can be written

1 1 1
M1 l12 2 + M2 l22 + m3 l32 m3 l2 l3 cos 2 + m3 l32 2

T =
2 2 2
     
+ M2 l1 l2 cos 2 m3 l1 l3 + cos 1 + m3 l3 l3 l2 cos ,

where M1 = m1 + m2 + m3 and M2 = m2 + m3 are combined masses and 1 = , 2 =


and 3 = + are combined angles.
The potential energy can be written

V = M1 gl1 cos M2 gl2 cos + m3 gl3 cos 3 ,

33
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0 l
1
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1
0
1 00
11
0 00
11
1
0
1
0 00 m1
11
1
0
1 11
00
0

1
0 00
11
1
0 00 m3
11
1
0
l2 1
0
1
0
1
0
1
0
1
0
10
1
0
l3
11
00 1
0
m11
00
00
11
2 1
0
1
0
1
0
1
0
1
0

Fig. 6.1: Representation of a playground swing as a triple pendulum.

where the zero potential energy level is set to be at the pivot of the swing.
The Euler-Lagrange equation of motion for and are then found to be
 
M1 l12 + M1 gl1 sin + C + S 2 2m3 l1 l3 sin 1 m3 l1 l3 cos 1 sin 1 2 = 0,

M2 l22 + m3 l32 2m3 l2 l3 sin + 2m3 l2 l3sin + M2 gl2 sin m3gl3 sin 3 + C S 2
 (6.1.1)

+m3 l3 (l3 l2 cos ) + l2 sin 2 = 0,

where C = M2 l1 l2 cos 2 m3 l1 l3 cos 1 and S = M2 l1 l2 sin 2 m3 l1 l3 sin 1 have been introduced to


simplify the equations.
The next step is to nondimensionalise the equations in order to simplify them and find the relative
magnitudes given by the dimensionless parameters. Similarly to what was done with the previous models,
the time is nondimensionalised to = t. The equations of motion therefore become

+ 2 sin + C + S 2 22 2 sin 1 2 2 cos 1 sin 1 2 = 0,




21 1 + 22 2 21 2 2 sin + 21 2 2 sin + 1 1 2 sin 2 2 2 sin 3 + C S 2




+2 2 (2 1 cos ) + 1 sin 2 = 0,
(6.1.2)
where 2 = 02 / 2 is the ratio of the natural frequency to the forcing frequency and the dimensionless
parameters are 1 = l2 /l1 and 2 = l3 /l1 , 1 = M2 /M1 and 2 = m3 /M1 . Also, C and S are the
nondimensional versions of C and S.
In order to study the initiation of motion from rest, these equations of motion could be linearised and
written in matrix form, which would clearly separate the parametric, direct and coupling terms. Instead,
in order to obtain an idea of the dynamics of the triple pendulum model and to compare the results with
the experimental ones, the full, nonlinear system shall briefly be analysed using numerical simulations.
This is easily done in Matlab using one of the built-in numerical integration schemes.
In keeping with the experimental results, the forcing angle was chosen to be a function of the swings
angle, or to be precise of the angular velocity, specifically

= A (1 H ( )) , (6.1.3)

34
0.8

,,
0.4

0.4

0 10
time 20 30

p/4

,,
0

p/4

time
Fig. 6.2: Left: Numerically produced flashlight traces. Right: Time series of (dotted), (dashed)
and (solid) from numerical simulation and experiment, above and below respectively. Values used for
simulations: 1 = 0.43, 2 = 0.21, 1 = 0.53, 2 = 0.31, A = /9

where H is the Heaviside function, and A is the amplitude of the forcing.


This forcing is clearly not the same as the one seen in the experiments, which was almost sinusoidal, but
is similar to the one considered when studying the canonical models in the autonomous description, and
turns out to be a first approximation that gives fairly good results, as seen in Fig. 6.2. An improved au-
tonomous forcing function could therefore be a similar, but smooth function such as the logistic function.
The numerically produced equivalent of the flashlight traces in Fig. 6.2, cf. Fig. 1.2, clearly shows that
the forcing is an oversimplification by the fact that the point m3 , which represents the centre of mass of
the child, follows a very different trajectory to that seen in the experiments.

