Model Diff
Model Diff
Model Diff
Modelling with
Differential and
Difference Equations
GLENN FULFORD;
TETER FORRESTER
ARTHUR, JONESP
Modelling with Differential
and Difference Equations
AUSTRALIAN MATHEMATICAL SOCIETY LECTURE SERIES
GLENN FULFORD
Department of Mathematics,
University College ADFA, Canberra
PETER FORRESTER
Department of Mathematics,
Melbourne University
ARTHUR JONES
Department of Mathematics,
Latrobe University
CAMBRIDGE
UNIVERSITY PRESS
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, SAo Paulo
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9780521440691
A catalogue record for this publication is available from the British Library
Preface
Introduction to the student
Part one: Simple Models in Mechanics
1 Newtonian mechanics
1.1 Mechanics before Newton
1.2 Kinematics and dynamics
1.3 Newton's laws
1.4 Gravity near the Earth
1.5 Units and dimensions
2 Kinematics on a line
2.1 Displacement and velocity
2.2 Acceleration
2.3 Derivatives as slopes
2.4 Differential equations and antiderivatives
3 Ropes and pulleys
3.1 Tension in the rope
3.2 Solving pulley problems
3.3 Further pulley systems
3.4 Symmetry
4 Friction
4.1 Coefficients of friction
4.2 Further applications
4.3 Why does the wheel work?
5 Differential equations: linearity and SHM
5.1 Guessing solutions
5.2 How many solutions?
5.3 Linearity
5.4 The SHM equation
v
vi Contents
ix
x Preface
the end of the year). Some chapters of the book assume a knowledge
of linear equations, complex numbers, vector algebra, or the elements of
probability theory. The mathematical prerequisites are listed at the start
of each chapter.
This book grew out of notes prepared for a first-year course given at
La Trobe University for each of the last six years. The total time allotted
to the course was 65 hours, including lectures and practice sessions. Not
all chapters were covered in the same year and some choice is possible.
A longer course could be organized by covering all the material in the
book and including some computer work where relevant.
Each year we refined the material and its presentation, based on
our experience in teaching the course during the previous years. Some
curious incidental difficulties were faced each year by some students.
These included (a) correct use of minus signs in setting up equations
(b) sketching diagrams to illustrate the choice of a particular coordinate
in a mechanics problem (c) distinguishing between the parameters of
a problem and its unknowns. We have attempted to address these and
other difficulties in both the text and the exercises. We have also analysed
the steps involved in solving various types of problem, at least in the
early part of the book, and we find this helps students to present their
solutions to exercises clearly.
We wish to thank Sid Morris and Ed Smith for assistance with the
overall planning of the original course and Alan Andrew, Jeff Brooks,
Peter Stacey and John Strantzen for improvements in certain sections.
One of the authors (G.F.) also wishes to thank Colin Pask for his
encouragement. We also thank Dorothy Berridge and Annabelle Lippiatt
for assisting with the typing.
1
2 Introduction to the student
the models when the problems are changed slightly. Carrying out these
modifications will help you to develop many of the skills which are used
in mathematical modelling.
Setting up a mathematical model of a problem involves the following
steps.
Identifying the quantities most relevant to the problem and then making
assumptions about the way in which the quantities are related. This
usually involves simplifying the original problem so as to emphasize the
features which are likely to be most important.
Introducing symbols to denote the various quantities, and then writing
the assumptions as mathematical equations.
Solving the equations and interpreting their solutions as statements about
the original problem.
Checking the results obtained to see whether they seem reasonable and,
if possible, whether they are in agreement with experimental data.
(a) (b)
Make simplifying Write these
assumptions about assumptions as
real-world problem mathematical
equations
(d) (c)
Do the Solve equations
predictions of the and interpret the
model agree with solutions in terms
experimental of the real-world
results? problem
YES
Use model to make further predictions,
and for control, in the real world
7
8 Newtonian mechanics
Aristotle claimed even more than is stated in the third of these assump-
tions, namely, the speed at which bodies fall is proportional to their weight.
Thus if a ball is twice as heavy as another one, it should reach the ground
in half the time (when released simultaneously from the same height).
Although these assumptions did not go unchallenged during the Mid-
dle Ages, the decisive break with Aristotle's ideas came during the
Renaissance with the work of GALILEO (1564-1642). Part of the folk-
lore surrounding Galileo concerns his dropping unequal weights from
the leaning tower of Pisa to disprove Aristotle's claim about the relative
speeds at which bodies fall. Less well known but equally deserving of
mention is an interesting `thought experiment' which Galileo suggested
to show that this claim of Aristotle led to absurdity. This is explained
later in the exercises.
To obtain further details about the motion of falling bodies, Galileo
rolled a brass ball down a wooden beam. This had the effect of slowing the
falling motion of the ball, making it possible to record, with reasonable
accuracy, the times at which the ball passed various points marked on the
beam. Whereas Aristotle's assumptions had referred to velocity, Galileo
expressed the result of this experiment in terms of acceleration (or rate of
1.1 Mechanics before Newton 9
increase of velocity). His results were consistent with the postulate that
bodies fall with uniform acceleration.
In place of Aristotle's claim that force was necessary to maintain a
body in motion, Galileo further postulated that
in the absence of forces a body need not be
at rest but can proceed with uniform speed.
A body moving in the absence of forces is said to be kept going by its
own inertia. Galileo's views on such inertial motion, however, were, by
later standards, rather timid. To account for the fact that objects do not
fly off the Earth into outer space, Galileo allowed the inertial motion to
take place along circles. He did point out, however, that motion along a
big enough circle would be indistinguishable from motion along a line.
Meanwhile Johannes KEPLER (1571-1630) had been investigating the
orbits of the planets. He believed he could explain the existence of six
planets (Neptune, Uranus and Pluto were then undiscovered) and the sizes
of their orbits from the geometry of the regular solids. Although none of
this is taken seriously today, Kepler devoted his life to elaborating these
ideas. Almost as a by-product he discovered the three laws of planetary
motion for which he is now famous.
1. The planetary orbits are ellipses with the sun as a focus.
2. The area swept out by a ray from the sun to the planet is propor-
tional to the time taken.
3. The squares of the periods of the planets are proportional to the
cubes of the major axes of their orbits.
Kepler's laws were based on the observations of the planets' motion by
the astronomer Tycho BRAHE (1546-1601). The basis for Kepler's laws
was thus empirical. There was at that time no mathematical model from
which these laws could have been deduced.
Anyone interested in learning more about the work of Aristotle, Galileo
or Kepler should consult the book by Cohen (1987), the relevant chapters
being 2, 5 and 6.
Exercises 1.1
1. Two balls have the same size but one is made of material which makes it five
times as heavy as the other. According to Aristotle's model, what is the ratio of
the speeds with which they fall?
2. Galileo refuted the idea that, in the absence of air-resistance, heavier bodies
fall faster than light ones by the following `thought experiment': Suppose it were
true that
10 Newtonian mechanics
(a) Of the two bodies shown below, which would fall faster, given the above
assumption?
2 kg
1 kg
(b) In view of (a), what should be the effect on the speed of the 2 kg body
if you placed the 1 kg body beneath it, as shown below?
2 kg
1 kg
(c) How does your answer to (b) contradict the original assumption?
Kinematics
Kinematics is concerned with the geometry of motion. What paths
do moving objects follow? What coordinate systems are best suited
to describing their paths? What are their velocities? What are their
accelerations? In other words, kinematics is concerned with the most
obvious aspects of motion the things we can see with our eyes.
Kepler's laws of planetary motion belong to kinematics since they give
a geometric description of the path traced out by a planet. Galileo's law
that bodies fall with uniform acceleration also belongs to kinematics.
The idea of a particle is basic for the kinematic model we shall be using
in this book. A particle is defined as a body which has zero dimensions
and hence may be regarded as occupying a single point. Clearly such a
body cannot exist in the real world, but the idea is none the less useful
in modelling, say, the solar system since the size of a planet is very small
compared with the interplanetary distances. On the other hand, once a
rocket ship gets close to a planet it is no longer appropriate to regard
the planet as a single point, although it may well be still appropriate to
1.2 Kinematics and dynamics 11
think of the rocket ship in this way. In this book the primary concern
is with situations in which it is appropriate to model the bodies by
particles.
The concepts of velocity and acceleration will be explained more fully
later in the book. For the present, however, it is sufficient to recall that
velocity measures how fast a body is moving, while acceleration measures
how quickly the velocity is changing. Velocity and acceleration must be
measured relative to a frame of reference. For example, Parliament House
is at rest relative to the Earth, but it is moving with a large velocity relative
to the sun. Hence its velocity depends on whether the Earth or the sun
is used as the frame of reference.
Strictly speaking, velocity and acceleration should be considered as
vector quantities since they have direction as well as magnitude. Initially,
however, we shall simplify matters by supposing that all motion takes
place along a line; hence the number of possible directions is reduced to
two (up and down, left and right, etc.).
Dynamics
Dynamics, in contrast to kinematics, is concerned with what causes bodies
to move in certain ways. It involves concepts like mass and force, which
are so basic that it is difficult to define them in terms of any simpler ideas.
Hence, instead of attempting to define these concepts in a general way,
we point to some of the everyday situations from which these concepts
have arisen.
As to the concept of force, it had already been mentioned in the
previous section that we exert a force when we push against objects, with
a view to moving them. Although we cannot see a force directly, we are
aware that a force is acting when we push against a table because we feel
the tension in our limbs. Galileo is generally regarded as the founder of
modern dynamics because he told what happens when there is no force
acting on the body: it moves with uniform velocity (whereas Aristotle
had claimed it would necessarily come to rest).
Like velocity and acceleration, force is a vector quantity, which has
direction as well as magnitude. For example, the result of kicking a
football depends not only on how big the kick is, but also the direction
in which the kick is aimed. It will be assumed, moreover, that if two
forces act at the same point of a body, then they will have exactly the
same effect as their sum, calculated in accordance with the rules of vector
algebra, as illustrated in Figure 1.2.1.
12 Newtonian mechanics
F2
+ve direction
Initially, all the action will be taking place along a fixed line and so
there will be only two directions to worry about. One of the directions
will be selected as the positive direction and the opposite direction will
be called the negative direction. Our convention for representing the
direction of a force acting along the line is illustrated in Figure 1.2.2.
On the left of the figure is shown a force of magnitude 3 units, acting in
the positive direction. On the right, however, the force has magnitude 2
units but acts in the negative direction. (Our convention is thus different
from the usual one, in which both arrows would be labelled with positive
numbers.)
The other dynamical concept, that of mass, arises from our everyday
experiences in which a given force does not always produce the same
effect. For example, kicking a football full of water produces a markedly
different result from kicking the same ball full of air. The motion is
smaller in the first case because of the increased mass of the ball.
Newton himself described mass as the `quantity of matter' in a body.
Implicit in this description is the idea that putting two identical bodies
together gives double the mass of each of the individual bodies. This
idea opens up the possibility of constructing a graded set of standard
masses against which the mass of any body (of moderate size!) can be
compared on a balance, as in Figure 1.2.3. (Strictly speaking it is their
weights which are being compared, but at a given point on the Earth's
surface, these are proportional to their masses.)
The above discussion is intended to provide a general idea of the
1.3 Newton's laws 13
AAA
Fig. 1.2.3. Scales are used to determine when masses are equal.
meaning of the concepts of force and mass, but it stops short of giving
precise definitions. The status of the concepts of mass and force in
mechanics is analogous to that of a point and a line in Euclidean
geometry. These concepts are so basic for geometry that it is not possible
to define them in terms of any ideas which are more basic. Instead,
there are the axioms or postulates of geometry, which tell us all we need
to know in order to make logical deductions about points and lines.
In mechanics, the role of the axioms is taken over by Newton's laws,
discussed in the next section.
Exercises 1.2
1. How many standard masses ('weights') do you need to be able to find the
mass, correct to the nearest kilogram, of any mass less than 10 kg?
applied to a body whose mass changes with time for example, a body
consisting of rocket plus fuel. We shall not use this form of Newton's
second law in this book. As to Newton's third law, a simple illustration is
provided by the propulsion of rocket ships. The rocket ship exerts a force
on the material which it expels backwards. This material exerts an equal
but opposite reaction on the rocket ship, which causes it to accelerate
forwards.
In later chapters, detailed models will be given for the forces which
act in various situations, such as those where friction or air-resistance
is present. There is, however, one type of force which acts, according
to Newton, on every body in the universe. Hence it deserves to be
mentioned here, alongside his laws of motion. Newton's law of universal
gravitation states:
Each pair of bodies in the universe exerts a force of mutual attraction
of magnitude
Gml m2
r2
where ml and m2 are the masses of the bodies, r is the distance
between them, and G is a constant (independent of the bodies).
On the basis of the above laws, Newton was able to derive Kepler's
laws of planetary motion one of the first big successes of Newtonian
mechanics.
A very readable account of Newton's work in celestial mechanics is
given in Koestler (1958), pages 504-517. A popular biography of Newton
is Andrade (1979). A critical discussion of Newton's own statement of
his laws of motion is contained in Westfall (1971), chapter 8.
Exercises 1.3
1. To each statement below attach the most appropriate name from the list:
Aristotle, Galileo, Newton.
1. A force is necessary to maintain a body in motion.
2. The negation of (a).
3. Bodies fall to earth with uniform acceleration.
4. Velocity is proportional to the force acting on the body.
5. Acceleration is proportional to the force acting on the body.
6. In the absence of forces, bodies move along straight lines.
7. In the absence of forces, bodies move with uniform speed but can move
around circles.
16 Newtonian mechanics
0 Mass m
Weight mg
GMm
r2
1.4 Gravity near the Earth 17
Fig. 1.4.2. Gravitational attraction of a solid sphere equals that of a single point
of the same mass.
where r is the distance of the particle from the centre of the Earth, as
shown in Figure 1.4.2.
It follows from this formula that, as the altitude of a particle increases,
the gravitational attraction of the Earth on the particle decreases. For
everyday purposes, however, the variation of gravity with height can be
ignored. The radius of the Earth is so large (about 6400 km) that changes
in height lead to very small proportional changes in the distance of the
particle from the centre of the Earth and hence to correspondingly small
changes in gravity. This is illustrated by the following example.
it had in Paris. Newton's explanation for this was that the Earth is not
a perfect sphere, but is flattened at the poles. The larger distance of a
particle from the centre of the Earth when the particle is on the equator
accounts for the smaller gravitational attraction there.
Some values of g obtained experimentally at various latitudes are given
in Cohen (1987), page 175. They range from about 9.78 metres/s2 at
the equator to about 9.83 metres/s2 at the poles. Unless stated otherwise
we shall ignore the variation in the value of g and use the value of 9.8
metres/s2. This will give a satisfactory model for the gravitational force
on a particle provided it stays near the surface of the Earth and the
duration of its motion is not too large.
Exercises 1.4
1. At what altitude is the weight of a given sample of material 1 % less than
when it is on the ground, directly beneath?
2. What is the percentage increase in the weight of a given sample of material
when it is taken from the equator to a pole?
There is also a rule for the mth power of a quantity, where m is a fraction :
20 Newtonian mechanics
force MLT-2
area L2
volume L3
Example 1. The formula for the period i of small oscillations of a simple pendulum
is
i =2n 0-1/9 (1)
where c is the length of the pendulum and g is the acceleration due to gravity.
Check that this formula is dimensionally correct.
Solution.
[LHS]=[i]=T
[RHS] = [2zc t /g] = [Cl g] 2 = (L/(Lr2)) = T .
Thus each side of the formula has the same dimensions, as required. Note that the
multiplicative numerical factor 2x in (1) is dimensionless.
Exercises 1.5
1. The kinetic energy of a particle of mass m which is moving with a velocity v
is defined to be
1
2 mv2.
What is the unit for kinetic energy when mass and velocity are measured in SI
units? What are the dimensions of kinetic energy?
2. Let F be a force, let m1 and m2 be masses and let ( be a length. Is the
following formula correct dimensionally?
ml
F= C.
m1 +m2
2
Kinematics on a line
for some twice differentiable function ¢. The discussion will then lead us
to define the velocity z of the particle at time t in terms of the derivative
of 0 by
The velocity and acceleration can then be calculated by using the rules
of differentiation.
Just as important for mechanics is the reverse problem : given the veloc-
ity or the acceleration as a function of the time, what is the displacement
at a given time? This problem will be solved by antidifferentiation and
will provide an introduction to the idea of a differential equation.
This chapter provides all that is needed on differential equations for
21
22 Kinematics on a line
Chapters 3 and 4, while Chapter 5 introduces some harder types of
differential equations, which will be used in Chapter 6.
I 4
30 km,
A B
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
1 Particle
0 x
Fig. 2.1.4. The coordinate x measures the displacement of the particle from the
origin.
Particle
Ground level
Fig. 2.1.5. Here the coordinate x measures the displacement of the particle above
ground level.
positive direction of the axis has been made already. The following
example illustrates how the origin of the coordinate axis and its positive
direction may be inferred from the description given in the problem.
for time; any instant prior to that is then labelled with a negative t value.
For example, if we choose today as day 0, then tomorrow is day 1 and
yesterday was day -1.
Some books use the same symbol x for the displacement and for the
function 0. This can lead to confusing notation such as `x = x(t)'. We
try to avoid such usage in this book, even though it means we have to
introduce extra symbols.
With the aid of the above ideas on displacement, we can now complete
our discussion of velocity.
Non-uniform velocity
In non-uniform motion, the velocity may change from one instant to
another. Hence there is a need to analyze the idea of velocity at an
instant or instantaneous velocity. At first sight, however, there appears to
be something paradoxical about this idea. An instant has zero duration,
hence during an instant the particle cannot move. How then can it have
an instantaneous velocity?
We resolve this paradox by looking at smaller and smaller time intervals
surrounding the given instant and taking the limit of average velocities
over these intervals. Let the time change from t to t + 6t and let the
displacement change from x to x + 6x. At the start of the change
Hence by subtraction
5x = 4(t + 5t) - 4(t).
This gives the change in displacement as the time changes from t to t + 6 t
and hence the average velocity of the particle during this period is
5x 4)(t + 60 - 4i(t)
6t 6t
Since 0 was assumed differentiable, this ratio will approach a limit as
bt -- 0 which is just the derivative of 0 at t. Hence we may define the
instantaneous velocity of the particle at time t as
6x
lira
8t--+o 6 t
26 Kinematics on a line
Thus we have been led to define the instantaneous velocity (or simply
the velocity) at time t as the derivative at t of the displacement-time
function. The function 0' is called the velocity-time function.
Notation
The velocity at time t is written in Leibniz's notation as
dx
dt
This notation helps recall the way in which the velocity was defined above
as the limit of a ratio : a small increase in x divided by a small increase
in t. It also reminds us that velocity is a rate of change : of displacement
with respect to time. For these reasons it is understandably popular with
those who are primarily concerned with setting up mathematical models
of physical problems.
On the other hand, Leibniz's notation ignores the functions which
link the physical quantities. There are many situations in mathematics
where it is necessary to recognize these functions quite explicitly and then
the notation 4'(t) has the advantage. Familiarity with both notations,
together with the ability to translate from one to the other, is therefore
desirable.
An alternative notation for the velocity the one used by Newton in
fact is
x.
This means that, during the interval when the time increases from t to
t + St, the net change of displacement of the particle has been in the
positive direction of the coordinate x. We may summarize this by saying
that
when z > 0 the particle moves in the positive direction for the x-
coordinate.
Similarly it can be shown that
when z < 0 the particle moves in the negative direction for the x-
coordinate.
The particle is said to be stationary when z = 0 because the velocity is
then instantaneously zero.
The following example illustrates the ideas that have been introduced
in this section.
Example 2. A particle moves along a horizontal line in such a way that its dis-
placement x metres to the right of an origin 0 at time t seconds is given by
x = t3 +2t2 -9t+9.
At the instant when t = 1, find the displacement and velocity of the particle. State
also the side of the origin on which the particle lies, the direction in which it is
moving, and its speed at this instant.
Solution. Here the displacement-time function is defined by 4(t) = t3 + 2t2 - 9t + 9.
By differentiation, the velocity at time t is ac = 41(t) = 3t2 + 4t - 9. Thus, when
t = 1,
x=3 and x=-2.
Since the particle moves to the right if x increases, this is the positive direction for
the coordinate x. Hence, when t = 1, the particle is to the right of the origin but
is moving left. Its speed is 2 m/s.
Exercises 2.1
1. A particle drops over the edge of a table and falls vertically downwards. The
coordinate x is chosen as the distance through which the particle has fallen at an
arbitrary time t.
(a) Which of the four diagrams below best conveys the meaning of x?
(b) Which is the positive direction for the coordinate x? Why?
x
28 Kinematics on a line
2. Repeat Example 2 in the text, but this time suppose that x metres is the
displacement of the particle to the left of the origin at time t seconds.
3. Let 0 be the displacement-time function for Example 2 in the text. Write down
each of the numbers 0(0), 4)(2), 4(t + 1). Write down also 41(0), 41(2), 0`(t + 1).
4. Two particles Pl and P2 are connected by a string (of constant length). P1 lies
on the table while P2 hangs from the rightmost edge of the table, shown below.
Let x be the distance of Pl from the rightmost edge of the table and let y be the
distance of P2 below this edge at time t.
P1
2.2 Acceleration
In everyday usage, a car is said to be accelerating when its speed is
increasing and decelerating when its speed is decreasing. In mechanics,
however, the word acceleration is used in a slightly more technical way,
being defined in terms of rate of change of velocity rather than speed,
and it may be positive or negative.
To frame a definition of acceleration we use the notation of the
previous section and we suppose that as the time changes from t to t + 6 t
the velocity changes from ac to is + 65c so that
with
is = 4'(t)
2.2 Acceleration 29
This gives the change in velocity as the time changes from t to t + 6 t and
hence the ratio
6z '(t + 6t) - '(t)
gives the change in velocity per unit time. This ratio is called the average
acceleration during the time interval from t to t + 6 t. Since 0' is assumed
differentiable, the ratio on the right approaches the limit 4)"(t) as 5t -- 0.
Hence we are led to define the instantaneous acceleration (or simply the
acceleration) of the particle at time t as
6t
Thus the acceleration of the particle at time t is the first derivative of the
velocity-time function or the second derivative of the displacement-time
function at this instant.
Notation
The acceleration of the particle at time t is written in Leibniz's notation
as
d dx d2x
or, more briefly, as .
dt dt dt2
Thus there is some divergence between the everyday usage of the word
acceleration and its use in mechanics. Our usage is summarized by saying
that :
Example 1. Suppose that a particle moves along a horizontal line in such a way
that its displacement x metres to the right of an origin 0 at time t seconds is given
by x = 4(t) where
4(t)=t2-t-2.
(a) Sketch the graph of the displacement-time function. At which times is the
particle
(i) at the origin ?
(ii) to the right of the origin ?
(iii) to the left of the origin?
(b) At which times is the particle
(i) stationary?
(ii) moving to the right?
(iii) moving to the left?
(c) Which is the leftmost point reached by the particle?
(d) Find the acceleration. In which direction is the particle accelerating?
which in turn helps us to sketch the graph in Figure 2.2.2 - a parabola with
vertex at the bottom.
2.2 Acceleration 31
0 X +ve direction
x-axis
Note that in sketching the graph we draw the x-axis vertically upwards even
though in the original problem it was horizontal. From the graph it is clear
that the particle is
(i) at the origin when t = -1 or t = 2,
(ii) to the right of the origin when t < -1 or t > 2,
(iii) to the left of the origin when -1 < t < 2.
(b) Differentiation gives
ic=4'(t)=2t-1.
The particle is thus
(i) stationary when is = 0, hence t = 2
(ii) moving right when is > 0, hence t > z
(iii) moving left when xc < 0, hence t < 2
(c) The particle reaches its leftmost point when is = 0, hence t = 2 and so
x = -2 . The leftmost point is thus 2 metres to the left of 0.
(d) Differentiation
a of xc with respect to t gives
a
z=4'(t)=2.
Thus the acceleration is 2 m/s2. Since z > 0, the particle is accelerating
towards the right.
32 Kinematics on a line
t=2
t
t=1
. x-axis
-3 -2 0 1 2 3 4
Exercises 2.2
1. In Example 1 in the text, how would the solution change if x were defined to
be the displacement of the particle to the left of the origin at time t?