35
Chapter 7

Conclusions and Outlook

Large part of this study of the dynamics of the playground swing, and in particular of the input of energy
into the system in order initiate motion, consisted of an in depth study of the two canonical models.
These models were analysed using similar techniques, allowing for an improved comparison, in order to
reach a better understanding as to which model best represents the playground swing.
While considering the parametric oscillator model, which is often the preferred one in the literature
[2, 6, 9, 27, 34], two important, but often overlooked features of the model were found. The first feature
is the extra term that appears due to the length not being constant in the derivation of the equation of
motion. This term is often overlooked, as the equation of motion is sometimes derived from the simple
one for the mathematical pendulum, by replacing the length by a function of time. However together
with the phenomenon of coexistence, which is the second less-known feature of the system, this term
affects the stability regions and cannot be ignored. Coexistence only occurs in the full equation of mo-
tion, Eq. (4.1.4), and not in the truncated equation for small amplitude parametric oscillations. The
reason being that this phenomenon may occur in Ince but not in Mathieu equations, due to the lack of
coefficients in the latter, see 2.2, meaning that the approximation made by truncating the equation
comes with its limitations and must be considered accordingly.
The next model that was considered was the double pendulum model, which in its simplest form involves
the child leaning backwards and forwards in order to pump the swing and is very different to the paramet-
ric model. The equation of motion for this model clearly shows both the direct and the parametric terms
expected, and upon simplification one further finds that these are of different orders in the amplitude of
forcing, namely first and second order. This result, which was already know to Case [7], together with
the fact that children usually pump a swing at its direct resonance and not the ideal parametric 2 : 1
resonance, leads one to conclude that the swing is effectively an example of a directly forced oscillator.
The parametric component of the forcing, which is expected to assume an increasingly important role as
the amplitude of the swing increases, as also seen in the experimental study of Post et al. [23], was found
to have an interesting Ince-Strutt diagram, see Fig. 4.3, due to the presence of instability pockets when
the forcing is sinusoidal.
The models were then considered using an autonomous description, which was found in the experiments
to be the more appropriate one. The two possible descriptions of the models were found to display similar
features, specifically the analysis of the parametric pendulum lead to an amplitude map that showed the
expected geometric increase in amplitude and the double pendulum analysis lead to a map that predicted
an arithmetic increase, also as expected. The method used to analyse the systems was taken from the
articles by Wirkus et al. [35, 36] and was found to be too simple even for a system such as the double
pendulum model, therefore precluding its use for more complicated systems such as the triple pendulum
model.

36
The results obtained from the analysis of the models were also validated by simple experiments of swing-
ing. These results clearly showed that the leaning back and forth method of pumping is the prominent
one, see Fig. 1.3, therefore supporting the double pendulum model, even though the non-rigid swing
supports make the system more similar to a triple pendulum. Interestingly, both the crouching and the
leaning methods of pumping were seen in the experiment, supporting the idea that a complete model
should include both [7, 36]. However, the parametric pumping was seen to play a secondary role such
that as a first approximation, a compound pendulum model without this form of forcing is felt to be
more than adequate. The experiment also gave results that seemed to validate the autonomous forcing,
as expected and seen in many of the models [10, 28, 35, 36], and also showed that the forcing is very much
harmonic in nature, thus rendering Heaviside-type forcing only an rough approximation.
The triple pendulum with a simple forcing from the child leaning back and forth was considered in Chap-
ter 6 and its equations of motion were derived and used for simple numerical simulations. The results
from this simple triple pendulum were seen to agree closely with those from the experiments, showing
that this model is probably the one which best represents the playground swing. In this respect, it is felt
that a large part of the possible further research could be done by analysing this model thoroughly and
possibly extending it. These extensions could include the addition of the so-called parametric forcing,
some of the other features seen in the experimental results and possibly explicit forces, from the child
pulling himself up to lean forwards, which could replace the somewhat artificial forcing angle. Ideally one
would be able to find an optimum model that displayed all the same results as the experiments, whilst
still being as simple as possible. With this in mind, what would be required is an in depth comparison
of the double and triple pendulum models in order to determine whether there are significant differences
and therefore whether the former, more simple model represents the playground swing sufficiently well,
which has been repeatedly claimed to be the case [7, 36].
Further analysis and modelling is felt to require further, more quantitative experiments, as the modelling
and analysis alone is at times not sufficient to predict and understand the swings dynamics completely.
The further experiments could easily be done using a method similar to the one used in this study, as it
was found to be easy both in terms of setup and analysis.
The direction taken in this study is however by no means the only one possible, another interesting
possibility for further study would be to extend the work of Piccoli & Kulkarni [15, 22], which analyses
the parametric oscillator model, and to consider the compound pendulum models using the theory of
optimal control to check whether the pumping of a swing in these models is also time optimal.

37
Appendix A

Analytical Methods for Parametric


Resonance

A.1 Methods to Find the Resonance Tongues


The transition curves of a Hill equation can be found using many different methods, as any method which
enables one to find periodic solutions will suffice. This is based on the knowledge gained from Floquet
theory, in Chapter 2, that the transition curves are given by the periodic solutions. Examples of methods
are found throughout the literature [14, 24, 37], a few possibilities being regular perturbations, harmonic
balance, multiple scales and the method of averaging. Some of these methods will be used throughout
the study and therefore briefly introduced here. In the explanation that follows, the canonical Mathieu
equation shall be used as an example.
The first method is that of regular perturbations, which gives approximations to the chosen order in b.
What is done is to assume small values of |b|, and then suppose that the transition curves are given by
a as a function of b. One then expands the variable a and the corresponding solutions of the Mathieu
equations in terms of b as follows

a = a (b) = a0 + ba1 + b2 a2 + ... , (A.1.1)


2
x (t) = x0 (t) + bx1 (t) + b x2 (t) + ... . (A.1.2)

These are then substituted into the Mathieu equation, and the knowledge that the solutions are periodic,
with period either 2 or 4, on the transition curves is used to get successive approximations to the
transition curves. To find the points on the a axis from which the Floquet tongues emanate, only the
first approximation is necessary, namely

O b0 : x0 + a0 x0 = 0.

(A.1.3)

Now since the harmonic and subharmonic solutions are the ones necessary, one requires that a0 = 41 n2 ,
n = 0, 1, 2, ... . This result agrees with what is seen in Fig. 2.2.
The second method uses harmonic balance, and again the knowledge that the solutions required are
periodic. One therefore expands the solution to the Mathieu equation as a Fourier series, in this case in
its complex form

1
X
x (t) = cn e 2 int . (A.1.4)
n=

38
This expansion being considered it covers both desired periodic solutions. Now replacing cos t by
1 it it

2 e +e and substituting the Fourier expansion into the Mathieu equation, one gets the follow-
ing equation
   
X 1 1 1 1
e 2 int bcn+2 + a n2 cn + cn1 = 0. (A.1.5)
n=
2 4 2

As this equation must hold for all t, the coefficients must all be be equal to zero, which gives the following
condition  
1 1 2 1
b cn+2 + a n cn + cn1 = 0, nZ (A.1.6)
2 4 2
This set of equations can be split into two sets, for n even and for n odd, which gives the conditions
for the harmonic and subharmonic solutions respectively. Now, these infinite sets of homogeneous linear
equations for the sequences {cn }, have non-trivial solutions if the infinite determinant formed by their
coefficients, known as the Hill determinant, is zero [14]. For the harmonic solutions, the determinant in
question is
... ... ... ... ... ... ...

... 1 1 1 0 0 ...