2. Suppose that a particle moves along a horizontal line in such a way that its
displacement to the right of an origin 0 at time t is given by x = 4(t) where
4(t) = t3 - 6t2 + 9t.
(a) Sketch the graph of the displacement-time function, showing where it
crosses the axes.
(b) Express ac and x as functions of time, and sketch their graphs.
(c) At which times is the particle
(i) at the origin?
(ii) to the right of the origin?
(iii) to the left of the origin?
(d) At which times is the particle
(i) stationary?
(ii) moving to the right?
(iii) moving to the left?
(e) At which times is the particle accelerating towards the right?
2.3 Derivatives as slopes 33
X-axis
Tangent
(f) Sketch a tracking diagram showing how the particle moves along the
line.
3. Repeat the above exercise, but with 4(t) = -t4 +2t2 -1. What is the rightmost
point reached by the particle?
t-axis
t-axis
1 2\.3 4 5 6 7
x = 0'(t)
-1
Example 1. Suppose that a particle moves along a line in such a way as to give
the graph in Figure 2.3.2 for the displacement x metres as a function of the time t
seconds. By examining slopes of tangents at various points obtain a rough sketch
of the velocity-time graph. Show also the general shape of the acceleration-time
graph.
Solution. The part of the graph for which t < 1 is a line segment of slope 2, while
the part of the graph for which t >_ 3 is a line segment of slope -1. In between,
as t increases from 1 to 3 the slope decreases steadily from 2 to -1. Hence the
velocity-time graph contains a horizontal line segment at height 2 for t < 1 and
another such segment of height -1 for t > 3. These are linked by a curved portion
which drops smoothly from height 2 to height -1 as in Figure 2.3.3.
The acceleration-time graph is now obtained by taking slopes on the velocity-
time graph. Along the constant parts of the velocity-time graph the slopes are zero.
Between t = 1 and t = 3 the slope drops from zero to some negative minimum value
and then comes back up again to zero. Hence the acceleration-time graph has the
general shape shown in Figure 2.3.4.
2.3 Derivatives as slopes 35
Exercises 2.3
1. A particle is moving along a horizontal line and its displacement to the right
of an origin 0 at time t is x. The displacement-time graph for the motion is
shown below.
= 0(t)
t-axis
What is the velocity when t = 1 ? In which direction is the particle then moving?
36 Kinematics on a line
2. Repeat the previous exercise, but this time suppose that x denotes the
displacement of the particle to the left of the origin 0 at time t.
3. The graph below shows the velocity-time graph for the motion of a particle
whose velocity and acceleration are both positive when t = 2.
.z=O'(t)
t-axis
2
=3
. x-axis
-4 -3 -2 -1 0 1 2 3 4 5 6
Time
Example 1. What can be said about the displacement x of a particle if its accel-
eration is given as a function of the time t by
.ic = t? (1)
4)'(t) = 1 t2 + ci
2
for some constant cl. A further antidifferentiation now gives
4)(t) = 61 t3 + cl t + c2 (2)
for some constant c2. But since the particle starts from rest, at the origin, the initial
values of x and ac are x = 0 and x = 0 at t=0. Hence 0(0) = 0 and 0'(0) = 0 from
which using (1) it may be deduced that ci = c2 = 0. Hence the required answer is
X= 1 t3.
6
Exercises 2.4
The exercises from number 4 onwards provide practice at applying differential equa-
tions to constant acceleration problems. In each of these exercises you must set up
a differential equation and then solve it - merely quoting a formula is not what
is wanted. You may assume the acceleration g due to gravity is 9.8 m/s2 vertically
downwards.
1. Use antidifferentiation to find the solution of the differential equation
x=t+1
which satisfies the initial condition x = 2 when t = 0.
2. Find the solution of the differential equation
x=e`+e
which satisfies the initial conditions x = 2 and is = 1 when t = 0.
3. A particle moves along a line and its displacement to one side of an origin
at time t is x. The acceleration is given as a function of time by the differential
equation
x = sin(2t).
Find x as a function of t, given that x = is = 1 when t = 0.
4. Read through the following problem, then answer the questions (a) to (e)
below :
2.4 Differential equations and antiderivatives 39
A stone was dropped from rest at the top of a cliff and a clunk
was heard 3 seconds later when it struck the ground at the foot of
the cliff. How high was the cliff`' With what speed did the stone
hit the ground ?
(a) There are two obvious points Pl and P2 either of which would be a
natural choice for origin. Which points are they?
(b) Show on a diagram how to choose a coordinate x for the particle, with
Pl as the origin for the x-axis.
(c) Write down a differential equation for x as a function of t and state the
initial conditions. Hence solve the original problem, stated above.
(d) Repeat parts (b) and (c) but this time choose the coordinate x so that
P2 is the origin for the x-axis.
(e) Which of the two choices of origin gives the neater solution?
41
42 Ropes and pulleys
+ ve direction
-T
B
Most people who have thought about it at all would probably answer
NO! The pull exerted by the team at one end of the rope will be the
same as the force felt by the team at the other end. A mathematical
model for the rope will be set up in which this can actually be proved
from Newton's laws of motion. First, however, it is necessary to analyse
the idea of tension in a rope.
Consider a length of rope which is being pulled at either end. Figure
3.1.1. shows a section through a rope at a point along its length. We
imagine this section as determining two separate pieces of rope, say A
and B.
The piece of rope A will have a force acting on it due to B of magnitude
say T > 0. Since the piece of rope B can pull, but cannot push (it would
go slack), the direction of this force on A must be in the direction shown
in Figure 3.1.1. By Newton's third law, the piece of rope B will have
a force acting on it due to A of the same magnitude T > 0, but with
opposite direction.
We define the tension in the rope at the point of section to be the
common magnitude T > 0 of the above forces.
A rope is said to be light if it has zero mass. Although such ropes
cannot occur in practice, we assume their existence as part of our idealized
mathematical model. They should be closely approximated in practice by
ropes whose masses are small compared with the masses of the objects
at either end. We also assume Newton's laws are applicable even when
3.1 Tension in the rope 43
-Ti T2
+ ve direction
the mass is zero. The assumption of a light rope gives the following
proposition, which is the key to modelling many practical problems.
Proposition 1 For a light rope, stretched taut by forces at either end, the
tension stays constant along the length of the rope.
Proof Let Ti and T2 be the tensions at any two points along the rope,
so that the forces acting on the piece of rope between these points are as
in Figure 3.1.2.
The net force acting on the rope is T2 - T1 and so, by Newton's second
law,
T2 - Tl = {mass of rope} x {acceleration}
= 0 x {acceleration}
=0.
Thus
T2 = T1.
This shows the tension is the same at the two points. As these two
points were chosen arbitrarily, the tension remains constant along the
entire length of the rope.
The tension in a rope may change as it passes around a pulley. For
example, if the axle of a pulley is not properly lubricated, a large part
of the force exerted by the worker in Figure 3.1.3 might be expended in
getting the pulley to turn, instead of in raising the load.
A similar result might ensue if the pulley were large and massive a
lot of the force would be wasted in getting the pulley to spin.
The contrary case, in which there is little friction at the centre of the
pulley and the mass of the pulley is small compared with that of the
load, will be modelled by an idealized pulley in which both friction and
mass are zero. Such a pulley is said to be smooth and light. It is possible
to argue although we shall not go into detail here that such a
pulley does not change the tension in the rope.
44 Ropes and pulleys
The upshot of this section is therefore that, if a light rope passes around
one or more smooth light pulleys, the tension will be the same at either end
of the rope.
Exercises 3.1
1. By modifying the proof of Proposition 1 show that for a rope not necessarily
light, stretched taut by forces at either end, the tension stays constant along the
length of the rope provided its acceleration is zero.
2. A light rope has a particle of positive mass firmly attached at a point
somewhere between its two ends, as shown below. Are the two tensions at either
end of the rope always equal when the rope is stretched taut? Give reasons for
your answer.
-Ti T2
While some students may wish to refer to the above list as a guide to
solving the problem, others may use it merely as a checklist at the end
to ensure nothing essential has been omitted from their solutions. The
following example illustrates how these steps are carried out in detail.
Solution.
STEP 1: Draw a diagram and set up notation. The particles are shown in typical
positions in Figure 3.2.1 at time t.
Let x1 and x2 be the coordinates of the particles as in the diagram. Thus
x1 = distance of first particle below centre of pulley
and similarly for x2.
Now note that, for either particle, if its coordinate increases then it moves down-
wards. Hence downwards is the positive direction for each coordinate.
STEP 2: Relate the coordinates. Note from the diagram that
T T
x2
xl
2 kg
l kg
-T
l kg
2 kg
lg
2g
1
But, since this does not change with time, xl + x2 must be constant and hence its
derivative with respect to t is 0. By the rule for differentiating a sum, applied twice
in succession, it therefore follows that
X1 + x2 = 0. (1)
STEP 3: Show forces on each particle. Since the rope can only pull, the forces it
exerts on the particles must be in the upwards direction, which is negative for each
coordinate. In addition, each particle has its weight acting on it in the downwards
direction. Hence the forces on each particle are as in Figure 3.2.2.
STEP 4: Apply Newton's second law. For each particle, the mass times the accel-
eration equals the net force, which can be read off Figure 3.2.2. Hence
1X1 = lg - T (2)
2X2 = 2g - T (3)
STEP 5: Solve the equations. The equations (1), (2) and (3) form a system of
three simultaneous linear equations in three unknowns X1, X2 and T. The standard
method for solving such systems is the Gaussian elimination algorithm. To apply
this procedure, first bring all the unknowns to the left-hand side of each equation
3.2 Solving pulley problems 47
to get
X1 +X2 =0
X1 +T=g
2x2+T = 2g
Next, the coefficients of the unknowns are read off and placed in a matrix alongside
the column of right-hand sides to give
1 1 0 0
1 0 1 g
0 2 1 2g
The aim now is to reduce the coefficient matrix to the unit matrix by using opera-
tions on the rows. This gives the following matrices.
1 1 0 0
0 -1 1 g new row 2 row 2 - row 1
0 2 1 2g
1 1 0 0
0 1 -1 -g new row 2 = - row 2
0 0 3 4g new row 3 = row 3 + 2 x row 2
1 1 0 0
0 1 -1 -g
ag
0 0 3
new row 3 = 3 x row 3
1 1 0 0
0 1 0 lg new row 2 = row 2 + row 3
ig
0 0 1
3
The final matrix is called the row echelon form of the original matrix. Reinser-
tion of the unknowns in the row echelon form makes the solutions obvious:
zl = __g (4a)
3
X2 = 31 g (4b)
T = 4g (4c)
3
Thus the first particle has an acceleration vertically upwards of magnitude g m/s2,
the second particle has a downwards acceleration of the same magnitude,3 and the
tension in the rope is g newton.
3
STEP 6: Check the answers. Certain features of the answers (4) could have been
predicted from Figure 3.2.1. For example zl should be negative and z2 should be
positive because the lighter particle will accelerate upwards and the heavier one
downwards. The magnitude of the accelerations should be less than g, moreover,
48 Ropes and pulleys
because the fall of the heavier particle is impeded by the upwards pull of the
rope.
The answer for the tension should (and does) lie between g and 2g since the
tension must be larger than the weight of the lighter particle (to accelerate it
upwards) and less than the weight of the heavier particle (to allow it to accelerate
downwards).
Example 2. Suppose that in Example 1 the particles are released from rest when
the heavier particle is at a height of 2 metres above the floor. Find how long it takes
to reach the floor (given that the rope is long enough for this to happen before the
lighter particle hits the pulley).
Solution. By Example 1, the distance x2 of the heavier particle below the centre of
the pulley, when considered as a function of time, satisfies the differential equation
1
x2= 3g.
Suppose that initially the particle is a distance a metres below the centre of the
pulley. The initial conditions may be written as
x2 = a and 5C2 = 0 when t =.O.
The problem is to find t such that x2 = a + 2, since the particle has to drop a
further 2 metres.
Antidifferentiation applied twice to the differential equation and use of the initial
conditions gives the solution
x2 = 61 g t2 +a.
Hence x2 = a + 2 when 2 = 6gt2 and hence when t = 12/g, as t > 0. Thus the
particle hits the ground after about 1.1 seconds.
Exercises 3.2
ED
3 [Tg]
2 kg
(a) Show the coordinates yl and y2 of the particles on the above diagram.
(b) Suppose now that the system is set in motion. What does your physical
intuition tell you about
(i) the sign of yl,
(ii) the sign of y2,
(iii) the range in which the tension in the rope must lie.
(c) Suppose the particles are initially released from rest when the second
particle is at a height of 2 metres above the floor. What are the values
of y2 and y2 when t = 0?
2. Repeat the solution of Example 1 in the text, but this time choose the
coordinates xl and x2 to be the heights above the ground of the respective
particles, at time t.
How have the answers changed with the new choice of coordinates? Does the
first particle still accelerate upwards?
3. Two particles, of mass 2 kg and 3 kg respectively, are attached to the ends
of a light rope which passes over a smooth light pulley, which is suspended at a
fixed distance below the ceiling. Find the accelerations of the particles and the
tension in the rope, by following the steps explained in the text.
4. Suppose that in the preceding exercise the particles are released from rest
when the heavier particle is at a height of 1 metre above the floor. Find how
long it takes to reach the floor (given that the rope is sufficiently long for this to
occur before the lighter particle reaches the pulley).
x2
xl
Example 1. Two particles of mass ml > 0 and m2 > 0 respectively are attached to
the ends of a light rope which passes over a smooth light pulley suspended a fixed
distance below the ceiling. Find the accelerations of the particles and the tension in
the rope.
Discussion STEPS 1-5: These may be followed much as in Example 1 of
Section 3.2. As always the first step is to draw a diagram to indicate which
coordinates are to be used, as in Figure 3.3.1.
The remaining steps then lead to the following answers for the accelerations
and the tension :
ml -m2
xl = ml
+ m2 9 (1a)
m2 -ml
x2= ml+m2g
2ml m2
T = (lc)
MI + m2
STEP 6: Because the answers (1) contain parameters, it is possible to use them
to make a wide range of predictions about the behaviour of the system in Figure
3.3.1 and to perform a variety of checks on the answers.
The simplest of the checks is just to substitute ml = 1 and m2 = 2 into the
answers (1) and observe that these are then the same as the answers (4) found
for Example 1 of Section 3.2. Also, it is easily verified that zl + z2 = 0 which
3.3 Further pulley systems 51
must be true since the length of the rope stays fixed. Another check is that T > 0
which must hold by the definition of tension.
Further checks are as follows.
(a) Dimensional checks. Recall that [g] = LT2. The answer (1a) gives for the
dimensions of xl
[xl] _ Ml -M2 g = MM-1 [g] = LT-2
Emi + m2
which are the correct dimensions for acceleration. The answer (1 a) gives
2m1m2 MLT-2
[T] = m1+m2 g = M2M-1 [g] =
which are the correct dimensions for a force.
(b) Equilibrium cases. If the two particles have equal masses, then their weights
should exactly balance. Hence the particles should stay at rest or move
with uniform velocity. The answers predict this will happen, since putting
m1 = m2 in (1 a) and (i b) gives
xl =x2 =0
while (1c) shows the tension is then the common weight of the particles.
(c) Limiting values of the parameters. If we keep the first mass fixed and allow
the other to approach zero, we expect to obtain answers appropriate to
free fall of the first particle. This is what happens since
ml -M2 g
x1 = m1+ m2 by (1a)
m1-0
as m2--+0
while
2m1m2
T= g by (I c)
m1 + m2
0
--+ g=0 as m2 -+ 0-
m1
(d) Further inequality checks. In the case m1 < m2, it is shown in one of the
exercises how to derive from the answers the following inequalities :
0< X2 <g (2)
These inequalities say that the answers must lie within certain ranges
which are very plausible on physical grounds. Thus the inequalities (2)
say that the heavier particle must accelerate downwards and that the
magnitude of the acceleration must be less than that for free fall. The
inequalities (3), on the other hand, say that the tension in the rope must
be larger than the weight of the lighter particle (to make it accelerate
upwards) and less than the weight of the heavier particle (to allow it to
accelerate downwards). A similar discussion applies in the case where
mt > m2.
52 Ropes and pulleys
The solution to the above example was based on assumptions about
the lightness of the rope and the pulley, and the smoothness of the
pulley. Examples encountered in practice only approximate our idealized
model, without satisfying our assumptions exactly. Hence it is desirable
to consider ways to test the accuracy with which our model approximates
the real world.
To this end, recall that the answers for the accelerations provide very
simple differential equations which can be solved to give the heights of
the particles as functions of the time. By adjusting the relative masses
of the particles, we could achieve small accelerations and hence, with the
aid of a stop-watch, get empirical plots of height against time. These
results could then be compared with those predicted by our model.
Alternatively, assuming the validity of the model, we could use the
measurements to determine the acceleration g due to gravity. When put
to this use, the mechanical system shown in Figure 3.3.1 is called Atwood's
machine.
The next example introduces some of the complexities that arise when
more than one pulley is involved.
xl
x2
m1
Pulleys find practical application in the use of `block and tackle' to lift
heavy loads. A small force exerted by a worker can thereby be converted
into a large force acting on the load. Further details are given in Exercises
7 and 8 below.
Exercises 3.3
1. Complete the solution of Example 1 in the text by carrying out Steps 2-5
(explained in Section 3.2).
To solve the linear equations you may use the fact that the matrix below on
the left has the row echelon form shown on the right, where it = (ml + m2)-lg.
m1 0 1 mlg 1 0 0 (m1 - m2)jz
0 m2 1 m2g 0 1 0 -(m1 - m2)jz
1 1 0 0 0 0 1 2m1m2µ
54 Ropes and pulleys
2. (a) Show that the answers obtained for zl and z2 in Example 1 in the text
can be written so that they involve the masses m1 and m2 only in the
combination m1 /m2.
(b) If the masses are both doubled, what happens to the accelerations?
(c) How would you choose the mass ratio ml /m2 to make z1 small and
positive?
4. Suppose that in Example 1 the particles start from rest with the second
particle at a height of 1 metre above ground level. If the second particle takes
1 second to reach the ground, what is the ratio of the two masses? Assume, of
course, that the first particle does not run out of rope.
[You are to solve this problem by solving a suitable differential equation with
the relevant initial conditions.]
5. Two particles, of mass m1 and m2 respectively, are connected by a light rope
passing over a smooth light pulley. A third particle, of mass m3, hangs by a light
rope from the second particle. The coordinates of the first two particles at time t
are their respective distances x1 and x2 below the centre of the pulley.
(a) Copy the diagram above and show the coordinates xl and x2 on it.
(b) Use your physical intuition to find, in each case, a necessary and sufficient
condition on m1, m2 and m3 to ensure that throughout the motion
(i) zl > 0 (ii) z1 = 0.
(c) Over which sections of the rope must the tension stay constant? How
many different tensions are there?
(d) Which two particles have the same acceleration?
3.3 Further pulley systems 55
6. For the mechanical system in Exercise 5, find the accelerations of the particles
and the tension in the rope by carrying out Steps 1-6 (explained in Section 3.2).
To solve the linear equations you may assume that the matrix below on the
left has the row echelon form on the right, where µ = (ml + m2 + m3)-1g.
m1 0 1 0 mlg 1 0 0 0 (mi - m2 - m3)µ
0 m2 1 -1 m2g 0 1 0 0 -(m1 - m2 - m3)µ
0 m3 0 1 m3g 0 0 1 0 2m1(m2 + m3)µ
1 1 0 0 0 0 0 0 1 2m1 m3µ
9. Complete the solution of Example 2 in the text by carrying out Steps 1-6
from Section 3.2.
To solve the linear equations you may assume that the matrix below on the
left has the row echelon form shown on the right, where µ = (ml + g. 4m2)-1
-
2 2
m1 ml m2
(a) z=
1
2m2
,+ 4m2 g (b) xi = g
m1 MI +M2
11. A light rope passes over two smooth light pulleys suspended at a fixed height
below the ceiling. Attached to the ends of the rope are two particles of mass m1
and m3 respectively. The central portion of the rope passes under a pulley which
supports a particle of mass m2 in such a way that the sections of the rope not
touching the pulleys hang vertically, as in the diagram below.
Choose coordinates for each particle so that their positive directions are
downwards and then find the accelerations of the particles and the tension in the
rope following Steps 1-6, explained in Section 3.2.
To solve the linear equations you may assume that the matrix below on the
left has the row echelon form on the right, where µ = (mlm2 + m2m3 + 4m3m1)-1.
12. Find a necessary and sufficient condition on the masses ml, m2, m3 in
Exercise 11 in order that the particles remain stationary if initially at rest.
13. Three particles of mass ml, m2 and m3 respectively are firmly attached to a
light rope - one at either end and the remaining one at an intermediate point
along the rope, as shown below. The portion of the rope joining particles one
and two passes over a smooth light pulley attached to the ceiling; that joining
particles two and three passes over a similar pulley.
3.4 Symmetry 57
Find the tension in the rope and the accelerations of the particles by following
Steps 1-6 of Section 3.2. [Note that this problem differs substantially from
Exercise 11. It involves five equations in five unknowns.]
To solve the linear equations you may assume that the matrix
ml 0 0 1 0 mig
0 m2 0 1 1 m2g
0 0 m3 0 1 m3g
1 1 0 0 0 0
0 1 1 0 0 0
has the row echelon form
1 0 0 0 0 (mi - m2 + m3)µ
0 1 0 0 0 -(MI - m2 + MOP
0 0 1 0 0 (m1 - m2 + m3)µ
0 0 0 1 0 2ml m2µ
0 0 0 0 1 2m2m3µ
14. (a) Refer to Example 1 in the text, and assume ml < m2. Show that
inequalities (2) and (3) hold. To show (3) observe first that T can be
written in each of the forms
ml +M1
T = m2 + m2 ml g T= m2$.
MI +M2 MI +M2
3.4 Symmetry
In everyday life, symmetry is seen in the patterns on wall paper, in
the design of furniture and in the architecture of great buildings. In
nature, symmetry is most evident in the crystalline structure of solids like
common salt and snowflakes. Some of the most interesting problems in
58 Ropes and pulleys
//
x2 x2
xl xl
formula valid; the formula for zi changes into the formula for z2 and
vice versa.
Recognizing symmetry in a mechanical system thus leads to an addi-
tional check on the answers. Thus, for example, the formula
2mim2
T = ml + 2rn2 $
cannot be the correct answer to the above problem since interchange of
m1 and m2 transforms the RHS into
2m 1 m2
9.
M2+ 2ml
This is not equal to what we had before the interchange and so the
required symmetry is lacking.
In more advanced courses, symmetry can occur in mechanical systems
in quite complicated ways. The systematic study of symmetry was closely
associated with the rise of modern algebra, particularly group theory.
Exercises 3.4
1. Is the mechanical system of Example 2 of Section 3.3 symmetric with respect
to the first and second particles? Do you expect the formula for T to stay the
same when ml and m2 are interchanged? Verify your answer by carrying out this
interchange in the formula for T.
2. Is the mechanical system of Exercise 3.3.11 symmetric with respect to the first
and third particles? Do you expect the formula for T to stay the same when ml
and m3 are interchanged? Verify your answer by carrying out this interchange in
the formula for T.
4
Friction
Together with gravity, friction forces are the ones which play the biggest
role in shaping everyday life. Without friction, we would be unable to
drive our cars, to walk, or even to hold our pens. The simple laws
of friction on which our model is based seem to have been first stated
by Leonardo da Vinci (1452-1519), who wrote prolifically about lots of
things.
Although these laws for friction are very simple, they provide useful
estimates and qualitative predictions for a wide range of behaviour as-
sociated with friction. More sophisticated models are sometimes used,
however, in specialized areas (such as the design of bearings in engineer-
ing).
60
4.1 Coefficients of friction 61
F 77"7 P
Block
+ve direction
Plane
The forces acting on the block which act parallel to the plane are shown
in Figure 4.1.1, where right has been chosen as the positive direction.
The pushing force is denoted by P (where P < 0), while the friction
force which opposes it is denoted by F (where F > 0). In the diagram it
is assumed that the friction force acts in the positive direction while the
pushing force acts in the opposite direction. While the block is at rest
(or moving with uniform velocity), Newton's second law gives P + F = 0
and hence F = -P.
Thus, as I P I increases from zero, IF I increases by an equal amount until
it reaches a maximum value, at which stage the block begins to slide in
the direction of the push. The maximum magnitude of the friction force
will be denoted by
Finax
The magnitude of the friction force depends on how hard one pushes the
block and can assume any value between 0 and Fmax.