... 0 0 1 0 0 ... = 0, (A.1.7)

... 0 0 1 1 1 ...


... ... ... ... ... ... ...

where n = b/2 a n2 and a is not equal to n2 .




The determinant for the subharmonic solutions is not shown as it is similar. Now the transition curves
are defined by the relations between a and b which are found in these Hill determinants. The problem
however being that the determinants are infinite. One way around this is to use recurrence relations
between determinants of finite and increasing size, starting with the mid-determinant, see Jordan &
Smith [14].
Another more straight forward way to use the Hill determinants is to substitute a regular perturbation
expansion of a, in terms of b, into a truncated determinant [24,25]. This uses a combination of ideas from
harmonic balance and regular perturbations, and turns out to be highly effective.

A.2 Method to Check for the Presence of Coexistence


The presence of coexistence in an Ince equation can easily be checked by using the Hill Determinants.
This Appendix shall explain the occurrence of the phenomenon and therefore also the method to check
for its presence.
The canonical Ince equation shall be used as an example. This example requires the assumption that
the canonical coefficients a, b and d depend upon a small parameter , and further that these coefficients
are all equal to zero when is equal to zero. This ensures that the Floquet tongues emanate from the
caxis.
The method of harmonic balance introduced in A.1 is the one used to study this phenomenon. However,
one needs to check Liouvilles condition, (2.1.7), to decide which solutions define the transition curves.
This cannot be done directly on the canonical Ince equation, due to the coefficient of the second derivative.
The way to proceed is therefore to transform the Ince equation into a Hill equation, and then check the
condition for the new equation, as the transformation does not affect the stability. This is not done
here, as it can be found in both Magnus & Winkler [18] and Rand [24]. The result is that the transition
curves are defined by the solutions of period 2 and 4, so the solution is expanded in the following

39
trigonometric Fourier series
    
X 1 1
x (t) = an cos nt + bn sin nt . (A.2.1)
n=0
2 2

Similarly to what was shown before, one gets four sets of algebraic equations on the coefficients an and
bn . These equations are decoupled and deal with the separate cases of aeven , beven , aodd and bodd. The
even cases being the harmonic solutions and the odd cases the subharmonic ones.
It has already been shown that these equation have non-trivial solutions if the Hill determinants vanish,
giving the following conditions


c Q (1) 0 0

2Q (0) c 1 Q (2) 0



aeven :
0 Q (1) c 4 Q (3) ... = 0, (A.2.2)
0 0 Q (2) c9



...


c 1 Q (2) 0 0

Q (1) c 4 Q (3) 0



beven : 0
Q (2) c 9 Q (4) ... = 0, (A.2.3)
0 0 Q (3) c 16



...

1


c + P (0) P (1) 0 0
4
9
P (1) c P (2) 0
4

aodd : 25
0 P (2) c P (3) ... = 0, (A.2.4)

4
49


0 0 P (3) c

4
...

1


c P (0) P (1) 0 0
4
9
P (1) c P (2) 0
4

bodd : 25
0 P (2) c P (3) ... = 0, (A.2.5)

4
49


0 0 P (3) c

4
...

where the notation has been simplified (following Magnus & Winkler [18]) as follows

d + bm am2
Q (m) = , m Z,
 2  (A.2.6)
1 4d + 2b (2m 1) a (2m 1)2
P (m) =Q m = , m Z.
2 8

Comparison of the even determinants, (A.2.2) and (A.2.3), shows that if the first row and column of the
determinant for a are removed, then the remainder of the determinant is identical to that for b. This is
a significant observation, as it implies that if one of the off-diagonal terms vanishes, that is if Q (m) = 0,
for some integer m, then coexistence occurs and an infinite number of possible tongues will disappear.
This can be understood better via an example. Therefore, suppose, as Rand [24] does that Q (2) = 0,

40
giving the following even determinants


X X 0 0 0



X X X 0 0

0 X X X 0 ...
aeven : = 0, (A.2.7)


0 0 Q (2) X X


0 0 0 X X


...