The value of Fmax depends on how hard the two surfaces are pressed
together. It is harder to make a heavy load slide than a lighter one. To
measure the extent to which two surfaces are pressed together, we note
that, by Newton's third law, the block and the plane exert equal and
opposite forces on each other in the direction normal (perpendicular) to
the plane. These forces are illustrated in Figure 4.1.2. Their common
magnitude N is called the normal reaction between the plane and the
block.
In the model we adopt for friction, it is assumed that the maximum
friction is a linear function of the normal reaction.
Law of static friction. For a given pair of substances (for the surfaces in
contact) there is a constant µS > 0 such that
Fmax = bus N.
Note that Fmax and N are both positive quantities since they are defined as
magnitudes.
The dimensionless constant us is called the coefficient of static friction
62 Friction
+ve direction
/Z 77
-N
Velocity: 4 Velocity: 0
F>0 4 -F
A
N
+ve direction
-mg
Fig. 4.1.4. On a horizontal surface the normal reaction force is opposite to gravity.
This table makes it clear why you should never brake your car so hard
that it begins to slide: the friction force is reduced to about 70% of its
value prior to sliding.
Applying the laws of friction involves calculating the normal reaction
between the block and the surface. In the special case in which the surface
is horizontal, this is particularly easy since both the normal reaction and
gravity act vertically, as in Figure 4.1.4. The net force on the block in
the vertically upwards direction is N - mg where m is the mass of the
block. The surface being fixed, the vertical acceleration of the block is
zero. Hence N - mg = 0 and so N = mg.
64 Friction
Exercises 4.1
2. A car of mass 2600 kg is parked across a driveway with the brakes on and
the owner has lost his keys. What is the magnitude of the force required to
(a) just start the car sliding?
(b) keep it sliding, once in motion?
Use the coefficients of friction given in the text.
3. Discuss the following statement (in which u denotes the coefficient of static
friction).
Values for µ depend on the materials in contact and the state of the surfaces,
and range from about 0.04 for ski wax on dry snow, through 0.4 for brake lining
on cast iron and 1 for rubber on a hard dry road, to values considerably greater
than 1 for very wide drag-racing tyres.
Velocity: -
X
Li_jF<O
T
0 +ve direction
1 v2
2 jig
which is the required distance.
66 Friction
x +ve x-direction
1
+ve y-direction
As a check on these answers note that both the time and distance
taken for the block to stop increase with the initial speed v, but decrease
as the friction /1 increases. The dimensions of the answers are correct
also.
2 kg
1_2g
Solution of the simultaneous linear equations (1), (2) and (3) for z, y and T gives
x = y = -0.7g, T = 0.6g.
Thus the block on the table has an acceleration of 0.7g m/s2 towards the pulley
and the tension in the string is 0.6g newton.
Case (b) : the block on the table is projected towards the pulley. The direction of
the friction force F is now reversed. Thus F acts in the positive x-direction and so
F = 0.1g. This leads to the answers
x = y = -0.63g, T = 0.73g.
Thus the block on the table has an acceleration of 0.63g m/s2 towards the pulley
and the tension in the string is 0.73g newton.
The above answers are valid while the blocks continue to move in the
original directions of projection. They will cease to be valid when the
velocity of the block on the table becomes zero or when the suspended
mass hits the ground.
Exercises 4.2
1. Repeat the solution to Example 1 in the text, but this time choose the positive
direction for the x-coordinate to be opposite to the direction of the velocity.
2. Verify that the answers obtained in Example 1 in the text, for the time and
distance needed by the block to come to rest, have the correct dimensions.
3. Suppose that in Example 2 in the text, the suspended block is given an initial
downwards velocity of 1 m/s. By solving a suitable differential equation find how
long it takes for the block to descend i metre.
4. As a generalization of Example 2 in the text, suppose that the block on the
table has mass ml > 0, the suspended block has mass m2 > 0
and the coefficient of friction is µ > 0. Show that
(a) when the block on the table is projected away from the pulley its
acceleration is
x= -µm1 - m2 g,
MI +M2
68 Friction
(c) In which case does the acceleration have the larger magnitude? Is this
reasonable?
F=O
the points of the wheel in contact with the road are stationary,
_,uNcFc1sN
(where js is the coefficient of static friction between the wheel and the
road, and N is the normal reaction). Thus, the model for stationary
friction permits the friction force to act in either direction and does not
exclude the possibility of its being zero. Some possibilities are illustrated
in Figure 4.3.2.
By way of contrast, note that to explain the friction forces on a sledge
being dragged along a road we would need to use the model for kinetic
friction, because slipping occurs between the road and the sledge. This
70 Friction
model permits only one value for the magnitude of the friction force,
namely AN, and so precludes it from having the value zero.
Exercises 4.3
1. Discuss the following statement.
The frictional force that opposes one body rolling over another is much less than
that for a sliding motion and this, indeed, is the advantage of the wheel over
the sledge. This reduced friction is due in large part to the fact that, in rolling,
the microscopic welds are `peeled' apart rather than `sheared' apart as in sliding
friction. This may reduce the frictional force by as much as 1000-fold.
[This quote is taken from Halliday and Resnick (1974), pages 80 and 81.]
5
Differential equations : linearity and SHM
71
72 Differential equations: linearity and SHM
x = 4(t). You can't antidifferentiate a function if you don't even know
which function it is!
Thus we are forced to look for a different way of solving differential
equations of the form (2). A crude, but none the less effective, way is to
guess a solution, then substitute it back into the differential equation to see
if it works. Even though your first guess may be not completely correct,
it will often be easy to see how to modify it to give a correct solution.
Constant solutions are particularly easy to guess.
Example 2. Find two solutions, other than the constant solution, of the differential
equation.
X = -X.
5.1 Guessing solutions 73
Solution. The second derivative is to be the negative of the function, suggesting sin
and cos. Choose x = sin(t), to get
LHS = .z = - sin(t)
RHS = -x = - sin(t).
Thus x = sin(t) is a solution. That x = cos(t) is a solution may be checked similarly.
Along with sin and cos, the exponential function figures prominently
in the solution of a lot of commonly occurring differential equations.
Successive differentiations give the following table
x=et x= e-t
ac=et x= -e-t
x =e .z= e-t
In each of the above cases, two differentiations give back the original
function. This enables us to guess the answer to the following problem.
Such results greatly extend the number of differential equations for which
we can guess non-constant solutions.
74 Differential equations: linearity and SHM
Exercises 5.1
In each case guess the required solutions and then substitute back into the differ-
ential equation. Work out both LHS and RHS and check whether they are equal.
1. In each case find all the constant solutions of the differential equation.
(a) x = x(x + 1) (b) X + x2 = 1
is of second order.
has the solution x = sin(t). From the theorem it therefore follows that,
for each c,
x = sin(t + c)
is also a solution. (You may and should also verify this by direct
substitution in the differential equation.) Thus the theorem has enabled
us to get infinitely many solutions from one solution -- not a bad trick !
As yet, however, we have not obtained all the solutions.
5.3 Linearity 77
Exercises 5.2
1. In each case give the order of the differential equation and say whether the
differential equation is autonomous.
(a) x=tx3 (b) X= X4
5.3 Linearity
A differential equation of the second order is said to be linear if it can
be written in the form
The function h(t) is then referred to as the right-hand side of the differ-
ential equation. A linear differential equation is said to be homogeneous
if its right-hand side is the zero function so that it can be written
X - f (t)ac - g(t)x = 0. (2)
Differential Homogenized
equation Order Linear Homogeneous version
N +2.z+x = 1 2 Yes No :z+2ac+x = 0
X = --x 2 Yes Yes
--t 2 Yes No z=0
z=x2+t 2 No
z=t2+x 2 Yes No :z=x
.z + cos(t)x = 0 1 Yes Yes
.z + cos(x)t = 0 1 No
Solution. This differential equation is linear and homogeneous. From Section 5.1, it
has the solutions x = cos(t) and x = sin(t). Hence, by the homogeneous superposi-
tion theorem for each choice of the constants cl and c2,
x = cl cos(t) + c2 sin(t)
is a solution. Since the differential equation is also second order, and cos and sin
are linearly independent, it follows that this formula gives all the solutions.
x = 4)P(t) + 4)H(t)
is a solution of the original linear differential equation. Each solution of
the differential equation can be obtained in this way, moreover, by suitable
choice of OH.
For a second-order linear differential equation, it follows that, if 01
and 02 is a linearly independent pair of solutions of the homogenized
equation, then every solution of the original equation can be written as
which gives
LHS=z+x=0+1 = 1
RHS=1.
Thus the guess is correct and (3) gives a particular solution.
Next, the homogenized linear differential is z + x = 0, which was solved in
Example 1 of Section 5.3, each solution being given by
x = cl cos(t) + c2 sin(t) (4)
for suitable cl and c2 E R.
Finally, the sum of the particular solution (3) and the solution (4) of the ho-
mogenized equation gives the solution
x = 1 + cl cos(t) + c2 sin(t).
In view of the inhomogeneous superposition theorem, each solution is given by this
formula for a suitable choice of cl and c2.
Exercises 5.3
1. Copy and complete the following table giving the classification of the
differential equations for x as a function of t.
Differential Homogenized
equation Order Linear Homogeneous version
:z+.z+x=et
x+7x+tx=0
x = sin(2t)
x = sin(2x)
2. Find all the solutions of the differential equation z + x = 2. Hence find the
solution which satisfies the initial conditions x = 1 and .z = 1 when t = 0.
5.4 The SHM equation 81
3. For each of the following differential equations (i) guess a linearly independent
pair of solutions, as in Section 5.1 and then (ii) find all solutions:
(a) z+ x= 0 (b) z- x= 0
(c) z=-4x (d) z=4x
(e) z=-9x (f) x+16x=0.
4. Find all the solutions of each of the following differential equations:
(a) z+x = 5 (b) :z - x = 1
(c) z=-4x+2 (d) x=4x+1
(e) x = -9x + 3 (f) x + 16x = 32.
where w > 0. Although (1) is called the SHM (simple harmonic motion)
equation, it is more accurately an equation involving a parameter and
represents infinitely many equations, one for each choice of the parameter
w. Examples which have arisen earlier in this chapter are the differential
equations
z = -x, -4x and z = -9x,
which correspond to the cases w = 1, w = 2 and w = 3, respectively.
With the aid of the results given in Section 5.1, it is easy to guess the
following pair of solutions to (1):
x = 41(t) = sin(wt) and x = 02(t) = cos(wt). (2)
There are now two ways in which to obtain the remaining solutions using
properties discussed previously.
for each choice of cl, c2 ER. As the pair of solutions (2) is linearly
independent and the SHM equation is second order, all its solutions are
given by (3).
82 Differential equations: linearity and SHM
The second way uses the fact that the SHM equation is also au-
tonomous. Given an arbitrary constant A >_ 0, we choose cl = A and
c2 = 0 in (3) to get the solution
x = A sin(cot).
x = A sin(w(t + E))
or, alternatively,
x = A sin(cot + S) (4)
I
Fig. 5.4.1. Graph of a solution of the SHM equation of amplitude A and period
2n/w.
Example 1. Find the amplitude and the phase of the solution of the differential
equation
5c = 2A cos(2t + 6)
Squaring these equations and adding them and then using cos2(b) + sin2(b) = 1
shows that A can only be ,,,[2-. Hence
x = Jsin(2t+i/4)
which clearly satisfies the required initial conditions. Hence the amplitude of the
desired solution is 2- and its phase is 7t/4.
84 Differential equations: linearity and SHM
Exercises 5.4
1. State which of the following differential equations are of the SHM type and,
if they are, give the period and frequency of their oscillatory solutions:
(a) z + 4x = 0 (b) z + 45c = 0
(c) x = -9t (d) x = -9x
(e) x=x (f) x-9x=0
(g) x + m x = 0 where k and m are positive constants.
2. In each case, find the solution of the differential equation
x=--9X
which satisfies the initial conditions :
(a) x=Oand is=3 when t=0
(b) x = 1 and .z=0 when t=0
(c) x = 1 and.z=3 when t =0
(d) x= 1 and .z = -3 when t = 0.
This chapter is based on a simple model for the force in a spring which
was first proposed by Robert Hooke, a contemporary of Newton. In this
model, the force exerted by a spring is assumed to be directly proportional
to the distance by which the spring is extended. It is quite a useful model
when the extension of the spring is not too large.
Some interesting mechanical systems arise when particles are attached
to the ends of springs. A consequence of Hooke's law is that the equations
of motion for such particles are linear differential equations, usually the
SHM equation or some simple variant of it. The solutions of these
differential equations can be expressed in terms of trigonometric functions
and hence the model predicts oscillatory motion for the particles.
Later in the book, when you have learned more about differential
equations, you will be able to include the effect of damping forces in the
model.
Oscillatory phenomena occur widely in nature: the alternate rising and
setting of the sun, the waxing and waning of the moon, the ebb and
flow of the tides are examples from physics, while the regular beating
of your heart is an equally familiar example from biology. Although
these phenomena may all be described by differential equations, these
differential equations turn out to be non-linear and hence much harder
to solve than the simple linear ones used in this chapter.
85
86 Springs and oscillations
OOOOOOOOOOOOOOOOOOOOOc
Force on hands:
used to pull wire doors shut are of this type. On the other hand, springs
used in the suspension of a car permit both expansion to longer than,
and compression to shorter than, their natural lengths. It is with springs
of this latter type that we shall be mainly concerned.
Hooke's law
If you attach one end of a spring to a fixed object, say the wall, and pull
on the other end so as to stretch the spring beyond its natural length,
the spring will exert a force on your hand which tends to pull it back
towards the wall as in Figure 6.1.1.
The further you stretch the spring, the larger this force becomes. If
you pull too far, however, the spring may become permanently stretched
and thereby be reduced to just a twisted piece of wire. In the problems
in this book, it will be assumed that the spring is never stretched to this
extent.
If, on the other hand, you push the spring back towards the wall so as
to compress it to less than its natural length, the spring will exert a force
on your hand which tends to push it back from the wall as in Figure
6.1.2.
The further you compress the spring, the larger this force becomes.
Eventually the spring becomes so compressed that each coil of the spring
touches the next one; after this, no further compression is possible. In
the problems in this book it will be assumed that springs are never
compressed as far as this.
Thus, whether extended or compressed, the force exerted by the spring
on your hand acts in a direction which tends to restore the spring to its
natural length. For this reason, the force exerted by the spring is often
6.1 Force in a spring 87
0000000000000000000000
Force on hands:
called a restoring force. For extensions and compressions which are not
too large, the magnitude of the force is given (to a reasonable degree of
accuracy) by the law first stated by Robert Hooke (1638-1703).
Hooke's law. The magnitude of the restoring force in a spring is directly
proportional to the length by which the spring is extended or compressed.
By introducing a constant of proportionality k > 0 we may write
Hooke's law as an equality:
The constant k is called the stiffness of the spring. Its dimensions are
those of force per unit length so that
MLT-2
[k] = -2
= MT
L
In the SI system the unit of k are units are newtons/metre or kg/s2. The
stiffness of a spring depends on the composition of the steel of which it
is made, the processes used in its manufacture, the thickness of the wire,
the number of coils in the spring, and so on.
Hooke established his claim to the discovery of his law by publishing
a famous anagram in 1676 consisting of the letters
c e i i i n o s s s t t u v,
which are a rearrangement of the letters in the Latin phrase
ut tensio, sic vis
which means: as the extension, so the force. It was not till two years later
that he revealed the meaning of the anagram to his colleagues.
Hooke's law provides the theoretical basis for the spring balance, which
is commonly used to measure weights and other forces. The fact that
88 Springs and oscillations
Exercises 6.1
1. How would you test the validity of Hooke's law for a particular spring and
how would you determine its stiffness?
Example 1. A light spring has stiffness k > 0. One end of the spring is attached
to a wall and the other end is attached to a particle of mass m > 0. The particle
and the spring lie on the floor, assumed smooth, and are free to move in a line
perpendicular to the wall. Find the equation of motion of the particle when the
coordinate for the particle is the extension of the spring.
+ ve direction
-----------------------------/+x ----------------------------- No.
OOOOOOOOOOOOOOOOOOOOOc
------------------- / -------------------- 4 --------- x ---------lo-
0m
F<0
+ ve direction
------------ /+X --------------- 0.
aaaaaaaaaaaaaaaaaaaaa
------------------- I ---------------------
1
Sm
F>0
Example 2. Describe the possible types of motion for the particle in Example 1.
92 Springs and oscillations
x-axis
Solution. Let the coordinate x for the particle be the extension of the spring beyond
its natural length t, at time t. Hence, by the solution to Example 1, the possible
motions are obtained by solving the differential equation
for x as a function of t. This is the SHM equation, discussed in Section 5.4, with
parameter co = k/m. The solutions are given by
x=Asin V k
t+b
m
where A, 6 E R are arbitrary constants with A >_ 0. The values of these constants
are determined from initial conditions.
It follows that the particle either remains at the equilibrium point where x = 0
(when A = 0) or oscillates about the equilibrium point with SHM of amplitude A
(when A > 0) as shown in the tracking diagram in Figure 6.2.4. It is assumed that
A is small enough to lie in the interval for which Hooke's law is valid. Changing E
corresponds to changing the origin of time and in the diagram we have put E = 0.
The period of the oscillations is 2n m/k. It is interesting to note that this period
is independent of the amplitude A of the oscillations.
As a check on our answer for the period, note that its dimensions are correct,
being given by
Our answer also shows that the period increases with the ratio m/k, which seems
physically plausible: particles with larger mass would take longer to complete an
oscillation, for a given stiffness.
6.2 A basic example 93
------------------- / -------------------- 0.
OOOOOOOOOOOOOOOOOOOOOa
-----------------------------/+x -----------------------------
x-axis
x = A x=0 x=A
t=z
t0
t-z
C
Fig. 6.2.4. Tracking diagram where the period i = 2n m/k.
Exercises 6.2
1. A particle moves on a line and the force F acting on the particle is given as
a function of its displacement x to the right of an origin 0 by the formula
F=kx, k>0.
(a) Which is the equilibrium point for the particle?
(b) In each of the cases x > 0 and x < 0 sketch a diagram showing the
particle in a typical position, the direction of the force and the direction
in which the particle would move if started from rest.
(c) Is the force restoring (in the sense that it always pushes the particle back
towards the equilibrium point)?
2. In each of the following cases, repeat Exercise 1 but use the new law of force :
(a) F = -kx2
(b) F = -kx3
(c) F = k(1 - x) (distinguish now the cases x > 1 and x < 1).
4. Repeat Example 1, but this time choose as coordinate for the particle its
distance y from the wall and so obtain the equation of motion
k k
y-}- y = mt.
m
[Hints: Follow the steps given at the start of the section, making whatever
changes are necessary in the notation. Be sure to get the correct formula for the
extension of the spring in terms of y and the natural length.]
5. How does the equation of motion obtained in Exercise 4 differ from that
obtained in Example 1 ? Solve the equation of motion in Exercise 4 and show
that the resulting description of the possible types of motion for the particle
agrees with that found in Example 2.
+ve
direction
m
Example 1. A light spring has stiffness k > 0. One end of the spring is attached
to the ceiling, and to the other end, hanging vertically below it, there is attached a
particle of mass m. Find the equation of motion of the particle when its distance
below the ceiling is taken as coordinate.
Solution.
STEP 1: Let t be the natural length of the spring, let y be the distance of the par-
ticle below the ceiling at time t. Since the particle moves downwards if y increases,
downwards is the positive direction for the y-coordinate. Let FS be the force on the
particle due to the spring, measured in the downwards direction.
STEP 2: Here we get a formula for the spring force FS by considering the possible
cases for y.
Case (a) : y > 1, spring extended. This case is illustrated in Figure 6.3.1.
The spring force acts upwards, opposite to the positive direction for the y-coordinate.
Hence FS < 0. The extension of the spring is y - t > 0. Hence, by Hooke's law,
the magnitude of FS is
{stiffness} x ( {extension} I = k(y - t)
which is positive. Since FS is negative, this implies
FS = -k(y - t) when y > t. (1)
Case (b) : y < e, spring compressed. This case is illustrated in Figure 6.3.2.
The spring force now acts downwards, in the positive direction for the y-coordinate.
96 Springs and oscillations
A
+ve
direction
y=-k(Y-t)+g
m
(4)
STEP 4: As a check, note that [k/m] = MT-2M-1 = T-2 so that each term in the
equation of motion has the dimensions of acceleration.
As another check, consider the equilibrium point of the system. This will be at
some distance, say d, below the natural length of the spring where the spring force
is balanced by gravity. At this point,
y=e+d and mg = U.
Thus d = mg/k and
y=I+mg/k.
Thus it lies a distance 21 g below the natural length position as we found earlier.
This physical argumentkthus gives the same value y = t' + mg/k as that obtained
by putting y = 0 in (4).
6.3 Further spring problems 97
00000000000000a
Fig. 6.3.3. Spring with two particles.
Example 2. Describe the possible types of motion for the particle in Example 1.
Solution. The equation of motion (4) obtained in Example 1 for the y-coordinate
of the particle may be written as
k k
y + m-Y = -m1+g .
A particular solution of this second-order linear differential equation is the constant
solution y = I+ mg/k. The homogenized differential equation is the SHM equation
k
+ m y=0
with the solutions y = A sin (/t where A, 6 E R with A >_ 0. Hence the
solutions of the equation of motion are given by
y= +mg+Asin
k -
kt+b
m
Putting A = 0 gives the equilibrium solution. For A > 0 and sufficiently small we
get a solution in which the particle oscillates up and down with SHM about the
equilibrium point with amplitude A and period 2n m/k.
The steps used to solve Example 1 may be adapted to help you with
the solution of more complicated spring problems. Such problems may
involve two particles attached to opposite ends of the same spring, as in
Figure 6.3.3, or several springs attached to the same particle, as in Figure
6.3.4. We assume the springs slide on a smooth horizontal table. The
modifications needed to solve these problems are as follows.
For the system consisting of two particles attached to the same spring,
two coordinates are needed one for each particle. The extension of
the spring can then be expressed in terms of these coordinates and the
natural length of the spring.
98 Springs and oscillations
oaoooooooooa v
D00000000000 `W'OOOOOOa
(a) (b)
If, as we assume, the spring is light, the forces exerted on the particles
at opposite ends of the spring will act in opposite directions. Since two
coordinates are involved, the motion of the particles will be described by
a pair of simultaneous differential equations.
For systems with several springs attached to the same particle, there
may be an increase in the number of cases needed to obtain all possible
combinations of extensions and compressions for the two springs. To
reduce the number of cases, we shall assume certain relationships between
the natural lengths of the springs.
In Figure 6.3.4, if for system (a) the two springs have the same natural
length, then they will both be extended or both be compressed (or both
be in the equilibrium position at any given time). Thus there are still only
two cases to consider. To simplify system (b), however, we assume the
sum of the natural lengths is equal to the distance between the walls. If
one spring is extended, the other is then compressed. Hence, once again,
there are only two cases to consider.
Exercises 6.3
1. Repeat Example 1, but this time choose as coordinate for the particle its
distance x below the equilibrium position and so obtain the equation of motion
z+ !x=0.
m
[Hints: Again follow the steps given. When sketching diagrams showing the spring
and the forces on the particle in the various cases, recall that the equilibrium
point is a distance d below the natural length where d = mg/k. When the
coordinate of the particle is x, the particle is a distance x below this again. Hence
obtain the extension of the spring in terms of x and d.]
2. How does the equation of motion obtained in Exercise 1 differ from that
obtained in Example I? Solve the equation of motion in Exercise 1 and show
that the resulting description of the possible types of motion for the particle
agrees with that found in Example 2.
3. Two springs are attached to a wall at one end and to a single particle at the
6.3 Further spring problems 99
other, as shown below. Their natural lengths are 11 and e2 respectively, where
el < 12. The springs lie in a line perpendicular to the wall and the distance of
the particles from the wall is y.
---------------- y ----------------- .
(a) Find the extension (or compression) of the springs in each of the cases
(i) 12 <y
(ii) 11 < y < ?2
(iii) y < 11 -
(b) Indicate the directions of the forces F1 and F2 exerted by the springs on
the particle in each of the above cases. What further information would
you need, to be able to write down their magnitudes?
+ ve direction
(a) Show, giving all relevant steps, that the equation of motion of the particle
is
m
(b) Solve the equation of motion and hence find the possible types of motion
for the particle.
(c) If the pair of springs were to be replaced by a single spring having the
same net effect, what should be its natural length? Its stiffness?
I 000000000000000000 `1 m t-0000000000000a
+ ve direction
(a) Find the equation of motion of the particle, giving all relevant steps.
(b) A student gets the answer for (a) as
x+kl k2x=0.
m
What is obviously wrong with this answer? Where did he go wrong?
000000000000000000000
---x.---
---------------------------y ----------------------------- b.,
(a) Show, giving all relevant steps, that in these coordinates the equations
of motion of the particles are
k
x= m(Y-x-
k
Y=-m(Y - x
(b) To `uncouple' these simultaneous differential equations introduce new
coordinates u and v by putting
u=y+x
v=y-x.
Since x and y are functions of t, so are u and v so you may differentiate
with respect to t to find iu and v in terms of z and Y. Hence show from
(a) that u and v satisfy the `uncoupled' differential equations
u=0
v+ 2kv=2ki.
m m
6.3 Further spring problems 101
Our interest in this problem stems from the fact that it leads to the idea
of a difference equation.
105
106 Difference equations
Month I Month 2 Month 3 Month 4
Fig. 7.1.1. A mature pair of rabbits (shaded grey) each month produce a new
pair of rabbits (shaded white). The new rabbits mature after two months.
The way in which the rabbits breed for the first few months is illustrated
in Figure 7.1.1. The pairs of rabbits are shown month by month within
their enclosure. It is assumed that none of the rabbits die and, for ease of
identification, each pair of rabbits is shown in the same position within
the enclosure from month to month. The newly born rabbits are shown
directly under their parents with an arrow pointing to them.
At the top of the enclosure each month is the original pair of rabbits
the only pair during the first month. This pair is assumed to be new
born when placed in the enclosure and hence produces a new pair of
baby rabbits in the second and in each subsequent month. The first pair
of baby rabbits, born in the second month, do not produce any offspring
till the fourth month.
Example 1. What is the total number of pairs of rabbits, each month, up to and
including the fourth month?
Solution. It is convenient to introduce some notation. Put
Yk = {number of pairs of rabbits in the month k}
for each integer k >_ 1. Inspection of the enclosures in Figure 7.1.1 shows that
Yi=1, Y2=2, Y3=3, y4=5. (1)
and thereby solve the original problem. This would be quite laborious,
7.1 Introductory example 107
Example 2. Develop a formula relating the number of rabbits present this month
to the number of rabbits present in the previous month.
Solution. Note first that, since we neglect rabbits' deaths, their total number can
only be affected by births. Hence, for the pairs of rabbits,
number presentj = number present + number born (2)
this month last month this month
Since the rabbits take two months to become productive and then produce only one
pair per month, the last term on the RHS of (2) is given by
number born = number present (3)
this month two months ago
provided the current month is at least the third month. Hence, substituting (3) back
into the RHS of (2) gives
number present = number present + number present (4)
this month last month two months ago
This is the desired relationship between the numbers this month and those for the
previous two months.
To express (4) in terms of the notation introduced previously, let the current
month be the kth month (k >_ 3). Hence the last month was the (k - 1)th and the
one before that was the (k - 2)th. Thus (4) may be abbreviated to
Yk = Yk-1 + Yk-2 (5)
where k = 3,4,5,... and we shall refer to (5) as the Fibonacci equation (despite
the fact that it was first written down explicitly by Kepler).
Equation (5), which relates the numbers this month to those in the two
preceding months, is an example of a difference equation.
To see how the difference equation helps in the calculation of the
remaining numbers, take k = 5 in (5) to get
Y5 =Y4+Y3
Since y3 and y4 are already known from (1), this gives y5 = 5 + 3 = 8.
The next step is to take k = 6 in (5) to get
Y6 = Y5 + Y4.
6
5
4
3
2
1
Months
0 1 2 3 4 5
Fig. 7.1.2. Graph shows the total number of pairs of rabbits present each month.
Because the time is discrete the graph consists of discrete points.
Exercises 7.1
In each problem, yk denotes the number of pairs of rabbits present in the
enclosure in the kth month.
1. For the original Fibonacci rabbit problem, it was shown in the text that
y5=8andy6=13.
(a) Use a suitable choice of k in Fibonacci's equation, (5) in the text, to find
Y7-
(b) In a similar way find in succession the numbers y8, Y9, YJo, Y11, Y12 thereby
completing the solution of Fibonacci's problem.
7.2 Difference equations - basic ideas 109
5. Repeat Exercise 4, but this time suppose the time taken for the rabbits to
become productive is four months. In part (b) you will need t choose for yl, Y2,
Y3, Y4 appropriately.
A rule which defines the kth member of the sequence in terms of the
(k - 2)th member (and possibly also the (k -1)th member or the number
k itself) is called a second-order difference equation. A unique solution of
such a difference equation is determined once the values of both yi and
Y2 are specified. Difference equations of the third and higher orders may
be defined in a similar way.
Solution. Taking successively k = 3,4,5,6 in the difference equation gives the fol-
lowing equations
Y3 = Y2 + Y1
Y4 = Y3 + Y2
Y5 = Y4 + Y3
Y6 = Y5 + Y4
The initial conditions, together with these equations used successively, give
y1=1, y2=1, y3=2, y4=3, Y5=5, Y6=8.
7.2 Difference equations - basic ideas 111
The above process of repeatedly substituting old values back into the
difference equation to produce new ones is known as iteration. It is clear
that this process will eventually produce yk for any prescribed value of
k.
Example 3. Guess a formula for the solution of the first-order difference equation
Yk=1+Yk-1+2V1 +Yk-i (k=2,3,4,...) (1)
which satisfies the initial condition yl = 0.
Solution. To ensure our guess will be an informed one, we first calculate the numbers
Y1, Y2, Y3, Y4. From the difference equation and the initial condition it follows by
iteration that these four numbers are respectively 0, 3, 8, 15. These numbers look
very close to the perfect squares 1, 4, 9, 16. More precisely
y1=12-1
Y2 =22- 1
Y3=32-1
y4=42-1.
This leads to the guess
yk =k2- 1 (2)
for allk>_ 1.
The formula (2) remains a guess at this stage since it has only been
verified to hold for four of the infinitely many possible values of k. The
following example shows how to verify it is valid for all k > 1.
Example 4. In the previous example, verify that the formula (2) for yk gives the
correct solution to the difference equation (1).
112 Difference equations
y-axis
25 F
p
20
15
10
5
'0,010 ,
10
10 10 10
0 k-axis
1 2 3 4 5
Fig. 7.2.1. Graph of the solution (2) of the difference equation (1).
Solution. It will be shown that the formula for yk satisfies both the initial condition
and the difference equation.
Putting k = 1 in (2) gives yl = 0; hence the initial condition is satisfied. Next,
to check that the difference equation is satisfied, first replace k in (2) by k - 1 to
get
(this replacement being valid since (2) was assumed to hold for all k >- 1). Sub-
stitution of these formulae in the difference equation (1) now gives, for k >_ It
The graph of the solution (2) is sketched in Figure 7.2.1. Since the
solution is a sequence, its graph is a discrete set of points rather than
a continuous curve. Note the reduced scale on the vertical axis to
accommodate the points.
7.2 Difference equations basic ideas 113
Exercises 7.2
1. In each case state the order of the difference equation and the number of
initial conditions needed to determine a solution uniquely :
2. Suppose that yk = Yk-1 for all k >_ 2. What can be said about the sequence
Y1, Y21 Y3, ...?
3. Suppose that Yk = (yk-1)2 for k >_ 2. For which values of k does it follow that
Yk+1 = (Yk )2
and shall write xn for the typical members of this sequence where n is
an integer >_ 0. The sequence's being constant means that xn does not
change with n and hence that
xn+1 = xn (n = 0, 1, 2, ...).
xn+1 = xn
2
(n=0,1,2, ...). (1)
Solution. If the solution is constant, then all members of the sequence have the same
value, which we denote by s. Hence xn+1 = xn = s for all n >_ 0. The difference
equation (1) implies that
s=s2
and hence s = 0 or 1. Thus the only possible constant solutions of the difference
equation are the sequences
xo = 0, x1 = 0, x2 = 0, ... and XO = 1, x1 = 1, x2 = 1, ... .
Conversely, these two constant sequences clearly satisfy the difference equation (1).
7.3 Constant solutions and fixed points 115
Example 3. Find all the fixed points s of the function g with g(x) = x2.
Solution. The fixed points s are the solutions of s = g(s), that is
s=s2
Hence the fixed points of g are the numbers 0 and 1.
116 Difference equations
R R
x-axis
Example 4. Show that the cosine function has just one fixed point s which lies in
the interval 0 < s < n/2.
Solution. The graphs of y = cos(x) and y = x are shown in Figure 7.3.3. They
intersect in just one point, whose x-coordinate lies between 0 and n/2. Hence there
is only one fixed point s and it satisfies 0 < s < n/2.
7.3 Constant solutions and fixed points 117
Fig. 7.3.3. Graph showing the single fixed point of the cosine function.
Exercises 7.3
1. In each case find all steady-state solutions of the difference equation using
the method of Example 1 in the text.
(a) xn+l = 2xn (n = 0,1, 2, ...)
(b) xn+1 = 2xn + 1
(c) xn+1 = 2(x2 - 3).
2. How many constant solutions does the following difference equation have?
Xk = xk_1 (k = 2, 3, 4, ...)
5. Suppose that a function g: R -> R assumes only positive values. What follows
about the fixed points (if any) of g?
6. Let g be the function corresponding to the difference equation
xn+1 = (xn)2-3 (n = 0, 1, 2, ...).
(a) Write down a formula giving g(x).
(b) Find the fixed points of g.
(c) Hence find the steady-state solutions of the difference equation.
118 Difference equations
8. In each case give a formula for g(x) where g is the function corresponding to
the difference equation.
9. In each case sketch graphs of y = g(x) and y = x on the same axes. Hence
decide whether the function g has any fixed points.
(a) g(x) = x2 - 1 (d) g(x) = e-x
(b) g(x) = x2 + 1 (e) g(x) = ln(x) (x > 0)
(c) g(x) = ex (f) g(x) = tan(x) (-In n < x < fir).
x-axis
(a) (b)
y = 8(x)
STEP 2: From Q project across horizontally until the point R is
reached on the line y = x.
STEP 3: Choose the next point to be the x-coordinate of R.
We call this the next-point procedure. The reasons why it gives the
correct point for xn+1 are as follows:
lying on the x-axis and forming a solution of the difference equation (1).
The pattern of arrows arising from use of the next-point procedure will
typically form a cobweb (although it might sometimes be better called a
zig-zag) path.
120 Difference equations
Solution. In each case the first step is to write down the function g for which the
difference equation has the form
xn+1 = g(xn)
The graphs of y = g(x) and y = x are then plotted relative to the same pair of axes
and the initial value xo is plotted on the x-axis. Repeated application of the next-
point procedure then produces the cobwebs shown in Figure 7.4.2 and the values
for
x0, X19 x2, x3, ..
(which check against those obtained directly from the relevant difference equation
by iteration).
Case (c) : no. As n increases, xn oscillates from one side of the origin to
the other with ever-increasing amplitude. Hence we say that the sequence
does not converge to any limit.
Case (d) : yes. Although xn oscillates as n increases, the amplitude dies
away and approaches 0. We can thus make xn as close as we like to 0 by
making n sufficiently large. Hence we say that the sequence converges to
the limit 0.
In the preceding example, the functions corresponding to the difference
7.4 Iteration and cobweb diagrams 121
(b)
(a)
y-axis
x-axis
x-axis
(c) (d)
Fig. 7.4.2. Cobweb diagrams for Example 1. Cases (b) and (d) converge to a
fixed point whereas (a) and (c) do not.
equations are all linear, in the sense of having straight-line graphs. While
this makes the graphs easy to draw, it also makes the behaviour patterns
for the solutions rather special. Thus, for example, if the function is
linear, it cannot have exactly two fixed points. The following example
provides a contrast with the linearity of the previous example.
for each of the following initial conditions: (a) xo = 0.9 and (b) xo = 1.1.
Solution. The function g corresponding to the difference equation (2) is given by the
formula g(x) = x2. Hence the cobweb diagrams are as in Figure 7.4.3. The solution
in case (a) converges to the limit 0 whereas the solution in case (b) diverges to oo.
122 Difference equations
x-axis
x4 x3 x2 x1 xo
Fig. 7.4.3. Cobweb diagrams for Example 2 show different behaviour for different
initial conditions.
Example 3. Use a cobweb diagram to discuss the behaviour of the solution of the
difference equation
xn+1 = -xn (3)
x-axis
The above sequence has the interesting feature that every second
member has the same value, even though the sequence is not constant.
This property may be expressed by saying that the solution has period 2
or is a 2-cycle. More generally, we say that a sequence
xo, X1 , x2, .. .
The smallest such p is then called the period of the sequence and the
sequence is called a p-cycle. In particular, a sequence of period 1 is a
constant solution.
Exercises 7.4
1. Sketch graphs y = x and y = x + 1, for 0 <_ x <_ 5, relative to the same pair of
axes. Hence draw a cobweb diagram for the solution of the difference equation
xn+1 =xn+1 (n=0,1,2,...)
which satisfies the initial condition xo = 0. Show x0, x1,.. . , x5 on the x-axis.
2. In each of the cases shown below, sketch a cobweb diagram for the solution
of the difference equation
xn+1 = g(xn) (n = 0, 1, 2,...)
124 Difference equations
(a) (b)
3. For each of the difference equations in Example 1 in the text, use iteration
to find x1, x2, x3, given that xo = 1. Hence check the statements made in Figure
7.4.2.
4. Repeat Example 1 in the text, but this time suppose the initial condition is
xo = -1.
5. Repeat Example 2 in the text, but this time choose the initial conditions to
be xo = -0.9 in case (a) and xo = -1.1 in case (b).
(b) Find all the numbers a * 0 for which the initial condition xo = a
determines a constant solution.
(c) Deduce that all non-constant solutions are 2-cycles.
where g is one of the two functions shown below. By experimenting with cobweb
diagrams corresponding to various initial conditions find:
(a) a solution of period 2 when g is given by the graph (i),
(b) a solution of period 3 when g is given by the graph (ii).
(i) (ii)
8
Linear difference equations in finance and
economics
126
8.1 Linearity 127
8.1 Linearity
The difference equations arising in this chapter are mainly first-order
linear ones, which have the form
STEP 1: Write down and solve the homogenized equation for (3). The homoge-
nized equation is
xk = 2k xo (5)
where the initial value xo can be chosen arbitrarily. Substitution of (5) back in (4)
shows that (4) is satisfied.
STEP 2: Guess one particular solution of the original equation (3). The RHS
of (3) involves only constants (apart from yk ). This suggests trying a constant
solution, in which yk+l = yk. The original equation (3) is then equivalent to
yk = 2yk + 3
and hence to
yk = -3. (6)
STEP 3: Add the solutions (5) of the homogenized equation to the particular
solution (6) of the original equation -- this gives all the solutions of the original
equation. Adding the solutions gives the formula
yk = -3 + xo2k (7)
for the solutions of the original difference equation (3), in terms of the parameter
xo.
yk=-3+5x2k (k=0,1,2,...).
Finally, it is a good idea to check that this formula satisfies both the initial
condition (put k = 0) and the original difference equation (substitute back in
(3)).
A general formula for the solutions of (1), which includes the solution
found in the above example as a special case, is given by the following
proposition.
8.1 Linearity 129
y=ax+b y=x
Yk+l = aYk + b,
with a and b constant, has its solutions given by the formula
1 ak(vo-lba)+iba if a* 1
Yk =
l yo ya=i
where k = 0, 1, 2, ... and where the initial value yo can be chosen as
desired.
a <-1
Exercises 8.1
In the exercises which ask you to `classify' a difference equation, state the order of
the difference equation and whether it is linear, homogeneous, has constant coeffi-
cients.
1. In each case, classify the difference equation and state whether Proposition 1 in
the text is applicable. If so, use it to get the closed-form solution of the difference
equation satisfying the stated initial condition yo = 3. Assume k = 0, 1, 2,....
(a) Yk+l = 3yk + 10 (b) Yk+1 = 2(Yk )3
(C) Yk+1 = Yk + 10 (d) Yk+2 = 2Yk+1 + 4Yk
2. Classify each of the following difference equations and state whether Propo-
sition 1 in the text is applicable.
(a) Xk+2 = 2kXk (b) Nk+1 = 3Nk
(C) Zk+1 = cos(Zk) (d) Yk+2 = -6 Yk+1 + 7 Yk + 5
3. For the difference equation
Yk+1 = I.lyk (k =0,1,2,...)
(a) write down in closed form the solution satisfying the initial condition
Yo = 100,
(b) use your calculator to find the integer k such that
yk < 200 < yk+1,
(c) state the last integer k for which yk is less than 200,
(d) state the first integer k for which yk exceeds 200.
(e) To which of the cases illustrated in Figure 8.1.2 does the above difference
equation belong?
10. This exercise illustrates how the methods of this section may be extended to
second-order difference equations. Consider the Fibonacci equation
Yk+2 = Yk+1 + Yk (k = 1, 2, 3, ...),
which is second-order linear homogeneous with constant coefficients. The solution
satisfying the initial conditions yl = 1 and Y2 = 1 will be found in closed form.
(a) Guessing solutions. The solutions for first-order equations suggest trying
exponential solutions, of the form
y = ak (k = 12293,... )
where a is a constant. Show that this satisfies the Fibonacci equation
only if a satisfies the quadratic equation a2 - a -1 = 0.
8.2 Interest and loan repayment 133
[Note that this gives two solutions, yk = (al )k and Yk = (a2 )k, where al
and a2 are the roots of the quadratic equation.]
(b) Superposing solutions. Check that
Yk = C1(a,)k + C2(a2)k
(k = 1, 2, 3, ...).
V-' 2
Give reasons.
The amount which the lender finally gets back depends, among other
things, on whether the money was lent at simple interest or at compound
interest. The annual rate at which interest is earned is expressed as a
percentage, say p% (or p% per annum). For compound interest, however,
there is also an interest conversion period, which is usually a year or some
fraction a of a year. How the interest is calculated and how the amount
on deposit grows will now be explained for each type of loan.
134 Linear difference equations in finance and economics
Simple interest
For loans of this type, the interest earned in any year is obtained by
applying the rate to the initial principle and thus it stays the same
throughout the duration of the loan.
Thus, for an annual interest rate of p%, the interest earned during any
one year is p% of the principal, that is,
{interest} = p x {principal}
100
and hence
amount on amount on
deposit after deposit after + x {principal}.
k -4-1 years k years 100
Sk+l = Sk + p So (k = o, 1, 2, ...).
100
This is a first-order linear difference equation of the type studied in
Section 8.1 with a = 1 and b = (p/ 100)So. Hence, by Proposition 1 of
Section 8.1, the solution is given by the formula
Sk = 1 + kp So (1)
100
which is called the simple interest formula.
The significance of this formula is that it gives the amount Sk into which
the principal So grows when it earns simple interest for k years at an annual
rate of P%.
Compound interest
This type of interest is more usual for loans made over longer periods.
Interest is added to the principal at regular intervals, called conversion
periods, and the new amount (rather than the principal) is used for
calculating the interest for the next conversion period. The fraction
of a year occupied by the conversion period is denoted by a so that
conversion periods of 1 month, 1 quarter, 6 months and 1 year are given
respectively by a = 12, a = 4, a = 2 and a = 1. Instead of saying that the
conversion period is 1 month, for example, we may say that the interest
is compounded monthly.
For an interest rate of p% and conversion period equal to a fraction
8.2 Interest and loan repayment 135
a of a year, the interest earned for the period is ap% of the amount on
deposit at the start of the period, that is,
finterestj - 100
ap X amount on deposit at the
start of the conversion period
and hence
amount on amount on amount on
deposit deposit « deposit
= +px
after k + 1
conversion
periods
after k
conversion
periods
1 afterk
conversion
periods
To express this as a difference equation, for each k let Sk denote the
amount on deposit after k conversion periods, giving
Sk+1 = Sk + ap-Sk
_ (1 + 100)
'xp S k (k = 0, 1, 2,...).
Sk = Soy (2)
100
which is called the compound interest formula.
This formula gives the amount Sk into which the principal So grows when
it earns compound interest for k conversion periods, each equal to a fraction
a of a year, at an interest rate of p%.
The formulas (1) and (2) show that the amount Sk increases as a linear
function of k when the interest is simple, but as an exponential function
of k when the interest is compound. The cobweb diagrams in Figure 8.1.1
illustrate the difference between these rates of growth, the exponential
growth being much greater than the linear one, at least in the long term.
When applying these formulae, recall that in (2) the letter k stands for
the number of conversion periods (rather than the number of years, as
in (1)). Thus to find the amount on deposit after 10 years at compound
interest with a monthly conversion period, take k = 120 in (2) (rather
than k = 10, as in equation (1)).
Loan repayments
Arguments similar to those used above may be applied to the study of
loan repayments. The particular scheme considered here is the one nor-
136 Linear difference equations in finance and economics
mally used for the repayment of housing loans and is called amortization.
Repayments are made at regular intervals, and usually in equal amounts,
to reduce the principal (the amount borrowed) and to pay interest on the
amount still owing.
It is supposed that compound interest at p% is charged on the out-
standing debt, with conversion period equal to the same fraction a of
the year as the period between repayments. Between payments, the debt
increases because of the interest charged on the debt still outstanding
after the last payment. Hence
1 debt after debt after interest on
k + 1 payments} = lk payments + { this debt } - {payment}.
To write this as a difference equation, let the initial debt to be repaid be
Do, for each k let the outstanding debt after the kth payment be Dk, and
let the payment made after each conversion period be R. Hence
a
Dk+1 = Dk + R
= (1+)Dk-R.
This is a first-order linear difference equation of the type studied in
Section 8.1 with a = 1 + ap/ 100 and b = -R. Hence the solution is given
by Proposition 1 in Section 8.1 as
Dk = (1 + ap-
l k CDO.100R 1 + 100R
(3)
100/ ap ) ap
Solution. Let R be the required monthly payment. In the loan repayment formula
choose
a= 1
12
(since a month is 121 of a year)
p = 15 (the annual percentage rate)
Do = 10 000 (the initial debt)
k = 60 (since 5 years = 60 months)
and hence get
D60 = 21071.81 - 88.57R.
But, for the loan to be repaid after 60 months, D60 = 0. Hence R = 237.91 and so
the required monthly repayment is $237.91.
Difference equations are derived for interest and loan repayment prob-
lems in Goldberg (1958), pages 87-90. Technical details concerning a
wide variety of financial calculations are given in Ayres (1963).
Exercises 8.2
The simple and compound interest formulae (1) and (2) in the text can be used
wherever they are relevant.
1. Find p so that interest compounded annually at p% produces the same
amount at the end of a year as 12% compounded quarterly.
2. Suppose that compound interest is earned on a certain investment, with an
interest conversion period of six months. Calculate the interest rate p% necessary
to double the initial principal in four years.
3. (a) Tabulate the amount on deposit at the end of each year for five years,
for an initial principal of $1000 earning simple interest at an annual rate
of 12.5%.
(b) Calculate the number of years it would take for the principal to double
at this rate.
4. Repeat Exercise 3, but with a compound interest rate of 12.5% and an interest
conversion period of one year.
5. Repeat Exercise 3, but with a compound interest rate of 12.5% and an interest
conversion period of six months.
6. Two insurance companies offer investors insurance bonds earning compound
interest at identical rates of interest. One company takes 5% of the initial
principal as their up-front charge while the other company leaves the entire
principal invested but takes 5% of the final amount as its withdrawal fee.
Does either scheme yield a better return to the investor? Give reasons for your
answer. You may assume interest rates stay fixed.
138 Linear difference equations in finance and economics
7. (a) By making any necessary changes to the derivation given in the text of
the compound interest formula, set up a difference equation to describe
the following situation:
An initial sum of money So is deposited to earn compound interest at
a rate of p% and the conversion period is a fraction a of a year. At
the end of each conversion period a further sum So is deposited, to
give a total amount Sk at the end of the kth conversion period.
(b) Find in closed form the solution of the difference equation found in (a).
9. In the loan repayment formula in the text, find the value of the repayment R
which just keeps the debt at its initial level. What happens if R is less than this
value? greater than this value? Argue from the formula.