X X 0 0 0



X X X 0 0

0 Q (2) X X 0
beven : = 0, (A.2.8)


0 0 X X X ...

0 0 0 X X


...

where the symbol X replaces all non-zero terms, and the reason for the priming will be clear later.
It is clearly seen that the vanishing of Q (2) disconnects the lower (infinite) portion and the upper
(finite) portion of the determinant. This gives rise to two possible ways in which these equations may
be satisfied, which shall be shown to both lead to the disappearance of an infinite number of Floquet
tongues:

For a nontrivial solution for the Fourier coefficients associated to the lower portion of the de-
terminant, this must vanish. Now, since the determinants are identical for both the coefficients,
coexistence occurs and the tongues emanating from c = 9, 16, ... all disappear. This ensures that
the coefficients an , bn for n = 6, 8, 10, ... will not in general vanish. However the upper portion of
the determinants will not in general vanish either, therefore one must use the knowledge that the
coefficients a0 , a2 , a4 and b2 , b4 will not be zero, as they depend respectively on b6 and b6 . This
can be seen by considering the disconnected determinants, (A.2.7) and (A.2.8), where the terms
that cause this dependence are primed.

The other possibility is that the lower portion of the determinant is not zero, which requires that the
infinite sets of associated coefficients aeven , beven vanish. Also the upper portion of the system now
becomes disconnected. For a nontrivial solution of the coefficients associated to the upper portion,
one therefore requires the 3 3 and 2 2 determinants to vanish. These give polynomials, which
can be solved for c in terms of the other coefficients in the Ince equation (2.1.3). If the resulting
equations for c are real, these give the transition curves emanating from c = 0, 1, 4.

To check for coexistence in an Ince equation, one therefore needs to check whether any of the terms in
the associated Hill determinants vanish, which can be done using Eq. (A.2.6).

41
Appendix B

The Method of Averaging

The method of averaging is an approximation method which leads generally to asymptotic series [31].
The procedures involved in this method are all straight-forward, but also algebraically demanding. For
this last reason, the method will be introduced here in all generality, such that it can be referred to
when required. The details presented here are those for second order averaging as these contain enough
information for one to extend the method to an arbitrary order. For a more complete introduction, one
may refer to Verhulst [31] and Rand [21, 24], which are the sources used for this brief introduction.
The general equation to which this method is applicable is the following

x + n2 x = F1 (x, x, t) + 2 F2 (x, x, t) + O 3 , x = x (t) .



(B.0.1)

If = 0, the solution is known and can be written is a number of ways. Even though the choice is just
one of coordinates, it will make a slight difference in terms of analysis. The choice here is that of a linear
combination of cos (nt) and sin (nt). When is not equal to zero, one therefore seeks for a solution to
the equation, Eq. (B.0.1), of the form

x (t) = u (t) cos (nt) + v (t) sin (nt) ,


(B.0.2)
x (t) = nu (t) sin (nt) + nv (t) cos (nt) ,

where u and v are expected to be slowly varying functions of t.


It should be noted that, in order for the transformation, Eq. (B.0.2), to hold, one requires the following
condition
u sin (nt) + v cos (nt) = 0, (B.0.3)

found by differentiating the first transformation and demanding that it is equal to the second.
Now, substituting the transformation into our system, Eq. (B.0.1), and using the condition, Eq. (B.0.3),
to solve for u and v, one gets

2
= sin (nt) F1 sin (nt) F2 + O 3 ,

u
n n (B.0.4)
2
= cos (nt) F1 + cos (nt) F2 + O 3 ,

v
n n

where Fi = Fi (u cos (nt) + v sin (nt) , nu sin (nt) + nv cos (nt) , t).
The treatment until until now has been exact, and is nothing more than the well known variation of
constants used to obtain solutions to inhomogeneous linear ordinary differential equations. One now

42
introduces the near-identity transformation

= u + w1 (u, v, t) + 2 v1 (u, v, t) + O 3 ,

u
(B.0.5)
= v + w2 (u, v, t) + 2 v2 (u, v, t) + O 3 ,

v

where w1 ,w2 and v1 , v2 are called generating functions [24], and are chosen such that the transformed
equations for u and v are as simple as possible. Substituting, the near-identity transformation, Eq. (B.0.5),
into the slow flow equations, Eq. (B.0.4), gives
   
w1 1 v1 1
u sin (nt) F1 (u, v, t) + 2 sin (nt) K1 (u, v, t) + O 3 ,

=
t n t n
    (B.0.6)
w2 1 v2 1
v + cos (nt) F1 (u, v, t) + 2 + cos (nt) K1 (u, v, t) + O 3 ,

=
t n t n

where K1 (u, v, t) = F1 (w1 , w2 , t) + F2 (u, v, t), using the fact that w1 ,w2 are themselves functions of u, v
and t.
Now the generating functions are chosen in order to simplify the equations, Eq. (B.0.6), as much as
possible. What is usually done, for O 1 say, is to trigonometrically reduce the equations and then


choose w1 and w2 to remove all the trigonometric terms in t [21, 24]. This effectively amounts to taking
the first term in a Fourier series of the O 1 terms, which was Lagranges approach [31]. The argument


for first order averaging is that the variables u and v change slowly over time, such that the contribution
of these changes, averaged over the period, is zero except for the first term in the Fourier series [31].
The result of first order averaging is a pair of equations of the form

1 RT
u = sin (nt) F1 (u, v, t) dt + O 2 ,

nT 0
(B.0.7)
1 RT
v cos (nt) F1 (u, v, t) dt + O 2 ,

= 0
nT

where u and v are kept constant over the integration.


Once w1 and w2 have been obtained, one can go to second order, where v1 and v2 are now chosen, again
in order to simplify the equations as much as possible.
Second order averaging results in a pair of equations of the form

u = G11 (u, v) + 2 G12 (u, v) + O 3 ,



(B.0.8)
v = G21 (u, v) + 2 G22 (u, v) + O 3 ,


that are known as the slow-flow equations.


The initial choice of not using the polar representation is due to it often leading to nonlinear slow-flow
equations for Matheiu type equations [31], which usually require a little more analysis to understand.
This transformation is however the most natural one for many application, and the one used by Tondl
et al to study autoparametric systems [29], as one is often interested in the change in amplitude of a
system.

43
References

[1] D.J. Acheson. A pendulum theorem. Proc. R. Soc. London Ser. A-Math. Phys. Sci., 443(1932):239
245, 1993.

[2] V.I. Arnold. Ordinary Differential Equations. Springer-Verlag, Berlin Heidelberg, 1992.

[3] G.L. Baker and J.A. Blackburn. The Pendulum, a case study in physics. Oxford University Press,
2006.

[4] H. Broer and M. Levi. Geometrical aspects of stability theory for Hills equations. Archive For
Rational Mechanics And Analysis, 131(3):225240, 1995.

[5] H. Broer and C. Simo. Resonance tongues in Hills equations: A geometric approach. Journal Of
Differential Equations, 166(2):290327, 2000.

[6] J.A. Burns. More on pumping a swing. American Journal of Physics, 38(7):9201, 1970.

[7] W.B. Case. The pumping of a swing from the standing position. American Journal of Physics,
64(3):215220, 1996.

[8] W.B. Case and M.A. Swanson. The pumping of a swing from the seated position. American Journal
of Physics, 58(5):4637, 1990.

[9] A. R. Champneys. The dynamics of parametric excitation. In R. Meyers, editor, Encyclopedia of


Complexity and Systems Science. Springer-Verlag, 2009.

[10] S.M. Curry. How children swing. American Journal of Physics, 44(10):9246, 1976.

[11] H. Goldstein, C.P. Poole, and J.L. Safko. Classical Mechanics. Addison Wesley, third edition, 2002.