10. The average loan on an Australian home is often quoted to be $55 000.
Calculate the monthly repayment necessary to have the loan repaid after 25 years
if the interest rate is 7.5%. What is the total amount paid back on the loan?
11. Bernoulli's inequality states that, for each real number x >_ -1 and each
positive integer n, (1 + x)" >_ 1 + nx. Use this inequality to prove that, for the same
principal and at the same rate of interest, compound interest always produces an
amount at least as great as that for simple interest.
12. Show that (1 + x/n)" is an increasing function of n > 0 where x is a fixed
real number >_ 0. Hence show that, for a given principal and rate of interest,
decreasing the conversion period can only increase the amount on deposit after
a specified time.
13. Discuss how you would modify some of the models from this chapter to
take account of inflation.
prices for a crop one season, will plant a larger area with that crop in
anticipation of a similarly high price the following season. Thus supply
lags one season behind the price of that particular crop. This provides
the motivation for assumption (a) in our description below of the cobweb
model. The assumption (b) relates demand to price in the same season, as
would normally be expected. The assumption (c) is relevant to perishable
goods like fruit and vegetables: the price will adjust to clear the market
by making demand equal to supply.
The cobweb model. This model of supply and demand for a commodity
makes the following assumptions relative to consecutive time periods:
In the above model, the functions have been assumed to be linear merely
for the sake of simplicity. Some economists use models in which the
functions may be non-linear.
To illustrate these assumptions we shall use the following notation for
the various quantities which are involved. For k = 0, 1, 2,... let
Example 1. Suppose that the supply and demand of potatoes are related to the
price by the straight-line graphs shown in Figure 8.3.1. Verify that assumptions (a)
and (b) of the cobweb model are satisfied. Show that, if (c) is assumed also, then
the price satisfies the difference equation
1
Pk = -2Pk-t + 1 (k = 1,2,3,...).
140 Linear difference equations in finance and economics
Sk ( kg) Dk ( kg)
1000 1000
750
500
Solution. The assumption (a) is satisfied because the supply Sk is a linear function
of the price pk-1 and, from the graph, Sk increases when Pk-1 increases. Similarly,
assumption (b) is satisfied because Dk is a linear function of pk and Dk decreases
if Pk increases.
It is easy to write down explicit formulae for the linear functions implicit in the
graphs. In each case we read off the slope from the graph (`rise over run') and then
add a constant to the RHS of the formula to make the graph pass through one of
the end points. This gives
Sk = 500pk-1 + 500,
Dk = -1000pk + 1500.
The assumption (c) translates to Dk = Sk and hence from the last two equations
1
Pk = -2Pk-1 + 1 (k = 1,2,3,...).
This is the required difference equation for the price.
Exercises 8.3
1. Suppose that in Example 1 in the text the price of potatoes is initially 90
cents.
(a) Sketch a cobweb diagram and deduce how the price behaves in the long
term.
(b) Find the solution of the difference equation for the price in closed form
and use it to check your answer in part (a) concerning the long-term
behaviour of the price.
2. Repeat Example 1 in the text, but this time suppose the graphs of supply and
demand against price are as shown below. Derive the difference equation for the
price,
Pk = -2Pk-1 + 5/2 (k = 1, 2, 3, ...).
Sk (kg) Dk (kg)
1000 1000
750
500
Yk = {national income}
Ck = {consumer expenditure}
Ik = {induced private investment}
Gk = {government expenditure} .
where a and b are constants with a > 1 and b < 0. The details are left
till later as a useful exercise. The difference equation (1) is first-order
linear constant coefficient of the type studied in Section 8.1 and hence its
solutions are given by
b b
Yk= a ( 0 _ (2)
I ) 1-a
Because a > 1, the formula (2) shows that Yk grows exponentially with
k (at least if Yo > b(1 - a)), but it does not oscillate. Hence Model I
does not predict the regular fluctuations in the national income which
we know occur in practice. To obtain more realistic predictions, some
modifications to' the model are needed.
An obvious point at which to try to improve Model I is at the
interpretation (c) of the accelerator principle. Knowledge of the latest
increase in national income comes too late for investors to take advantage
of it. Hence it would be more realistic to introduce a time lag and to use
the increase from one period back. This leads to the following model for
the national income.
Model II consists of the statements (a), (b) and (d) from Model I together
with the modified statement (c') of the accelerator principle:
(c') Induced private investment in any period is directly proportional to
the increase in the national income for the previous period above
the national income for the period before that.
This model leads to a second-order difference equation for the national
income of the type
Yk + aYk-1 + bYk-2 = c (3)
144 Linear difference equations in finance and economics
where a, b and c are constants satisfying certain inequalities. An example
appears in the exercises which shows such an equation can have a solution
showing cyclic (periodic) behaviour. Hence, in Model II, our objection
to Model I has been removed.
Samuelson's models for the national income are described in Gandolfo
(1971), pages 63-73, Kenkel (1974), pages 241-259, and in Pfouts (1972),
pages 116-119. Kenkel gives some interesting comments as to how these
models may be made more realistic.
Exercises 8.4
1. (a) In Model I write each of the statements (a), (b), (c) and (d) in terms of
the symbols introduced in the text for the various quantities. (Use the
letters A and B for the constants of proportionality in statements (b)
and (c) respectively.)
(b) Deduce from your answers to part (a) of this exercise that the national
income satisfies the difference equation
(1 - A - B)Yk= -BYk-l +Go (k = 1,2,3,...).
3. (a) In Model II write each of the statements (a), (b), (c') and (d) in terms of
the symbols introduced in the text for the various quantities.
(b) From your answers to part (a) show that the national income satisfies a
difference equation of the form
Yk+aYk-i+bYk-2=c (k=2,3,4,...)
where a, b and c are constants, as claimed in the text.
(c) In the case a = -1, b = 1, c = 100 show that the solution of this difference
equation satisfying the initial conditions Yo = 100 and Yl = 101 exhibits
cyclic (periodic) behaviour.
4. Consider the model for the national income consisting of statements (a), (c)
and (d) in the text together with the following assumption:
8.4 National income: `acceleration models' 145
146
9.1 Linear models for population growth 147
Cell division
In nature, species typically compete with other species for food and are
themselves sometimes preyed upon. Thus the populations of different
species interact with each other. In the laboratory, however, a given
species can be studied in isolation. We shall therefore concentrate, at
first, on models for a single species. The first example we shall study is
a population of yeast cells which reproduce by dividing into two. The
rate at which yeast cells divide is governed by environmental factors such
as nutrient availability and temperature. In the laboratory these factors
may be controlled so that the rate of dividing is constant.
Example 1. (Cell Division). Suppose a single cell divides every minute. Assuming
that none of the cells die determine how many minutes it will take before there are
more than one million cells.
148 Non-linear difference equations and population growth
Solution. We measure time in integer multiples of one minute. Let Nk be the number
of cells after k minutes (where k is an integer). So Nk+i will be the number of cells
one minute later. If each cell divides into two there will be twice this number of
cells one minute later. Hence we may write down the difference equation
Nk+i = 2Nk with No = 1. (1)
This linear difference equation, which is of the type studied in Chapter 8, has the
solution
Nk=2 k (2)
which simplifies to
k 61n(10)
= 19.93. (4)
In(2)
Thus in 19 minutes less than a million cells are produced and in 20 minutes more
than a million cells are produced. So 20 minutes is the required time.
Nk+l -
current number number + number
of individuals - who die born (6)
or
Thus we interpret the population after the first season as 1313 individuals and after
the second season as 1746 individuals.
The difference equation (7) is of the linear type studied in the previous
chapter. Its closed form solution, given the initial number No, is
N k = (1 + a - f3)kNo, (k = 0, 1, 2, ...).
r=a-J3. (8)
5 10 15 20
k (Breeding seasons)
Fig. 9.1.1. Population growth as given by a linear difference equation with growth
rate r.
Exercises 9.1
1. How many cells will a single cell produce after 10 divisions?
2. Suppose that a single yeast cell divides every 2 minutes. Also suppose 75%
of yeast cells survive to divide in the next generation.
(a) What is the growth rate?
(b) How many yeast cells will there be after 3 hours?
3. A population of birds on an island has a constant per-capita birth rate a and
a constant per-capita death rate /3 (per individual per year). Also, a constant
number I of birds migrate to the island each year.
(a) What is the appropriate time period to be used here?
(b) Formulate a suitable difference equation.
(c) Obtain the closed-form solution given the initial number of individuals,
No.
4. In a certain type of female insect population all the adults die before the eggs
hatch. Each adult contributes a constant number of eggs b, of which a fraction
f survive and develop into adult females.
(a) Set up a difference equation for the population after the kth hatching.
(b) What is the minimum number of eggs that should be laid so that the
population does not become extinct, given the fraction which survive is
20%.
1500
1500 300
1000
1000 200
100
500
500
0 5 10 0 5 10 0 5 10
(a) (b) (c)
Nk
Fig. 9.2.2. Growth rate R(Nk), plotted against Nk, for the discrete logistic equa-
tion.
s (I - s ) = 0. (5)
Other models
There are many examples of populations raised in laboratories which
have growth rates greater than 3. Thus the fact that the logistic equation
produces negative population for r > 3 means that this model is not
suitable for such populations.
We must not forget that the discrete logistic equation is based on an
assumption of simplicity namely the straight-line form of R(Nk). In
practice the form of R(Nk) can be more complicated. There are many
choices that we can make for the form of R(Nk) based on experimental
evidence. One such model, which still has the advantage of being fairly
simple, is popular in the biological literature. In particular it has been
used for fish populations; for example, see Greenwell and Ng (1984).
This model uses an exponential curve, yielding a difference equation
Nkea(1-Nk/K),
Nk+1 = (7)
1. Use a calculator to find the population for the first five breeding seasons:
(a) using the discrete logistic equation with K = 1000, r = 0.5, No = 200;
(b) using equation (7) in the text with K = 1000, No = 200 and a = ln(1.5).
3. A model for insect populations (in which all adults are assumed to die before
next breeding) leads to the difference equation
2Nk
Nk 1 = 1 + aNk
Nk
1500 1 1500 1
r = 0.2 r = 0.8
Fig. 9.3.1. Discrete logistic equation for r = 0.2 and r = 0.8. The population
increases then levels out to an equilibrium value.
Nk Nk
1500 1500
r= 1.6 r= 1.9
0
sao 500-
fl 0
1
10 20 30 40 0 10 20 30 40
k (Time intervals) k (Time intervals)
Fig. 9.3.2. Discrete logistic equation with r = 1.6 and r = 1.9. Damped oscilla-
tions.
1500 1500 7
r = 2.1 r = 2.4
o g o w w g c Q Q w o w Q g c Q F 9 Q
1000 iooo
O y : N ;i ii i ji
vi
500 500 4
0 10 20 30 40 0 10 20 30 40
k (Time intervals) k (Time intervals)
Fig. 9.3.3. Discrete logistic equation with r = 2.1 and r = 2.4. Each case
demonstrates a 2-cycle.
... ..
.8 o
A.
.4 so
.. , . ... ,... .... 9.
.
1.
aoo ob
00 .0 Go
e d :i c ;: 4 b o o 's: c :i a i o
a
500
10 20 30 40
k (Time intervals)
Nk
1500 1500 r=3
r = 2.6
o p
QTY Q Q;
9
A
1000 I. 1000
a
a
soo b
ad soo 6
i
0 10 20 30 40 0 10 20 30 40
k (Time intervals) k (Time intervals)
Discussion
We have now completed our numerical investigation of the effect of
varying the parameter r since, when r > 3, Nk becomes negative and
the model ceases to apply. Varying the parameter K does not provide
any useful insight since this corresponds merely to a change of scale in
the vertical axis. Varying the initial population No does not normally
affect the long-term behaviour of the population except when the the
behaviour is chaotic. This is also investigated in the exercises.
Numerical experimentation has now become an accepted part of ap-
plied mathematics. The value of numerical experiment is illustrated by
the results of the computer experiments, for the discrete logistic equa-
tion model, in Figures 9.3.1-9.3.5 since they raise some very interesting
questions. For instance, we might ask how realistic the predictions of
our model are compared with the results shown in Figure 9.2.1 for beetle
populations. Also, what is the exact value of r for which the popu-
lation first executes oscillations (2-cycles) ? But most importantly, the
results lead us to look for underlying causes of the oscillations and their
biological significance.
The key to understanding why the population oscillates depends on
two factors. Firstly, the population is self-regulating through the pop-
ulation-dependent growth rate. (Note that we found, in the previous
section from studying the difference equation, that population decreases
in the next time interval when it is greater than the carrying capacity
and increases when it is less than the carrying capacity.) Secondly, the
regulating effect is felt in the next time interval but actually determined
by the population in the current time interval. In other words there is a
natural delay in the population responding to overcrowding. As a result
when the ideal growth rate r is sufficiently high the population responds
by overcorrecting itself which leads to oscillations and sometimes chaos.
Note that for continuously breeding populations (described by differential
equations) the growth rate responds instantaneously to the population
and oscillations do not normally occur, except through external seasonal
factors.
162 Non-linear difference equations and population growth
One major limitation of the discrete logistic model is that it predicts
negative populations when r > 3, which is clearly unrealistic since many
real populations have growth rates exceeding 3. Despite this limitation the
model is still useful since it predicts the qualitative range of behaviour
as seen in Figure 9.2.1 and it does illustrate the importance of the
growth rate. These considerations are certainly important to population
biologists.
For general discussions on chaos see the books by Gleick (1987) and
Stewart (1990). A more technical, but still very readable reference, is
the book by Devaney (1986). Also the book by Tuck and De Mestre
(1991), which discusses how to use the programing language BASIC to
do computer experiments in population dynamics and ecology, is pitched
at an introductory level.
Exercises 9.3
1. (Computer Simulation). Use a computer to iterate numerically the equation
Nkea(1-Nk/K)'
Nk+1 =
where K is the carrying capacity and a is a positive constant. You are given
K = 1000, No = 100 and you should use your own choice of values of a. Describe
the types of behaviour which appear for different values of a. In particular:
(a) When do 2-cycles first appear?
(b) When do 2-cycles become 4-cycles?
(c) Does the model exhibit chaos?
for the insect population Nk. Using a computer, sketch solutions for the following
values of the parameters:
A b
(a) r = 0.5,
(b) r = 2.5,
(c) r = 3.
Start with No = 100 and then try No = 101 and No = 99. What happens for
each value of r?
4. The discrete logistic equation can be put into another form. The following
shows how to do this.
(a) Write the discrete logistic equation
Nk+1 =Nk+rNk 1- Nk
K
in the form
Nk+1 = ANk 1- Nk
K1
Measles epidemics
Measles is a highly contagious disease, caused by a virus and spread
by effective contact between individuals. It tends to affect mainly chil-
dren. Epidemics of measles have been observed in Britain and the
United States roughly every two to three years. They occur more
frequently in developing countries. An important problem is to un-
derstand what factors affect the timing and severity of measles epi-
demics.
Let us now look at the duration of the disease for a single child. A
child who has not yet been exposed to measles is called a susceptible.
Immediately after the child first catches the disease there is a latent period
where the child is not contagious and does not exhibit any symptoms of
the disease. This is because the virus has not yet multiplied sufficiently.
The latent period lasts, on average, 5 to 7 days. After this the child
enters the contagious period. The child is now called an infective since it
is possible for another child who comes in contact with the infective to
catch the disease. This period lasts approximately one week. After this
time red spots appear on the skin of the child for a few days after which
the child recovers. During this period, and subsequently, the child is
immune to the disease and cannot be reinfected. Note that an individual
cannot become an infective until a week after they have been infected,
due to the latent period.
9.4 A coupled model of a measles epidemic 165
Example 1. Suppose there are 10 infectives and 1000 susceptibles present in the
kth week. Suppose that each infectives infects two susceptibles. How many infectives
and susceptibles are there in the (k + 1)th week if we ignore births and deaths?
Ik+l =,fsklk
(3)
Sk+l = Sk " f SkIk + B
where B and f are constant parameters of the model. Note that the
higher the value of f the more easily the disease is spread between
individuals.
Computer results
In an article, Anderson and May (1982) use this model to discuss measles
epidemics in a typical city in Britain and in Nigeria. They assume
initial values of infectives, and susceptibles of Io = 20 and So = 30000
respectively. They also choose f = 0.3 x 10-4 it can be shown that this
value of f corresponds approximately to one infective infecting a single
susceptible, during one week. In their article the number of new births,
in one week in the typical British city, is given as B = 120.
Numerical iteration of (3) with these values gives the results shown in
Figure 9.4.1. This shows that there is a dramatic increase in the number
of infectives every 130 weeks (roughly 2-3 years). This corresponds to
the epidemics observed every 2-3 years in Britain.
Now let us examine what happens when we change the parameter B
to 360 (three times its previous value) corresponding to the birth rate
of a developing country such as Nigeria. Here the birth rate is much
higher. The results are shown in Figure 9.4.2. In this case the model
9.4 A coupled model of a measles epidemic 167
k (weeks)
Fig. 9.4.1. Numerical simulation of a measles epidemic for a typical British city,
with B = 120 births per week.
predicts epidemics every year and the epidemics are much more severe.
This result is consistent with observation.
Steady-state solutions
Steady-state solutions are sometimes useful for understanding simple
models. Finding steady-state solutions for coupled difference equations is
the same, in principle, as for single difference equations. Now, however,
both the quantities Ik and Sk are to assume constant values I and S, say.
This leads to a pair of simultaneous equations for I and S. Care must be
taken that all solutions are found. It is also a good idea to verify results
by substitution back into the original equations.
20 000
10000
2000
1500
1000
40-4
500
.104 woe
0 50 100 150 200 250 300
k (weeks)
Solution. Let us denote the steady-state number of infectives by I and the stead y-
state number of susceptibles by S, where I and S are constants. Then,
Ik+l = Ik = I and Sk+1 = Sk = S (4)
are the steady-state solutions. Substituting (4) into (3) we obtain
I=fSI,
S=S-fSI+B,
or
I(1 - f S) = 0, (5a)
f SI - B = 0. (5b)
Our aim is to solve for I and S. From equation (5a) there are two cases to be
considered : I = 0 or S = 11f.
Case (a) : I = 0. Substituting I = 0 into (5b), to determine S, yields B = 0. But
this contradicts the fact that B is a positive constant. Thus, there is no solution of
both (5a) and (5b) corresponding to I = 0.
Case (b) : S = 11f. Substituting S = 1/f into (5a), to determine I , yields I = B.
Hence the pair, S = 1 If , I = B, is a solution of the coupled system (5a) and (5b),
which is easily verified by substitution.
One can also try to find solutions of (5b) first and then substitute into (5a) but
this does not yield any additional solutions.
9.4 A coupled model of a measles epidemic 169
Discussion
The first of the equations (3) shows that if Sk is below 1 If then
Ik+1 <
Ik
so Ik decreases as k increases. Now the second of the equations (3) shows
that for small Ik (after an epidemic is over), Sk increases due to births
until Sk eventually becomes greater than 1 If and the number of infectives
begins to increase again. This signifies the important part played by the
birth rate, as seen in the numerical simulations. Also one can easily see
the advantage in trying to keep the number of susceptibles down below
11f, through vaccination, to prevent epidemics.
A comprehensive discussion of this model is given in the article by
Anderson and May (1982). This article is certainly accessible to students.
Further models of epidemics will be discussed in Chapter 19 of this book.
Exercises 9.4
1. In some diseases infected individuals do not become immune but return to
the susceptible class. Ignoring births, put together an argument which gives
I k+1 = f Sklk
Sk+1 = Sk + Ik - f SkIk
4. Modify the measles model in the text given that a constant fraction y of those
recovered are reinfected. You will need to introduce an additional variable.
5. Modify the measles model in the text given that a constant fraction 6 of the
new births are vaccinated.
6. In the model in the text we assumed that the number of new infected
individuals per week was given by f IkSk. Another model assumes a constant
probability p of contact between two individuals selected at random.
170 Non-linear difference equations and population growth
(a) Given the probability p above, what is the probability that a given sus-
ceptible does not have contact with a single infected, picked at random?
Hence find the probability that a given susceptible does not have contact
with any of the Ik infected and hence argue that the expected number
of new infected individuals during week k is
Sk-Sk(1-P)lk
(b) Put e -Y = 1- p and show that the expected number of new infections is
given by
Ik+1 = Sk [1 - e-Y'k ]
Hold Sk constant and sketch a graph of the number of new infections
as Ik varies. Is this model better than the one used in the text when the
number of infectives is large?
7. Using the result of Exercise 6 show that the modified model from Exercise 1
becomes
Ik+l = Sk [I -e-71k
Sk+1 =Ik+Ske".
Hence show that
Ik+l = (M - Ik) [1 -e-ylk]
where M is a constant.
8. (Computer Simulation). Modify the model in the text for a measles epidemic,
using the result of Exercise 6. Use y 0.3 x 101. What differences do you
observe?
9. In a host-parasite system, a parasite searches for a host on which to deposit
its eggs. Define
Nk = {number of host species in kth breeding season}
Pk = {number of parasite species in kth breeding seasons}
f = {fraction of hosts not parasitized)
c = {average number of eggs laid by parasite which survive}
A = {host rate, given that all adults die before their offspring can breed).
Argue that Nk and Pk satisfy
Nk+l = Af Nk and Pk+1 = cNk [1 - f].
for what values of the growth rate r will the population tend to the carrying
capacity K if the initial population is close to K?
To answer this question the scaled variable
Xk = Kk (2)
Xk+1 = Xk + rXk (1 - Xk )
(4)
The solutions of the linear difference equation (5) are only approximate
solutions of (4) but they have the advantage that they can be found in
closed form.
The closed-form solution of equation (5) is
Yk =(1-r)kYo
where Yo can be calculated from the initial population No using equations
(2) and (3). Hence an approximate solution to the original equation (1)
is given by
Nk K +K(1 -r)"Yo. (6)
Now for ' 1 - r I < 1, the term (1 - r)k tends to zero as k approaches
infinity. Thus the approximate solution converges to the steady-state
solution. This means that for 0 < r < 2 the approximate solution is
attracted to the steady-state, while for r > 2 the solution is repelled.
This is in precise agreement with numerical experiments on the original
non-linear difference equation, provided that the initial population No is
sufficiently close to the carrying capacity K.
The types of behaviour just discussed occur not only for the discrete
logistic equation, but for many other difference equations as well. In
general, given a non-linear difference equation, a steady-state solution s is
called an attractor, or is said to be stable, if all the approximate solutions
obtained from the linearized equation converge to the steady-state. If, on
the other hand, all these solutions (apart from the steady-state solution)
diverge away from the steady-state, then it is said to be a repellor, or is
said to be unstable.
9.5 Linearizing non-linear equations 173
Linearization by differentiation
The above discussion illustrates the main ideas involved in the lineariza-
tion of a difference equation. The method used, however, is only suitable
when the function on the RHS of the difference equation is a polynomial
of low degree. A more generally applicable technique uses a formula
involving the derivative to approximate the RHS. This new technique
will now be explained.
Consider a general first-order non-linear difference equation
Xk+1 = g(Xk) (k = 0, 1, ...) (7)
where the sequence members Xk are real numbers and g is a sufficiently
smooth function. Our interest is in the following problem : obtain an
approximate solution of the non-linear equation (7) if the initial value Xo
is close to a steady-state solution Xk = s.
The first step is to introduce a new variable Yk such that
Xk = S + Yk. (8)
The new variable Yk equals 0 when Xk is at the fixed point s and is small
when Xk is close to the fixed point. We shall assume that Yk stays small
enough to ensure that
Yk2 is very small compared with Yk. (9)
Yk+1 = g'(S) Yk
Yk = (g'(s))k Yo
and hence, after substituting for Xk using (8), we see that the desired
approximate solution of the non-linear equation
Xk+1 = g(Xk),
when Xo is close to the fixed point s, is
Xk =S+ ((gF(s))' (Xo - S). (12)
It is simple to use this approximate solution to determine under what
conditions the fixed point is an attractor (see Exercise 4). The method of
linearization by differentiation is summarized in Table 9.5.1.
Properties of the solution: If Ig'(s)I < 1 then the fixed point is an attractor
and the solution converges to s, while if Ig'(s)I > 1 then the fixed point
is a repellor and the solution does not converge to s.
Xn+1 = g(Xn)
with
g(x) = x2.