[12] B.F. Gore. The childs swing. American Journal of Physics, 38(3):3789, 1970.

[13] B.F. Gore. Starting a swing from rest. American Journal of Physics, 39:347, 1970.

[14] W.P. Jordan and P. Smith. Nonlinear Ordinary Differential Equations. Oxford University Press,
fourth edition, 2007.

[15] J.E. Kulkarni. Time-optimal control of a swing. In 42nd IEEE Conference On Decision and Control,
Vols 1-6, Proceedings, pages 17291733. IEEE, 2003.

[16] L.D. Landau and E.M. Lifshitz. Mechanics. Pergamon Press Ltd., 1960.

[17] C.C. Lin and L.A. Segel. Mathematics applied to deterministic problems in the natural sciences.
MacMillan Publishing Co., Inc., 1974.

[18] W. Magnus and S. Winkler. Hills Equation. John Wiley and Sons, Inc., 1966.

44
[19] N.W. McLachlan. Theory and Application of Mathieu Functions. Oxford University Press, 1947.

[20] J.T. McMullan. On initiating motion in a swing. American Journal of Physics, 40(5):764766, 1972.

[21] L. Ng and R. Rand. Nonlinear effects on coexistence phenomenon in parametric excitation. Nonlinear
Dynamics, 31:7389, 2003.

[22] B. Piccoli and J.E. Kulkarni. Pumping a swing by standing and squatting. IEEE Control Systems
Magazine, 25(4):4856, 2005.

[23] A.A. Post, G. de Groot, A. Daffertshofer, and P.J. Beek. Pumping a playground swing. Motor
Control, 11(2):136, 2007.

[24] R. Rand. Lecture Notes on Nonlinear Vibrations. published online by The Internet-First University
Press (2004), http://dspace.library.cornell.edu/handle/1813/79, 2005. Version 52 available online at:
http://www.tam.cornell.edu/randdocs.

[25] G. Recktenwald. The Stability of Parametrically Excited Systems: Coexistence and Trigonometrifi-
cation. PhD thesis, Cornell University, 2006.

[26] J.R. Sanmartin. O botafumeriro - parametric pumping in the middle ages. American Journal of
Physics, 52(10):937945, 1984.

[27] A.E. Siegman. Comments on pumping on a swing. American Journal of Physics, 37(8):8434, 1969.

[28] P.L. Jr. Tea and H. Falk. Pumping on a swing. American Journal of Physics, 36(12):11656, 1968.

[29] A. Tondl, T. Ruijgrok, F. Verhulst, and R. Nabergoj. Autoparametric Resonance in Mechanical


Systems. Cambridge University Press, 2000.

[30] M. van Noort. The Parametrically Forced Pendulum. PhD thesis, Rijksuniversiteit Groningen, 2001.

[31] F. Verhulst. Nonlinear Differential Equations and Dynamical Systems. Springer-Verlag, Berlin
Heidelberg, 1996.

[32] F. Verhulst. Perturbation analysis of parametric resonance. In R. Meyers, editor, Encyclopedia of


Complexity and Systems Science. Springer-Verlag, 2009.

[33] J. Walker. Flying Circus of Physics with answers. John Wiley and Sons, Inc., 1977.

[34] J. Walker. How to get the playground swing going - a first lesson in the mechanics of rotation.
Scientific American, 260(3):106109, 1989.

[35] S. Wirkus, R. Rand, and A. Ruina. Modelling the pumping of a swing. Newsletter for the Consortium
for Ordinary Differential Equations Experiments (CODEE), (Winter-Spring):711, 1997.

[36] S. Wirkus, R. Rand, and A. Ruina. How to pump a swing. College Mathematics Journal, 29(4):266
275, 1998.

[37] R. Zounes and R. Rand. Transition curves for the quasi-periodic Mathieu equation. SIAM Journal
on Applied Mathematics, 58(4):10941115, 1998.

45

Anda mungkin juga menyukai