The fixed points s are given by the solutions to the equation
s=g(s)=S2
and so are s = 0 and s = 1. Thus the non-zero fixed point is s = 1.
Now
Ig'(1)I = 2
and thus
Ig'(S)I > 1.
177
178 Models for population genetics
explained. In genetics, the laws are due to the monk, amateur scientist
and plant breeder Gregor Mendel, who performed a series of famous
experiments in the mid 1800s.
A and a.
In any given individual in the species, the alleles can occur in just one of
the combinations
Reproduction
Our aim is to determine how the proportions of the three genotypes in
a population vary with the time. To achieve this aim, it is necessary to
know how the genes are transmitted during reproduction. The only type
of reproduction we shall consider is the most common one in which the
male and female both contribute one gamete (an egg or a sperm) to the
offspring. The gamete contains only one of the alleles A or a, the allele
being chosen in accordance with the following law.
Mendel's first law. The allele in the gamete is chosen at random from
the two alleles in the genotype of the parent.
0
as
B
Fig. 10.1.3. Genotypes of a hypothetical population of six individuals.
Thus the possible genotypes of the offspring are AA and Aa, noting that aA is the
same as Aa. Thus both AA and Aa occur with the same probability 112, in this
example.
X(AA)
G(AA) = (1a)
N
where G(AA) is the proportion with genotype AA, .K(AA) is the number
with genotype AA, and N is the total number in the population. Similar
10.1 Some background genetics 181
0
00 00
00
0
00 00
Fig. 10.1.4. Gene pool for the population in Figure 10.1.3.
1(aa)
G(aa) = N (1c)
P(A) = (3a)
2N
Similarly
2N
As an illustration, we apply the formulae (3) to the population in
Figure 10.1.3, for which X(AA) = 1, .N'(Aa) = 3 and .N'(aa) = 2. This
gives
2 x 1+ 3 5
P(A) 2x6 12
and
7
Pa =2x2+3
() 2x6 12
in agreement with the answers stated earlier.
Proportions as probabilities
The proportions of genotypes and alleles in the population can be re-
garded as the probabilities of certain events. For example, since G(AA)
is the proportion of AA genotypes in the population, we may also write
probability that an individual selected
G(AA) (4)
at random has genotype AA
Again, since P (A) is the proportion of A-alleles in the gene pool, it is
also the probability that an allele selected at random from the gene pool
happens to be an A-allele. By Mendel's first law this implies that
probability that a gamete from a randomly
P(A) = selected individual in the population (5)
contains an A-allele
Interpretations similar to (4) or (5) also apply to G(Aa), G(aa) and P(a).
10.1 Some background genetics 183
Example 2. Suppose that 40% of a population are of genotype AA, 40% are of
genotype Aa, and 20% of genotype aa. Find the probability that a gamete from a
randomly selected individual in the population contains an A-allele. Find also the
probability that it contains an a-allele.
Example 3. For the problem in Example 2 state the possible genotypes of the
offspring and find the probability of each genotype occurring, given that the two
parents are chosen at random.
Solution. The possible genotypes for the offspring are AA, Aa and aa. For offspring
of genotype AA, each parent must contribute a gamete with the allele-A. Whether
this occurs for the male parent is independent of whether it occurs for the female,
the probability of occurrence being P(A) in either case. Hence the probability of
both occurring - and giving offspring of genotype AA - is the product
P(A)P (A) = [P (a)] 2 = 0.36.
Similarly, for offspring of genotype aa, the probability is
p(a)p(a) = [P(a)]2 = 0.16.
Finally, offspring of genotype Aa may be obtained by either the male parent con-
tributing the A-allele and the female the a-allele, or vice versa. Hence the probability
of offspring having genotype Aa is
P(A)P(a) + P(a)P(A) = 2P(A)P(a) = 0.48.
Thus we have obtained the probabilities of all the possible genotypes.
184 Models for population genetics
Once the offspring are born, the genotype composition of the popula-
tion alters and this must be taken into account in predicting the genotypes
of those born subsequently. Various assumptions which enable us to do
this will be considered in the following sections.
Finally, recall that the sum of probabilities of mutually exclusive events,
exhausting all possible cases, is 1. This leads to two useful identities :
G(AA) + G(Aa) + G(aa) = 1, (6)
You should also check that these identities hold in the previous examples.
This section has covered all the background theory from genetics
needed for the task of predicting the changes in genotype proportions.
You may well be interested, however, in learning more about genetics.
If so, you should consult one of the more specialized books on genetics,
such as Hexter and Yost (1976) or Hartl (1980).
Exercises 10.1
1. Match each of the symbols on the left, which were introduced in the text,
with the appropriate description on the right :
(a) G(AA) (a') total population
(b) P(A) (b') number with genotype AA
(c) X (AA) (c') proportion of A-alleles
(d) N (d') proportion with genotype AA
Interpretation as probabilities were given in the text for G(AA) and P (a). Write
down similar interpretations as probabilities for G(Aa), G(aa) and P(a).
2. Verify the identities stated in the text,
G(AA) + G(Aa) + G(aa) = 1 and P (A) + P (a) = 1,
by using the definitions (1) and (3) in the text.
3. A population consists of seven individuals : two of genotype AA, two of
genotype Aa and three of genotype aa.
(a) Draw a diagram like that in Figure 10.1.3 showing the genotypes and
then write down the proportions of each genotype in the population.
(b) Draw a diagram like that in Figure 10.1.4 showing the gene pool for the
population and then write down the proportions of each of the alleles A
and a in the gene pool.
(b) Verify your answer to part (a) by using the formulae (3) in the text.
5. A population consists of 20% genotypes AA, 40% genotypes Aa, and 40%
genotypes aa. Calculate the allele proportions P (A) and P (a).
6. In humans there is a blood group known as the `MN group'. It is composed
of individuals with the genotypes MM, MN or NN, each genotype consisting of
a combination of the genes M and N.
(a) What are the possible genotypes of the children of an MM and an NN
parent? What is the probability of occurrence of each genotype?
(b) As in (a), but now suppose both parents are of genotype MN.
7. Suppose that, in a survey of people from the MN blood group, it was found
that 115 were of genotype MM, 115 were of genotype MN, and 3/5 of genotype
NN.
(a) Calculate the proportions of the genes M and N respectively, in the gene
pool.
(b) Give the probabilities for each of the possible genotypes occurring in
children of those surveyed, assuming the mating is random.
8. The colouring of fur in rabbits is determined by a single pair of genes, to be
denoted by A and a. The allele A is responsible for the pigmentation of the fur,
while the allele a is responsible for white fur. The allele A is dominant to the
allele a.
(a) What genotypes have coloured fur?
(b) Suppose that, in a population of rabbits, 25% of genotypes are AA, 50%
of genotypes are Aa, and 25% are aa. Calculate P (A) and P (a).
(c) Deduce the probability that a particular offspring is a white rabbit,
assuming the mating is random.
9. Show from appropriate formulae in the text that
P (A) = G(AA) + 1 G(Aa),
2
(a) mating occurs at random: the choice of mate does not depend on
the mate's genotype.
kth generation
Born Mates
Born Mates
Similar notation will be used for the values of the other quantities
associated with the population, at that time. The first of our models will
now be described.
The other major assumptions, to be used in this section, are equal
survival and equal fertility. These assumptions are as follows.
(b) Equal survival: each genotype has the same chance of surviving from
the fertilized egg to the end of the generation, where mating occurs.
(c) Equal fertility: each couple produces on average, the same number
of viable sperm and eggs, regardless of the genotypes in the couple.
Example 1. Find an expression for Gk+1(AA) in terms of the the allele proportions
from the kth generation.
Solution. Recall that, by definition,
probability that an individual at the beginning
Gk+1(AA) = of the (k + 1)th generation has genotype AA
Hence, by assumption (c), equal fertility,
probability that an offspring from a given
Gk+l (AA) = couple in the kth generation has genotype AA
This, in turn, is equal to the product of the separate probabilities that each parent
contributes an A-allele. Each of these latter probabilities is equal to the proportion
of A-alleles in the gene pool at the time of mating (by assumption (a)) and hence
to Pk(A). Thus
G;+1(AA) = Pk(A)Pk(A) = [Pk(A)]2.
Similarly, results are easily obtained for Gk+1(Aa) and G;+1(aa). These
are summarized as
G;+1(AA) = [Pk(A)]29
Gk+1(Aa) = 2Pk (A)Pk (a), (1)
Gk+1(aa) = [Pk(a)]2)
These equations are always valid for random mating.
The British geneticist R.C. Punnet deduced the formulae (1) from a
diagram which is now called the Punnet square. In the Punnet square,
shown in Figure 10.2.2, the alleles present in the gametes of the male
parents and their expected proportions are arranged along the top of
the square, and those of the female along the side. The genotypes of
the offspring and their expected proportions are contained within the
square.
Thus the genotype proportions of the offspring, just after birth, can be
read from the Punnet square as [Pk (A)] 2 for the AA genotype, [Pk (a)] 2
for the as genotype and 2Pk(A)Pk(a) for the Aa genotype. Note that the
cross terms in the Punnet square add to give the Aa genotype since Aa is
the same as c A.
10.2 Random mating with equal survival 189
A Pk(A) a Pk(a)
A AA Aa
Pk(A) Gk+I(AA) = (Pk(A))2 Gk+I(Aa) = Pk(A)Pk(a)
a aA as
Pk(a) Gk+I(aA) = Pk(a)Pk(A) Gk+I(aa) = (Pk(a))2
Example 2. Find a difference equation for the recessive gene a given that a constant
fraction r of each genotype of fertilized egg survives to the end of each generation.
Solution. We now carry out the two steps in detail for Pk+l(A) in terms of Pk(A).
STEP 1: Define Nk+1 as the total number of fertilized eggs giving rise to the
(k + 1)th generation and let .Kk+i (AA), .Kk+i (Aa) and .Kk+i (aa) denote the num-
ber of each genotype of the fertilized eggs. Now X;+i (AA) = G;+1(AA)N;+, with
190 Models for population genetics
similar results for the other genotypes. Thus, from (1), valid for random mating,
X;+1(AA) = [Pk(A)]2Nk+1,
.Kk+1(Aa) = 2Pk(A)Pk(a)Nk+1, (2)
Xk+10IOI) = [Pk(a)]2Nk+1
From assumption (b), equal survival, a fraction r of each genotype of fertilized egg
survives to the end of the (k + 1)th generation, giving Xk+1(AA), Xk+1(Aa) and
"k+1(aa). Thus
.1k+1(AA) = r.Kk+1(AA) = r[Pk(A)]2Nk+1,
.Kk+1(Aa) = r.Kk+1(Aa) = 2rPk (A)Pk (a)Nk+1 (3)
STEP 2: The right-hand side of (2) contains the allele proportions Pk(A) and Pk(a)
(measured at the end of the generation). To obtain a difference equation we need
to obtain the allele proportions at the end of the (k + 1)th generation. Now
{number of a-alleles}
Pk+1(a)
{total number of alleles in gene pool
2.Kk+l (aa) + V1k+1(Aa)
(AA) + 241k+1(0)
Substituting from (3), we thus obtain,
[Pk(a)]2 + Pk(A)Pk(a)
Pk+1(a ) = .
[Pk (A)] 2 + 2Pk (A)Pk (a) + [Pk (a)] 2
To eliminate Pk(A) we use the identity
Pk(A) + Pk(a) = 1
obtaining, after some straightforward algebra, the rather simple difference equation
Pk+1(a) = Pk(a) (4)
Thus we have obtained a difference equation for Pk(a) as required.
Po(a) = 1 (5)
2
The solutions of the difference equation (4) are all constant functions of k. From
the initial condition (5) it therefore follows that
Pk(A) = 1 (k = 09 19 2, ...
2
Equations (1) now give the expected genotype proportions as constant function of
k:
From assumption (b), equal survival, the genotype proportions cannot change
during a generation, so it follows that
1 1 1
Gk+l (AA) = 4, Gk+I (Aa) = 2, Gk+l (aa) = (k = 09 1, 2, ...)
4
Note that Gk+l (AA) + Gk+l (Aa) + Gk+l (aa) = 1.
Exercises 10.2
1. In each of the following cases state whether the assumption is consistent with
the assumptions in the equal survival model (Model I). If the answer is `no', state
which assumption of Model I would be violated.
(a) Between birth and parenthood 10% of each of the genotypes AA, Aa, as
die.
(b) Between birth and parenthood 5% of the genotype AA die, 10% of Aa
and 15% of xa die.
(c) The genotypes AA mate only with the genotypes AA or Aa.
(d) The genotypes AA have, on average, twice as many offspring as the geno-
types xx.
4. Suppose, as seems likely, that the simple random mating model is applicable
to members of the MN blood group in humans. Suppose, furthermore, that the
genotype proportions have already reached equilibrium values, as in Exercise
3. Given that 36% of the members of this blood group are of genotype MM,
calculate the expected percentages of genotypes MN and NN.
10.3 Lethal recessives, selection and mutation 193
STEP 2: Now
{number of a-alleles}
A_ (a) =
(total number of alleles in gene pool)
_ 2A1k+l (aa) + Xk+1(Aa)
2Xk+l (AA) + 2.N'k+l (Aa) + 2Ak+1(aa)
Substituting from (2), we obtain
Pk+1(a) = Pk(A)Pk(a)
[Pk(A)]2 + 2Pk(A)Pk(a)
Finally, to eliminate Pk(A), the identity
Pk(A) + Pk(a) = 1
is used and we obtain
k+1
Xk+1 = 1+Xk (4)
'
which is the required difference equation.
x
X1 i xo __ Xo
X2 =
1+X1 1+ 1+
XI
0
x
X3 __ X2 1+2X0 Xo
1+X2=1+X0=1+3Xo.
1+2X0
Xk = 1 +XokXo (5)
To verify that this is the solution we check that it satisfies both the initial condition
and the difference equation. The initial condition is satisfied since putting k = 0 in
(5) gives Xo. To check the difference equation is satisfied note that when (5) is
used in (4)
Xo
LHS =
1+(k+1)Xo
Xk
RHS =
X
1 + Xk
I+kXo Xo
X
1 + 1+kXO
1 + (k + 1)Xo
Note that we assume that both male and female are equally likely to
survive. To illustrate the new assumption we consider the frequently
studied example of the Peppered Moth Biston betularia. The dominant
forms, AA and Aa, are a coal black colour whereas the recessive form,
aa, is a pale speckled colour. For moths living in industrial cities the pale
speckled coloured moths are less well camouflaged and thus are more
likely to be eaten by predators. We thus say that the pale speckled moths
are at a selective disadvantage compared with the coal black moths in
industrial cities.
The details of setting up the model and deriving a difference equation
is left to the exercises, where it is shown that the proportion of a-alleles,
Xk = Pk (a), satisfies the non-linear difference equation
(p -1)Xk -4- Xk
Xk+1 - 6
1 + - 1)Xk
where P is given by P = r2/rl. The number P is called the relative fitness
of the genotype as and measures the fitness to survive of the recessive
as genotype relative to the AA and Aa genotypes. Note that the special
case P = 1 corresponds to the equal survival example from Section 10.2
(Model I) and the case and P = 0 corresponds to the lethal recessive
gene model (Model II) of this section.
A numerical iteration of equation (6) has been carried out for various
values of the parameter P starting with an allele distribution of 90%
recessive and 10% dominant. The results are shown in Figure 10.3.1.
Note that as P decreases the recessive allele proportion tends to decrease
10.3 Lethal recessives, selection and mutation 197
Pk(a)
1.0-1 03=1
t -
0.8--
''©.
°''b -©..© /3 = 0.9
0.6 -
GL
00
1.
0.4 -I ©
* o.
© ©.. = 0.7
0.2-
/3= 0.3
4.44 0
5 10 15 20 25
k (Generation)
Fig. 10.3.1. Recessive allele proportions for Model III for various values of the
parameter /3 starting from Pk (a) = 0.9.
very rapidly within the first few generations. The decrease is more gradual
for higher values of P.
It has been estimated that the relative fitness for the pale speckled
Peppered Moth in Manchester, UK, compared with the coal black moth
is approximately /j = 0.7. Figure 10.3.1 shows that there is a rapid
decrease from 90% to 40% in only 10 generations.
Another interesting application of these ideas is to the genetic dis-
order sickle cell anaemia in West Africa. Here a defective recessive
gene causes a minor chemical change in the blood cells. Those who
inherit the recessive gene from both parents have a low survival rate.
However, the gene is not wholly recessive. The hybrid genotypes Aa
are slightly affected, but not enough to cause a fatal condition. In fact
the hybrids have an enhanced resistance to malaria, which is prevalent
in West Africa. Thus the hybrids, Aa, are the most likely to sur-
vive, followed by the pure dominants, AA, and then the recessives, aa.
To model this situation it is necessary to postulate different survival
fractions for each genotype. This case is considered, in detail, in the
exercises.
198 Models for population genetics
Mutation model IV
Another factor which can affect the distribution of genes in a population
is mutation. This happens when external factors (for example, background
radiation) cause an allele A to change into an allele a. The rate at which
mutation of genes occurs is usually very small (typically a fraction 10-5
or 10-6 of the alleles per generation). Also, mutation more commonly
acts to change a dominant gene into a recessive one.
Model IV for population genetics assumes random mating, equal sur-
vival of genotypes and also incorporates mutation of A-alleles into a-
alleles. Incorporating mutation into the modelling of population genetics
requires us to modify Step 2 of the procedure set out in Section 10.2.
This is illustrated in the following example.
Example 3. Find a difference equation for Pk(a) given equal survival and that the
A-alleles mutate to the a-alleles at a rate of y alleles per allele per generation.
Solution. Assume that, for each genotype, a fraction r survives from the beginning
of the generation to the end of the generation, and assume that during a single
generation a fraction p of the A-alleles mutates into a-alleles.
STEP 1: This step is exactly the same as in Example 2 of Section 10.2 and we thus
obtain
Xk+1(AA) = r[Pk(A)]2Nk+1,
Vk+1(Aa) = 2rPk (A)Pk (a)Nk+1, (7)
Note that, even though some of the alleles have changed due to mutation, the total
number of alleles remains the same so the denominator is still twice the sum of
Ak+1(AA), .IVk+1(Aa) and Xk+1(aa).
Substituting (7) into (8), and letting Xk = Pk(a), we obtain
Exercises 10.3
1. (a) For Model II (lethal recessives) show that the proportion of A-alleles
satisfies the difference equation
1
Yk+1 =
2 - Yk
where Yk = Pk(A).
(b) Iterate and hence guess the solution of this difference equation. Prove
that your guess is a solution.
Xk+1= Xk
1 + Xk
By making the transformation Zk = 1 /Xk in the difference equation show that it
reduces to a linear difference equation. Hence obtain the solution of the original
difference equation.
3. Consider Model III (natural selection) where the recessive genotype as has a
lower survival rate than the other genotypes.
(a) Show that Xk = Pk(a) satisfies
(/3-1)Xk+Xk
k+1 1 + ($ -1)Xk
10. (a) Find all the steady-state solutions for the difference equations for Pk (a)
for Models I-IV in the text.
(b) For those who have studied Section 9.4 of Chapter 9: determine which of
the steady-state solutions in (a) are attractors.
Part three
Models with Differential Equations
11
Continuous growth and decay models
In this chapter some problems of growth and decay will be studied for
which differential equations, rather than difference equations, are the
appropriate mathematical models. Such problems include :
The differential equations which arise from the above problems are all
of the first order. The two methods of solution which we explain are
sufficient to solve all the differential equations which arise in the next
three chapters. The theoretical background for these two methods is
contained in Chapter 5. The first of the two methods, which applies only
to linear differential equations, is very similar to the method already given
in Section 8.1 for solving linear difference equations. The continuous
models used in this chapter are similar to the discrete models discussed
in Chapter 9.
203
204 Continuous growth and decay models
ac=ax+b (1)
Homogeneous equations
The following example illustrates how to find all the solutions when the
differential equation is homogeneous. It corresponds to the choice a = 2
and b = 0 in (1).
Example 1. Find all the solutions of the first-order linear homogeneous constant-
coefficient differential equation
ac=2x. (2)
Solution. To guess a solution, recall that the exponential function is its own deriva-
tive. Hence our first guess is x = et. This gives ac = x, however, which is out by a
factor of 2. Hence our next guess is
x=e2t (3)
RHS = 2x = 2e2t.
Hence both sides will be equal if 2 = 2. So once again we get the
particular solution x = e2t for the differential equation (2). The rest of
the solutions are then obtained as in Example 1.
Inhomogeneous equations
It is now easy to solve an inhomogeneous equation, such as
.z=2x-6. (6)
Second, add the solutions for the homogenized equation .z = 2x, which
were obtained as equation (2) in Example 1, to get
x=3+Ce2t
This formula gives all the solutions of (6) (by the superposition theorem
for inhomogeneous equations in Section 5.3).
Variables separable
These differential equations have the form
dx=
dt f ( xg(t)
) (7)
is of the variables separable type since it has the form (7) with f (x) _
x(1 - x) and g(t) = t. The reason for the term `variables separable' in
describing these equations will become clear later, after the procedure for
solving them has been explained.
Constant solutions
These solutions, which are also called equilibrium or steady-state solutions,
are important because they are easy to find and they provide a framework
for the study of other solutions. To find these constant solutions, solve
the equation f (x) = 0; in the example (8) this gives
x=0 orx=1
It is easy to check that each of these is a solution of the differential
equation since, when substituted in (8), each gives
LHS=.z=O
RHS=x(1-x)t=0.t=0
The graphs of these constant solutions x = 0 and x = 1 are horizontal
lines, as shown in Figure 11.1.2.
If x = ¢(t) is any other solution, its graph cannot intersect either
of these lines (by the existence-uniqueness theorem of Section 5.2) and
11.1 First-order differential equations 207
x-axis
X=0
-----w t-axis
hence its graph must lie entirely within one of the three horizontal strips
determined by these lines.
Other solutions
The procedure for finding non-constant solutions of variables separable
differential equations will now be illustrated. The idea behind the proce-
dure is to try to `separate' the variables so that x appears on one side, t
on the other.
Solution. The constant solutions are x = 0 and x = 1. The initial condition places
the solution x = 4(t) in the horizontal strip between the lines x = 0 and x = 1, as
in Figure 11.1.2. Hence, for all time, this solution satisfies
0<x<1. (9)
STEP 1: Divide both sides of the differential equation by f (x) = x(1 - x), which
is non-zero. This gives
1 dx
x(1 - x) dt t
STEP 2: Integrate both sides, with respect to t, from the initial time 0 to the
current time t. This gives
t=
1 dx
dt = ` t dt.
I It 0 x(1 - x) dt r-0
208 Continuous growth and decay models
STEP 3: Apply the rule for integrating by substitution to the LHS. This gives
(since 0(0) = i and ¢(t) = x)
I I
x(1-x)dx= tdt.
Jx=3 r=0
X= 1
1 +1x dx=it2
2
ln(1- x) + ln(x)Ix_ 1 = i t2
X- 2 2
Note that the arguments of log are both positive since 0 < x < 1, by (9). Hence
In ( 1"x) - ln(1) = 2t2
I-X = e2 (10)
x= 1 - -11 2
1+e2`
STEP 5: Check the solution. Substituting (10) back into the differential equation
gives
dx to 3 `2
LHS =
dt (1 + '
e2`2 1
RHS=x(1-x)t= 1 t2 It2
1+e 1+e
Thus the differential equation is satisfied. Also, (10) shows that, when t = 0,
x=1- 1
1
e°2'
= 1-
2
1 1
so that the initial condition is satisfied. Thus (10) does indeed give the required
solution of the differential equation.
Exercises 11.1
1. Copy and complete the following table to show the classification of the given
first-order differential equations.
210 Continuous growth and decay models
ac+2x=0
x = tx3/2
2. From the solutions found in the text for the differential equation ac = 2x,
(a) find the solution which satisfies the initial condition x = 3 when t = 0,
(b) show that the solution which satisfies the initial condition x = xo when
t=0is
x = xoe2t.
3. (a) Use the method of Example 1 in the text to find all the solutions of the
differential equation ac = -3x.
(b) To which types of differential equations is the method used in (a)
applicable?
(c) Find the solution of the differential equation ac = -3x + 6 which satisfies
the initial condition x = xo when t = 0.
(b) Which constant solution does the differential equation have? Find the
solution which satisfies the initial condition x = 1 when t = 0 by
separating the variables as in Example 2 in the text.
dy
dt
= A2y + IoAie-A1`, Al * A2
where Al and A2 are positive constants.
(a) If there is no iodine 131 in the thyroid initially solve the differential
equation and show that
10 ,1 [e1t - e-A2c
A2 Al
(b) After a nuclear test in Colorado in the USA in 1964 the following
estimates of Al and A2 were made for a Colorado deer population. They
were
A 1 = 0.126 and '2 = 0.107.
Use the model to estimate the maximum percentage increase in the
amount of iodine 131 in the deers' thyroids.
212 Continuous growth and decay models
Continuous models
The study of population models helps bring into focus the distinction be-
tween discrete and continuous models. The population models discussed
in Chapter 9 were discrete, being appropriate for species of animals which
breed during specific breeding seasons, equally spaced. The population
models of interest in this chapter, however, are more appropriate for
large populations which reproduce continuously, rather than at regular
intervals. Human populations are naturally modelled in this way, as are
certain types of microscopic organisms.
In the continuous models, the number of individuals in a population at
time t will be modelled by a solution N = 4(t) of a differential equation.
Thus both of the variables N and t will assume all the real values in
some interval, fractional and irrational values included.
The justification for using N as a real variable in this way is that, when
the population is sufficiently large, differences of one or two individuals
are of little consequence. At the end of the problem, we simply round
N to the nearest integer value. A similar justification applies to our
use of t as a real variable : in the absence of specific breeding seasons,
reproduction can occur at any time; for a sufficiently large population,
it is then natural to think of reproduction as occurring continuously.
Microscopic organisms
Populations of microscopic organisms are attractive to model since they
may be grown in the laboratory. This enables data concerning their
growth to be collected easily and the environment to be controlled.
Examples of such micro-organisms for which data are available are
yeast cells and E. coli. The former are involved in brewing and in the
commercial production of certain vitamins; the latter are a species of
bacteria which occur in the intestines of man and other animals.
Both of these examples are single-celled organisms which reproduce by
11.2 Exponential growth 213
dividing into two a process known as binary fission. The cells absorb
nutrients which are dissolved in the liquid in which they are immersed.
Thus the amount of nutrients available to the cells can be controlled.
Thus the assumption (1) implies that the population of cells grows
exponentially.
The formula (4) implies that the population grows indefinitely large at
an ever-accelerating rate, as can be seen from the typical graph sketched
in Figure 11.2.1.
Such unrestricted growth is impossible in practice since eventually the
214 Continuous growth and decay models
N = Noea`
No
t-axis
population runs out of space and nutrients. The above `J' curve is often
observed, however, during the initial stages of a population's growth.
Hence
e10a = 1.5,
lOa = ln(1.5).
Thus, to two decimal places,
a = ln(1.5)/ 10 = 0.04 (per minute per cell).
11.2 Exponential growth 215
In (N)-axis
ln(N)=1n(No)+at
ln(NO)
t-axis
.
4
.
2 . ZO
I
Time (hours)
0 1 2 3
Human populations
In adapting the preceding growth model to human populations, we must
take into account deaths as well as births. A plausible assumption is that
births and deaths both occur at a rate which is proportional to the size
N of the population at any time t. Hence we may write
dN
= aN - #N = (a - #)N (6)
Tt
where a and /3 are positive constants denoting the average rate of births
and deaths, on average (per head of population per year). If a - /3 > 0
then (6) has the same form as (2) with a = a - P, and so the model
predicts exponential growth, given by (4). It is left as an exercise to show
that, if a population grows exponentially, then there is a fixed length of
time that it takes to double.
An interesting historical account of theories of population growth is
given in Hutchison (1978). The idea that populations grow exponentially
can be found in the works published by Graunt (1662) and Malthus
(1798), who made estimates of the times taken for various populations to
double. All sorts of data concerning the growth of populations around
the world can be found in the Demographic Yearbook published by the
United Nations. For example, the following annual growth rates for
the world's population, shown in Table 11.2.1, were obtained from this
source.
An article in The Age newspaper (25 May 1989) quoted United Nations
sources as stating that the world's population growth, after having slowed
11.2 Exponential growth 217
down in the 1970s, was speeding up again, and that the current population
of about 5.25 billion people would double in 39 years, at present rates.
The above figures are in rough agreement with an exponential growth
model for the world's population for the period considered with
a growth rate a of about 0.02 new individuals per head of population
per annum. The exponential growth model cannot be valid in the long
term, however, as the population would run out of food and space. A
more realistic model of population growth, taking such limitations into
account, will be described in the next section.
Exercises 11.2
1. Let a be any real constant. Find all the solutions of the differential equation
dN = aN
dt
by using the appropriate method from Section 11.1. Hence show that the solution
which satisfies the initial condition N = No when t = 0 is
N = Noe°`.
5. Find approximately the value of the growth rate a for the population in
Figure 11.2.3.
218 Continuous growth and decay models
Date Population
1901 3773801
1911 4 455 005
1921 5 435 734
1933 6 629 839
1947 7 579 358
1954 8 986 530
1961 10 508 186
1966 11 550 462
1971 12 755 638
1976 13 548 472
1981 14 576 330
Can you think of any historical reasons for any anomalies in the data?
K
N
Fig. 11.3.1. Logistic decline in growth rate.
or
dN
dt
=aN 1- N
K
(2)
N= K ai
(3)
_ 1)e +1
Note that this formula gives the correct initial condition N = No when
t = 0. As t -+ oo, moreover, the term involving a-°1 approaches 0 as a is
positive. Hence
lim N = K.
t-)-oo
220 Continuous growth and decay models
N-axis
A
K N=K
N=O
t-axis
It also follows from (2) that dN/dt is positive. Hence the graph of the
solution (3) is the `S'-shaped curve shown in Figure 11.3.2.
Thus the model predicts that the population increases steadily from
the value No and approaches the carrying capacity K as the time be-
comes arbitrarily large. The relevance of the model to various types of
populations will now be discussed.
Microscopic organisms
The logistic model is reputed to give reasonably good predictions for
the behaviour of populations of yeast cells, bacteria, and protozoans
(the most primitive form of animal life), when grown under suitable
laboratory conditions.
To test the relevance of the model to the growth of such populations,
we shall refer to Table 11.3.1, which is based on actual laboratory
measurements of Carlsen (1913).
When the points in Table 11.3.1 are plotted, as in Figure 11.3.3, they
are seen to lie along an `S'-shaped curve. In this respect, at least, they
are in agreement with the predictions of the logistic model.
Graphs similar to that in Figure 11.3.3 for the growth of the population
of yeast cells may be seen for example in Emlen (1984), page 43, Emmel
(1976), page 103, Hutchinson (1971), page 24, and Kormondy (1976),
page 78.
How can we further test the agreement with the logistic model? A
simple geometric answer is provided by Figure 11.3.1: the logistic model
11.3 Restricted growth 221
Time Number of
in hours yeast cells
t N
0 10
2 29
4 71
6 175
8 351
10 513
12 584
14 641
16 651
18 662
N-axis
700
1W
600
500
400
JO
300
200
100
t-axis
0 2 4 6 8 10 12 14 16 18 (hours)
Table 11.3.2.
Time in hours t 2 4 6 8 10 12 14
QN AN)
dt
N
29 71 175 351 513 584641
this table have been plotted in Figure 11.3.4. It can be seen that the
points do indeed lie approximately on a straight line, in accordance with
the logistic model.
Fisheries management
Models for population growth find ready application in the fishing in-
dustry, which aims at maintaining a permanent supply of fish. Too
much fishing in a particular year might so deplete the population that
it would take a long time to recover or it might even become extinct.
Too little fishing, on the other hand, might leave the population intact
but result in a smaller harvest than necessary. Biologists employed by
the fishing industry are therefore interested in determining the maximum
rate, somewhere in between these two extremes, at which fish can be
harvested without reducing the population in the long term. Models for
population growth play an important part in determining this maximum
rate, as we now show.
11.3 Restricted growth 223
aN11-K I.
l
A little calculation shows that the desired choice is N = K/2. This means
that the population should be maintained at half the carrying capacity.
To get the maximum value of dN/dt we now substitute this value of N
back into the quadratic to get
(dN aK
dt max 4
This is the maximum rate at which fish can be harvested, if the population
is to be kept at a constant size.
A discussion of how the answer for the maximum value depends on the
particular population model is given in Ginzburg (1985), pages 130,131.
Some pros and cons of the logistic model, when used in this context, are
discussed in Walter (1981).
Human populations
The data in Section 11.2 suggest that the world's population is currently
growing exponentially with a growth rate of about 0.02 per year. If this
were to continue for the next three centuries, however, the population
would increase by a factor of (1.02)300, which is about 380. Hence the
average density of the world's population (over the surface area of the
inhabited countries) would increase from its 1985 value of 36 people per
km2 to about 13 680 per km2. This latter figure is truly fantastic : less
than one square metre of land for each living person. Long before this
happened, of course, the food supply would have been exhausted and
the population would have exceeded its maximum sustainable size.
As a more realistic alternative, the logistic model was proposed for the
growth of human populations by Verhulst in the middle of the nineteenth
224 Continuous growth and decay models
century. He used the logistic model to estimate the maximum values for
the population of various countries.
The use of the logistic model to study human populations was revived
in 1920 by Pearl and Reed. They compared the census figures for the
population of the USA from the years 1790 to 1910 with the values which
could be predicted from the logistic model. The remarkable agreement
between the actual and predicted values is shown in Table 11.3.3.
To get the predicted values, Pearl and Reed assumed that the logistic
equation (2) was satisfied with N denoting the population of the USA at
a time t years after some initially chosen year; hence the population is
given as a function of the time by the solution (3) of the logistic equation.
They chose the parameters a and K in such a way as to make the formula
(3) give the actual values of the population in the years 1790, 1850 and
1910. As explained in the exercises, the values they obtained for these
parameters were
a = 0.03134 per year,
K = 197 273 000 individuals.
The formula (3) then gives the remaining predicted values in Table 11.3.3.
As figures from later censuses became available, the remarkable agree-
ment between actual and predicted values for the population of the USA
continued, as can be seen from the first column of Table 11.3.4. After
1950, however, the predicted values consistently underestimated the ac-
11.3 Restricted growth 225
Table 11.3.4. Good then bad news for the logistic model.
tual size of the population, and by 1980 the actual population was well
in excess of the previously estimated carrying capacity, K.
The failure of the Pearl and Reed model to give a realistic prediction
of the maximum sustainable population for the USA, highlights the dif-
ficulties of making predictions about the growth of human population
in the long term. An obvious difficulty is that human beings can change
their environment in such a way as to invalidate the values of the param-
eters previously relevant to the model. Thus, for example, technological
advances in agricultural production and distribution can improve the
supply of food and thereby increase the carrying capacity. Advances
in medical science can decrease the death rate and thereby increase the
growth rate. It is easy to think of many other factors under the control
of human beings which can affect the values of the parameters.
Human population models are discussed in Braun (1983), Section 1.5,
Hutchinson (1978), pages 22-23, and in Keyfitz (1977), pages 213-220.
Exercises 11.3
1. This exercise is about the population of yeast cells from which the data in
Table 11.3.1 were obtained.
(a) Estimate the growth rate a and the carrying capacity K for this popula-
tion by comparing Figure 11.3.4 with Figure 11.3.1. What is the accuracy
of your estimates?
(b) Hence compare the number N of yeast cells at time t predicted by the
formula (3) in the text with the observed values given in the table, for
t = 0, 2,14.1... ,18 hours.
226 Continuous growth and decay models
2. Let a, K, and No be positive real numbers such that 0 < No < K. Use
separation of variables to show that the solution of the logistic equation
dN
dt
_ aN I- NK
which satisfies the initial condition N = No when t = 0 is
N= K
(_1)e_t2t+1
No
3. Repeat Exercise 2, but this time suppose that No > K. Sketch the graph of a
typical solution.
(a) What does your graph tell you about what happens in the long term if
the population initially exceeds the carrying capacity?
K = N1
- No) - NO(N2 - N1)
1 Ni - N0N2
5. (a) Show that if the population N reaches half its carrying capacity when
the time t = t1 then the solution of the logistic equation in Exercise 2
may be written
K
e-a(t-tl) + 1
(b) On page 32 of Braun (1975) it is stated that Pearl and Reed calculated
that the population of the USA reached half its carrying capacity in
April 1913. It follows from the above formula that in year t
197 273000
N= e-0.03134(t-1913.25) + 1
Does this formula give the values predicted in Table 11.3.2? Comment.
The growth rate, instead of decreasing linearly with V (as in the logistic model),
decreases exponentially with t, its value at time t being found empirically to be
oce-At
(c) Sketch the graph of a typical solution. What happens to the volume as
[The law of growth for the tumour is called the Gompertz growth law. Further
discussion may be found, for example in Braun (1983), and Rubinow (1975),
page 43.]
Radioactive decay
A typical example of such a process is the decay of a radioactive element,
such as radium. Since the decrease in mass is caused by the emission
of alpha particles, the decay is really a discrete process. The mass of
an alpha particle, however, is very small compared with the mass of the
sample and hence it is appropriate to regard the mass as a quantity which
can change continuously. On average, the larger the sample, the greater
will be the number of alpha particles emitted per unit time. The simplest
way to model this is to assume that, at any time,
rate of decrease mass of sample
is proportional to
of mass of sample still present
To express this as a differential equation, let m denote the mass of the
sample still present at time t and so obtain
dm _
dt - -km (1)
228 Continuous growth and decay models
where k is a positive constant. The minus sign ensures that the derivative
of m with respect to t is negative; hence the mass of the sample decreases
as time goes on.
Since this differential equation is linear homogeneous with constant
coefficient, it can be solved by the method of Section 11.1. The solution
satisfying the initial condition m = mo when t = 0 is found in this way
to be
m= moe-kr.
(2)
Drug absorption
Another process which also leads to an exponential decay model is the
absorption of drugs from the bloodstream into the body tissues. When
a drug is administered by an injection, it mixes with the blood. As time
goes on, the amount of the drug in the bloodstream diminishes, being
absorbed by the body tissues or excreted from the body. When medical
staff administer a drug it is important for them to know how much to
give in the next injection too little and the drug is ineffective, too
much and undesirable side effects could result.
The significant quantity to monitor is the concentration of the drug in
the bloodstream, which is defined as the amount of drug per unit volume
of blood, and is usually measured in mg/litre. For most drugs the rate
of absorption from the bloodstream increases with higher concentration.
As with radioactive decay, the simplest model consistent with this is the
assumption that, at any time,
f rate of decrease {concentration}.
of concentration is proportional to
Exercises 11.4
1. In the differential equation (1) in the text:
(a) What quantities do the symbols m, t, and dm/dt stand for?
(b) Why is there a minus sign?
(c) Assume mass is measured in grams and time in years. What are the SI
units for the decay constant k? What are its dimensions?
2. Obtain the solution of the differential equation (1) in the text which satisfies
the initial condition m = mo when t = 0. What makes the solution decrease more
rapidly : large k or small k ?
3. Show from the solution (2) in the text that, if the mass of a radioactive sample
decays from mo to mo/2 in time T, then T = ln(2)/k.
230 Continuous growth and decay models
4. Given that the half-life of radium is 1600 years, what is the value of its decay
constant k ? How long does it take for the mass of a given sample to decrease to
3 of its value? a of its value? n of its value?
5. Carbon-14 dating. While a plant or animal is living, the ratio of 14C to 12C
in its tissues is a small constant, the same for all living tissue. When a plant
or animal dies, however, this ratio decays exponentially with a half-life of 5730
years.
A sample of charcoal was found at the cave at Lascaux in France containing
the famous prehistoric painting, for which the ratio of 14C to 12C had decayed
to 14.5% of its original value. How many years ago did the wood grow?
(Further information about carbon-14 dating is given in Braun (1983), Sec-
tion 1.3.)
6. (a) Match each symbol occurring in the differential equation (1) in the text
with the symbol which plays a similar role in the differential equation
(4).
(b) Why is there a minus sign in the differential equation (4)?
(c) Suppose the concentration is measured in mg/litre and time is measured
in hours. What are the units for µ?
(d) Use the solution given in the text for (1) to write down the solution of
(4) which satisfies the initial condition c = co when t = 0.
(e) How long does it take for the concentration to reduce to half its initial
value? Express your answer in terms of µ.
(d) What is the shortest safe time that a second injection may be given so
that side effects do not occur?
10. One method of administering a drug is to feed it continuously into the blood
stream by a process called intravenous infusion. This may be modelled by the
linear differential equation
dc __µc+D
dt
where c is the concentration in the blood at time t, p is a positive constant, and
D is also a positive constant which is the rate at which the drug is administered.
(a) Find the constant (or equilibrium) solution of the differential equation.
(b) Given c = co when t = 0, find the concentration at time t. What limit
does the concentration approach as t -+ oo? Compare with your answer
to part (a).
(c) Sketch the graph of a typical solution.
12
Modelling heat flow
Some typical processes from everyday life which involve the flow of heat
from one region to another are the heating of beverages, food and living
areas, and the cooling of foodstuffs in refrigerators. The flow of heat
involved in such processes is best described by mathematical models.
This chapter introduces some simple mathematical models which are
based on Newton's law of cooling and Fourier's law of heat conduction.
These laws lead to very simple differential equations of the type studied
in Chapter 11. At the end of this chapter these ideas are used to model
the loss of heat from an insulated water pipe. The model makes some
unexpected predictions.
The only concept from physics which is assumed initially is that of
temperature which indicates the hotness of a body, and is measured
with a thermometer.
232
12.1 Newton's model of heating and cooling 233
The model
Although the model for cooling applies to any heated object, we shall
stay with the cup of coffee as an illustration. We aim at predicting how
the temperature of the coffee changes with time.
The intuitive starting point for modelling this problem is the idea that
the greater the difference between the temperature of the coffee and that
of the surrounding room, the greater will be the rate of cooling of the
coffee. The simplest mathematical model consistent with this requirement
is to have
1 temperature of
us - t the surrounding room f '
du__
T 2(u-us) (2)
Example 1. A cup of coffee is initially at boiling point, 100 °C. The temperature
of the room is 20 °C. Find the temperature of the coffee as a function of the time.
Solution. Let u be the temperature of the coffee after time t. Since us = 20, the
di, fJrerential equation (2) is now
du = -A(u - 20) (3)
dt
with the initial condition u = 100 when t = 0. The differential equation, being
linear constant coefficient, may be solved by the method of Section 11.1.
First, we find all solutions of the homogenized equation
du
(4)
dt
Try u = em' where m is a constant to be determined. By substituting in the homog-
enized equation (4) we find that m = -A. Hence one solution of (4) is u = e-)J.
As (4) is homogeneous, every solution therefore has the form
u = Ce-A` (5)
for some real constant C.
Second, we guess the original equation (3) has a particular solution in which u
is a constant. This means that du/dt = 0 and hence by (3) that u - 20 = 0. Thus
a particular solution of (3) is the constant solution
u = 20. (6)
Finally, all the solutions of the original equation (3) are obtained by adding the
solutions (5) and (6) to get
u=Ce-A`+20. (7)
It is left as an exercise to check that (8) satisfies both the differential equation (3)
and the initial condition. This formula gives the temperature u of the coffee as a
function of the time t.
12.1 Newton's model of heating and cooling 235
Temperature
Behaviour of solutions
The constant solution (6) obtained in the course of the above working
has the interesting physical interpretation that if the coffee is initially at
room temperature 20 °C, then it will stay at this temperature indefinitely.
It is therefore called the steady-state temperature for the coffee.
Note that the formula (8) does not provide a complete answer to
Exercise 1 since the value of the parameter A has not yet been specified.
In spite of this, however, it is possible to indicate the general shape of a
typical graph of temperature against time, as in Figure 12.1.1.
The graph was obtained from (8) by observing that
u = 100 when t = 0, (9)
du
<0 for t >_ 0 (10)
dt
and
The property (11) means that, as the time approaches oo, the temperature
of the coffee approaches the steady-state temperature, equal to that of
the surroundings.
at some other time, besides the initial one. A second way, which will
be explored in the next section, is to reformulate the model, taking into
account the physics of heat transfer. This will show how the parameter 2
depends on such factors as the mass of the heated object, the material of
which it is composed, and its surface area. Such information will make
the model more versatile and will play an essential role later in our study
of the effect of insulating a hot water pipe.
Exercises 12.1
1. An object is at a temperature u which is colder than the temperature us of
its surroundings. Let % > 0. Which of the following differential equations predict
that the temperature of the object will increase with time?
(a) u = Au (b) iu = A(us - u)3
(c) iu = -A(u - us)2 (d) u = -AI us - uI
2. Let %, us and uo be constants. Show that the solution of the differential
equation
du =-A(u-us)
dt
which satisfies the initial condition u = uo when t = 0 is
u = (uo - us)e zt + us
[Hint : Use the method of Example 1 in the text, which utilizes the fact that the
differential equation is linear constant coefficient]
3. Repeat Exercise 2, but use the method of separation of variables.
4. A cold beer is at a temperature of 10 °C. After 10 minutes the beer is at
a temperature of 15 °C. Find how long it takes for the beer to warm to 20 °C,
given that the temperature of the room is 30 °C.
5. Sketch the general shape of the graph of temperature against time in Exercise
2, assuming . > 0 and uo < us.
6. From the expression in Exercise 2 for temperature as a function of time, say
what happens when uo = us. Explain this physically.
7. Suppose that, instead of Newton's law of cooling (2) in the text, the law of
cooling is
du_
=f(u-us)
dt
for some function f. Explain the physical interpretations of each of the following
conditions on the function f and state whether they seem realistic.
(a) f (0) = 0. (b) f (x) > 0 for x > 0
(c) f (x) < 0 for x < 0 (d) f (x) = -f(-x).
8. A student was seen to enter a tutor's office at 4.00 p.m. The tutor was later
12.2 More physics in the model 237
at the bar at 4.30 p.m. drinking heavily. At 6 p.m. the cleaners discovered the
student's body in the tutor's office and called the police. The police first measured
the temperature of the corpse at exactly 6.30 p.m. as 30 °C and later at 8.30 p.m.
as 27 °C. The temperature of the office remained at a constant 25 °C. [Hint:
Assume Newton's Law of cooling, and use the solution obtained in Exercise 2.]
(a) What is us ?
(b) Why is it a good idea to set t = 0 to correspond to 6.30 p.m.? What is
the initial temperature?
(c) Write down an expression for the temperature at time t and hence
determine 2 from the information given in the question.
(d) Hence determine the time of death, assuming that the temperature of the
student just before the murder was 37 °C (normal body temperature).
9. As mentioned in the text, Newton's law of cooling assumes that air at room
temperature is blown past the cooling body (`forced cooling'). For cooling in still
air ('natural cooling') a better model is to assume that the rate of temperature
decrease of the cooling body is directly proportional to the 5/4th power of
the difference between the temperature of the body and the temperature of the
surrounding air.
(a) Introduce appropriate notation and thus write the law for natural cooling
as a differential equation. Is it linear?
(b) Show that the temperature u of the cooling body at time t is given by
the formula
which can flow from a hotter substance to a colder one, thereby raising
its temperature. The microscopic origin of heat is the motion of atoms
and molecules which compose the substance. Heat is a form of energy,
which in the SI system is measured in joules (J).
The change in the heat of a given substance depends on both the mass
of the substance and the change in temperature, in a manner which we
now describe.
As to the dependence on mass suppose, for example, that 1 joule of
heat flows into 1 kg of the substance, raising its temperature by 1 'C. It
then seems reasonable to suppose that 2 joules of heat will be required
to raise the temperature of 2 kg by the same amount. More generally,
we assume that when the temperature of a given substance is raised by
a fixed amount
{change in heat} is proportional to {mass of substance}. (1)
If, furthermore, we keep the original mass of the substance fixed but
wish to raise its temperature by 20C we might expect that it would
take twice as much heat energy and, in general, for a fixed mass of the
substance,
{change in heat} is proportional to {change in temperature}. (2)
in (3) occur during a time interval of length bt. Dividing both sides of
(3) by 6t and then letting 6t approach 0 gives in the limit
dH du
dt - cmdt (4)
This gives the desired relationship between rate of change of heat and
rate of change of temperature at any given instant.
The constant c depends on the type of substance being heated (alu-
minium, brick, glass, etc.) and is called the specific heat of that substance.
Specific heats of some common substances are shown in Table 12.2.1. In
the table the specific heats are said to be `taken at 20 °C' to indicate that
the formula (3) is valid provided the temperatures stay close to 20 °C.
Thus we can see from Table 12.2.1 and equation (3) that metals such
as aluminium, copper and iron have lower values of c and thus require
much less heat energy to raise their temperature than does water and
food products such as butter fat, lamb and potatoes (which contain a
substantial proportion of water).
We now model how the heat is lost to (or gained from) the surround-
ings. The key quantity to consider is the rate of change of heat energy
contained within the object, which we have denoted by dH/dt.
First, since the heat loss occurs at the surface of the object, it seems
reasonable to suppose that
f rate of changel
is proportional to !surface area l (5)
1 of heat f t of object f
Second, when cooling is expressed as a loss of heat, Newton's law of
cooling says that
rate of change Itemperature)
of heat } is proportional to difference (6)
Denoting A as the surface area of the object, (5) and (6) may be combined
to give
dH -hA(u
= - us) (7)
Tt
where h is a positive constant of proportionality.
The constant h is known by several different names - the convec-
tive heat transfer coefficient, surface conductance, and the surface convec-
tion coefficient. Experimentally determined values of this coefficient are
shown in Table 12.2.2 under various circumstances. The unit for h is
watt metre -2 °C-I (where the watt is the unit of power, or 1 Joule s-1).
We can now use the relationship (4) between heat and temperature to
write (7) as the following differential equation for the temperature.
du _ hA
(u - us)
dt me
Exercises 12.2
1. Discuss the significance of the minus sign multiplying the RHS of the
differential equation (7) in the text.
2. Find the parameter A in Newton's law of cooling for an iron plate whose total
surface area is 2 m2 which is cooling in a stream of air flowing over the plate at
35 m/s. Assume the mass of the plate is 2 kg.
3. Consider a 3 kg plate of iron with total surface area 2 m2, initially at a
temperature 150 °C. In each of the following cases, find how long it takes to cool
to a temperature 100 °C if the temperature of the surroundings is 200C and
(a) the plate is in still air,
(b) air flows over the plate at a speed of 35 m/s.
Useful data are given in various tables in the text.
Steady-state conduction
We will consider the conduction of heat through insulating material
between the inner and outer walls of a house, as in Figure 12.3.1 below.
Suppose that the inner wall is at a temperature of 20 °C and the outer
wall is at a temperature of 10 °C. Thus heat flows from the inner wall
242 Modelling heat flow
Inner wall Outer wall
Heat Insulating Heat
flowing material flowing
out:
Jjoules/min
to the outer wall. The temperature inside the insulating material varies
continuously between the temperatures of the inner and outer walls.
As heat flows from the inner to the outer wall, some of the heat
goes into raising the temperature of the insulation. We suppose that
eventually the temperature at each point inside the insulation reaches a
steady-state; this steady-state temperature is independent of the time but
varies continuously with respect to distance, from inside to outside. When
the temperatures inside the insulation have reached this steady-state, the
rate of flow of heat going into the insulation must equal the rate of flow
of heat coming out.
Intuition also suggests that the rate of heat flow, along the x-direction,
12.3 Conduction and insulation 243
Cross-section
Outer wall
Heat
flows out
will depend on the drop in temperature per unit length in this direction.
A larger drop per unit length will produce a larger rate of heat flow. The
simplest model consistent with this idea is to assume that
J is proportional to {temperature gradient}. (3)
Solution. Let J joules per minute be the rate at which heat flows across a cross-
section of area 3 m2 parallel to the walls at a distance x metre from the inner wall
(as in Figure 12.3.2). We assume steady-state temperatures, hence J is a constant,
which is to be determined.
Let u be the temperature at distance x. Hence by (4) the differential equation
du J
dx kA
is satisfied where A = 3 m2. The initial condition is u = 500'C when x = 0. Solving
the differential equation by antidifferentiation gives
J
u=-kAx+500
12.3 Conduction and insulation 245
Exercises 12.3
1. Complete the solution to Example 1 in the text by verifying the claims made
concerning the solution of the differential equation satisfying the given initial
condition.
2. Give reasons why the following would not be suitable to use in place of
Fourier's Law.
du 2
(a) J=-A du
dx
.
du .
(b) J = _A2
dx
[Hint: Consider dimensions.]
Exercises 3,4,5,6,7,8 refer to the rectangular slab of material shown below. It may
be regarded as a wall of a heated room or of a furnace.
Cross-section
Inner face
The slab, of thickness C, has an inner face and an outer face, each of area A.
The inner face is at a uniform temperature ua, and the outer face is at a cooler
temperature ub. Heat is assumed to flow straight through the slab from inner to
outer face. All points on a cross-section at distance x from the inner face have the
same temperature u. Heat flows through this cross-section at the rate J. Assume the
temperature has reached the steady-state. In Exercise 7(c) below, you will express
J in terms of the other parameters.
3. In terms of the notation introduced above, what is the value of the temperature
246 Modelling heat flow
(a) when x = 0,
(b) when x = t?
4. In our model we have assumed that the heat flows straight through the
slab from inner to outer face, none of it escaping out the other sides. Is this
assumption more appropriate when t is large or when t is small?
5. Given that the temperature has reached the steady-state, what follows about
the value of J as x increases from 0 to t ?
6. On the basis of physical intuition, decide the effect on the value of J of each
of the following separate changes.
(a) Increasing ua.
(b) Increasing ub.
(c) Decreasing t.
(d) Increasing A.
(e) Replacing the material by one with greater thermal conductivity k.
(a) What does this tell you about the shape of the graph of u against x?
Hence state why this graph is as shown in the diagram below.
(b) Use the diagram below and the interpretation of the derivative as a slope
to obtain du/dx in terms of Ua, Ub, and t.
(c) Deduce that
J =kA
t
and hence verify your answers to Exercise 6.
u-axis
x-axis
B
8. (a) Write down Fourier's law as a differential equation for the temperature
u as a function of the distance x. The differential equation will involve
the parameters k, A, and J. Why can the differential equation be solved
by antidifferentiation?
12.3 Conduction and insulation 247
(b) Find the solution of the differential equation which satisfies the initial
condition u = ua when x = 0.
(c) Now use the value of the temperature u at the outer face to find J in
terms of the parameters k, A, Ua, Ub, &"-
(d) Does your answer agree with that found in Exercise 7(c)?
9. Suppose a stone slab has a surface area of 10 m2 and thickness 2.7 m. The
inner and outer faces are at steady temperatures of 20 °C and 0 '*C respectively.
Given that the thermal conductivity of stone is 2.7 W m-1 °C-1, calculate J from
your answer to Exercise 7(c).
Exercises 10,11,12,13,14,15 below refer to the figure below (which gives an end-on
view of the slab in the diagram above).
Slab
Besides the assumptions already listed for Exercises 3,4,5,6,7,8, the following
additional assumptions also apply. The outer face of the slab, at a temperature of
ub, is cooled by a stream of cold air at a temperature of usb.
Newton's law of cooling applies to the loss of heat from the outer face of the
slab to the cold air. From (7) of Section 12.2, this means that
dHb
hbA(ub - usb)
dt
where -dHb/dt is positive and denotes the rate at which heat is being lost to the
cold air at the outer face. hb denotes the heat transfer coefficient between the outer
face and the cold air. Similarly, the inner face of the slab, at a temperature of ua,
is heated by a stream of hot air at a temperature of usa. Newton's law of cooling
applies to this face also. In Exercise 15 you will express the rate of heat flow J in
terms of the temperatures of the hot and cold air (rather than the temperatures of
the inner and outer faces, over which we have no direct control).
10. (a) Arrange the temperatures ua, ub, usa, usb in increasing order.
(b) What are the signs of (i) ub - usb and (ii) ua - usa ?
(c) Extend the diagram shown in Exercise 6 to show also the temperature
of the hot air (corresponding to points with x < 0) and to show the
temperature of the cold air (corresponding to points with x > C).
248 Modelling heat flow
(d) At which points is there a discontinuity in the graph you have drawn in
(c)?
11. Check, from the formula given above for dHb/dt and your answer to
Exercise 10(b), that -dHb/dt is positive.
12. (a) The steady-state temperatures having been attained, what is the rela-
tionship between the rate J at which heat is arriving at the outer face
and the rate -dHb/dt at which heat is being lost from the outer face?
Deduce that
J = h6A(ub - usb)
(b) Hence find the temperature u6 at the outer face in terms of the temper-
ature usb of the cold air (and the parameters J, hb, A).
13. Newton's law of cooling at the inner face may be written, with a suitable
choice of notation, as
dH
_ haA(ua - u.).
dt
What is the sign of dHa/dt? What is the physical significance of this quantity?
14. (a) The steady-state temperature having been attained, what is the relation-
ship between the rate J at which heat is entering into the slab from the
inner face and the rate dHa/dt at which heat is being transferred to the
inner face from the hot air? Deduce that
J = haA(usa - ua).
(b) Hence find the temperature Ua of the inner face in terms of the temper-
ature u. of the hot air (and the parameters J, ha, A).
15. (a) From your answers to Exercises 7(c), 12(b) and 14(b), show that
A(usa - ub)
J= ha-1 +hb-1
+tk-i
(b) What features of this answer agree with your physical intuition?
Exercise 16 refers to the diagram below, which shows two slabs of material joined
together.
Inner slab
12.4 Insulating a pipe 249
The notation to be used for each slab is similar to that used in the previous
exercises. The inner and outer faces of the combination will be assumed to be at
the respective temperatures ua and ub, where ua > ub. The thermal conductivities
of the respective slabs are denoted by kl and k2. Assume the temperatures have
reached the steady-state; hence J is the same for each slab of material. For the
first slab
J_-k1A du (0<x<C).
dx
16. (a) Write out in words the meaning of the above differential equation for
the temperature in the first slab.
(b) Write down a similar differential equation for the temperature in the
second slab.
(c) Solve these two differential equations, using the fact that u = ua when
x = 0 and U = ub when x = 21.
(d) Hence show that
A(ua - ub)
e(kl-1 + k2-l )
a
b
flowing through the insulation and then escaping to the surrounding air.
The problem is to determine the extent to which the insulation reduces
loss of heat from the pipes.
The model
Our model will be based on Fourier's law of heat conduction (to describe
the flow of heat outwards through the insulation) and Newton's law
of cooling (to describe the loss of heat from the outer surface of the
insulation to the surrounding air). Instead of imagining the heat as
flowing across plane faces as in Section 12.3, however, it will now be
regarded as flowing across cylindrical surfaces, which we now describe.
For each r > 0, the points which are at the same distance r from the
axis of the pipe form a cylinder of radius r, as shown in Figure 12.4.2.
(The cylinder is a surface, not a solid.) We assume the cylinder has the
same length L as the pipe. The cylinder coincides with the outer surface
of the pipe when r = a and with the outer surface of the insulation
when r = b. If r lies between a and b, then the cylinder lies inside the
insulation.
Note that, since a circle of radius r has circumference 21rr, the surface
area of this cylinder is given by
Let
J temperature at each point
u - of the cylinder of radius r,
(3)
dr (2 kL) r
Because the steady-state has been attained, J and -J/(2irkL) are con-
stants; hence (3) is a very simple differential equation for u as function
of r.
The inner boundary of the insulation is assumed to be at the common
temperature of the pipe and the water. Hence the initial condition for
252 Modelling heat flow
(since in the steady-state the rate at which heat is being lost to the air
must equal the rate at which heat is flowing through the insulation). The
last two equations give
J = hA(b)(ub - us) (6)
Solution. We convert all quantities to SI units. Thus for the first pipe
Example 2. Repeat Example 1 but now assume the pipe has outside diameter 5 mm
and has 2 mm of asbestos insulation. For asbestos take k = 0.11 W m-1 °C-1 and
h = 8Wm-2oC-1.
Solution. We obtain J = 13.5 W with the insulation and J = 11.3 W without it.
- b-axis
a k Outer radius of insulation
h
Fig. 12.4.3. Graph showing rate of heat loss verses outer radius of insulation.
Practical considerations
To obtain a rule of thumb as to whether a pipe should be insulated or
not we graph the rate of heat loss J against the outer radius b, as given
in equation (8). The details have been left to the exercises and the graph
is presented in Figure 12.4.3. The turning point is at
b = k* (9)
h
If b is below this value then the rate of heat loss initially increases as b
increases. If b is above this value, however, then adding more insulation
decreases the rate of heat loss.
One way to guarantee that adding insulation always decreases the rate
of heat loss is to make sure that the outer radius of the pipe is greater
than the critical value given in equation (9). Thus in Example 2, this
condition requires the pipe to have outside diameter a > k/h = 0.11/8
m or 13.75 mm. Alternatively, for a given pipe size we should choose the
type of insulating material to give a critical value (9) less than the radius
of the given pipe.
This chapter has been based on a model introduced in the Open
University module : `Modelling Heat' (1975).
Exercises 12.4
1. (a) State briefly the meaning of each of the symbols occurring in equation
12.4 Insulating a pipe 255
dr 2nkL r
(c) Steady-state temperatures having been attained, which symbols in the
above differential equation are constants? What are the dependent and
independent variables? Solve the differential equation subject to the
initial condition u = u,y when r = a.
(d) Hence derive the formula (7) in the text,
Ub
__ J
In
b
a + uw.
2. (a) Explain the meaning of the symbols occurring in Newton's law of cooling
(5) in the text,
dH__
T hA(b)(ub - us)
J = 2n(u,, - us)hL b
l+kbin(Q)
(d) The water being at a higher temperature than the surrounding air, what
is the sign of uw - us ? Given the geometrical interpretations of a and b
in the text, what is the sign of ln(b/a) in the above formula?
3. (a) What do you expect to be the effect on the rate of heat loss from the
heated pipe of each of the following separate changes?
(i) The temperature of the water is increased.
(ii) The temperature of the surroundings is increased.
(iii) The length of the pipe is increased.
(iv) The thermal conductivity of the material insulating the pipe is de-
creased.
(b) Check your answers to part (a) of this exercise by using the formula
from Exercise 2(c).
4. (a) Suppose that f is a differentiable function which does not assume the
value 0 at any point of its domain. Show that (c/ f)'(x) = 0 precisely for
those x such that f'(x) = 0, where c is any non-zero constant.
256 Modelling heat flow
Statement of problem
In a dye factory a large vat is used to mix dye and water. The water
flows in at a rate of 6 litres/minute and the dye flows in at a rate of 2
litres/minute. The mixture is drawn off at a rate of 8 litres/minute. Ini-
257
258 Compartment models of mixing
Pure water
100 1
Initially Later
tially the vat contains 100 litres of pure water. How does the strength
of dye in the water change with time?
We shall set up a compartment model for the mixing of dye and water
within the vat to produce a mixture of the two liquids. In this problem,
since the total flow rate of ingredients into the vat equals the flow rate
of mixture out of the vat, the volume of mixture in the vat is constant.
However, the amount of dye in the tank changes with time.
We now introduce some relevant terminology for problems involving
mixing of liquids. We then look at some introductory examples of
the small time-interval technique which we use to formulate differential
equations for this type of problem.
Concentration
In mixing problems in general we refer to the mixture as the solution, the
substance we are introducing as the solute and the liquid which dissolves
the solute as the solvent. In our problem the dye-water mixture is the
solution, the dye is the solute, and the pure water is the solvent.
To obtain a measure of the strength of dye in the dye-water mixture
we introduce the concentration of the mixture, which is the ratio of
the amount of solute to the amount of solution. In our problem it is
convenient to define the concentration as a ratio of volumes :
{volume of solute}
{concentration} _ (1)
{volume of solution}
This definition assumes that the solute is homogeneously mixed in the
solution. Note that the maximum concentration, which is unity, occurs
when the whole of the mixture is pure dye; the minimum concentration,
which is zero, corresponds to pure water.
13.1 A mixing problem 259
Solution. Firstly,
volume of mixture
flowing out = 86t (litres).
Now let us obtain an expression for the change in the volume of dye
in the mixture over some small time interval St. Let
x
volume ofofdye
olumeume dye _ volume
volume of dye
(2)
entenng vat f
entering vat l leaving
leaving vat
vat
Note that the volume of dye is the appropriate measure of the amount
of dye here since the concentrations in the problem are given as the
ratio of volume of dye in the mixture per unit volume of the mixture. In
the following example we obtain approximately the volume of dye which
leaves the vat in a small time interval b t.
Example 2. Find the approximate change in volume of dye bx in the vat shown in
Figure 13.1.1, during a small time interval b t.
Having accounted for the input and output of dye to the vat we are
finally ready to derive a differential equation for the volume of dye in
the vat at time t. First we divide (6) by b t and then we let b t -+ 0. Since
dx=limbx
dt 6t-+o 6t
we obtain
dx 6 8x
(7)
dt 100'
But the problem was posed in terms of finding how the strength, or
concentration, of the dye in the mixture varied with time. To answer this,
a differential equation for the concentration of dye in the vat at time t is
more relevant.
Since the volume of mixture in the tank remains constant at 100 litres,
then from (1)
x
C = or x = 100c. (9)
100
Substituting (9) into (7) we obtain
d
dt(100c)
= 6 _ 8 x 100c
100
which simplifies to
do
= 0.06 - 0.08 c. (10)
dt
The initial condition for this differential equation is
c=0 at t=0 (11)
----------------------- Steady-state
concentration
t-axis
Time
Fig. 13.1.2. Graph of how the concentration of dye varies with time, from
equation (12).
dc
differential equation by setting
do
T = 0. From (10) we see that, if c < 0.75,
then T > 0. Thus all concentrations with co < 0.75 increase with time.
They all tend to the steady-state value 0.75.
Summary
A summary of the procedure for formulating a differential equation for
mixing problems is now given.
Dye flowing in
Volume of dye
in the vat
Mixture flowing out
Exercises 13.1
2. In each case determine the amount of the substance entering the compartment
in a time interval b t. Give the appropriate units.
(a) Pure dye flows in at the rate 9 litres/min. Give the amount of dye.
(b) Dye, at a concentration 5%, flows into a vat at a rate of 2 litres/min.
Give the amount of dye.
(c) Water containing 0.5 kg of salt per litre enters a tank at a rate of 3
litres/min. Give the amount of salt.
3. In each case below assume that there is an amount x of the substance present
at time t and then determine the amount bx of the substance entering and the
amount leaving the compartment in the small time interval from t to t + bt.
Indicate units, and also indicate where your answer is exact, and where it is
approximate for very small b t. Hence determine the differential equation for the
amount of substance and then for the concentration in each case.
(a) Pure dye flows into a vat at a rate of 4 litres per minute and pure water
flows into the vat at a rate of 8 litres per minute. Initially there are 50
litres of pure water in the vat. The well stirred mixture is drawn off at a
rate of 12 litres per minute.
(b) Dye of concentration 40% by volume flows into a vat at a flow rate of
2 litres per minute. Initially there are 100 litres of pure water in the vat.
The well-stirred mixture flows out at the same rate.
(c) A salt-water mixture containing 15 grams of salt per litre flows into a
lake of volume 2000 litres at a flow rate of 10 litres per minute. The
flow rate out of the lake is 16 litres per minute and pure water flow into
the lake at a flow rate of 6 litres per minute. Assume the mixture is well
stirred.
6. At 6.00 pm on a Friday night a public bar opens and is rapidly filled with
clients of whom the majority are smokers. The bar is equipped with ventilators
which exchange the smoke-air mixture with fresh air. Unfortunately cigarette
smoke contains 4% carbon monoxide, CO, and a prolonged exposure to a
concentration of more than 0.012% of CO can be fatal. The bar has dimensions
of 20 m by 15 m by 4 m and it is estimated that smoke enters the room at a
constant rate 0.006 m3 /minute. The ventilators remove the mixture of smoke and
13.2 Modelling pollution in a lake 265
air at 10 times the rate that smoke is produced. The problem is to find the time
when the concentration of CO reaches 0.012%.
(a) Formulate a differential equation for the concentration of CO at time t.
(b) By solving the differential equation find at what time the lethal concen-
tration will be reached.
7. Using the data of the previous exercise determine the rate at which the
air-conditioners should operate if the concentration is never to reach the lethal
level.
8. A dam contains 106 litres of water. Fresh water enters the dam at a rate of
104 litres per day and the same amount flows out each day. Suppose someone
spills a barrel of pesticide into the dam.
(a) Set up a differential equation for the concentration of pesticide in the
dam.
(b) Suppose that at a given time the concentration is 5 times the safe level
for use by stock. How long before the stock can use the dam?
9. A 100 litre tank originally contains 50 litres of fresh water. Beginning at time
t = 0, water containing 50% of pollutant flows into the tank at a rate 2 litres per
minute and the well-stirred mixture leaves at a rate of I litre per minute. This
exercise involves a situation where the volume of the mixture is not constant.
(a) Find the volume of mixture in the tank as a function of time.
(b) Formulate a differential equation for the volume of pollutant in the tank.
(c) Hence show that the concentration of pollutant at the time the tank
overflows is approximately 48%.
Fig. 13.2.1. The Great Lakes system on the eastern USA and Canadian border.
Fig. 13.2.2. Simple compartment model for removal of pollution from Lake
Superior.
The model
The mechanism for clearing the lake of pollution is the inflow of pure
water which dilutes the water-pollutant mixture. This process is what we
wish to model. We start with the following assumptions :
From the second of these two assumptions we see that the volume of the
lake remains constant.
The quantity of interest is the amount of pollutant in the lake at any
given time, measured in tonnes. In this problem we choose initial time to
be when pollution input to the lake stops. A compartment diagram for
this situation is shown in Figure 13.2.2.
13.2 Modelling pollution in a lake 267
m = fi(t) = 1t in
mass of pollutant)
lake at time t
and define the constants
V = {volume of lake} = 1.2 x 1013 litres (1)
and
F _ flow rate of mixture 6. 5 x 1010 litres/year . (2)
out of lake
We also define the concentration of pollutant in the lake c by
_ m
c
V
We would expect the concentration to be a decreasing function of time
in this problem.
Now,
volume of mixture
flowing out of lake = FSt litres (4)
in time 8t
since F is the flow rate of mixture in litres per unit time. Thus
mass of pollutant volume of mixture fraction of
flowing out of lake = flowing out of lake x pollutant
in time St in time St in mixture
since m/V gives the mass of pollutant per unit volume of the mixture.
Hence by (3) and (5) we obtain
bm ^ -Fm
V
bt.
()6
Now, dividing by b t and letting b t - 0, we obtain
268 Compartment models of mixing
dm fF
(7)
dt V) m
as the differential equation for the mass of pollutant in the lake at time
t.
We define the concentration
mass of
concentration of pollutant
in lake m
c = pollutant in lake = (8)
at time t J volume V.
l of lake f
Substituting into (7) we obtain the differential equation
dc (F) (9)
Thus to find the time t = T when the concentration reaches 10% of its
initial value, we obtain the equation
CO e-(F/V)T.
1 0 = c0 ( 11 )
10
Solving for T (omitting the algebra) yields
T = V 1n(10). (12)
F
Substituting the values for F and V for Lake Superior from (1) and (2)
we find that our model predicts that it takes approximately 425 years
for the concentration of pollutant to be reduced to 10% of its initial value,
even if no more pollutants are put into the lake.
The flow of water through the lake depends on the existence of stagnant
regions cased by eddy currents, thermal layers and wind. Thus the
assumption of a well-stirred mixture is not always true.
Some pollutants may settle on the bottom of the lake.
Bacterial action can affect the concentrations.
The volume of the lake may not be constant over a whole year.
Despite these limitations, the model does provide a starting point
for further investigations. A detailed discussion of this model and its
applicibility is given in the article by Rainey (1967). In the exercises we
also look at simple extensions. These include a model where pollutants
are fed into one lake from an outflow from another lake and a model
where bacteria consume some of the pollution.
Exercises 13.2
1. For Lake Erie the flow rate of water into and out of the lake is 1.75 x 1011
litres/year. The volume of the lake is 4.6 x 1011 litres. How long does it take for
the concentration of the lake to reduce to one quarter of its initial value?
2. For Lake Ontario, about 84% of its inflow comes from Lake Erie. Using the
data in Exercise 1 together with the volume of Lake Ontario as 1.6 x 1011 litres,
find an expression for the concentration of pollutants at time t. [Hint: You will
neeed to account for a variable concentration in the input of pollutant to Lake
Ontario.]
The next two exercises involve a slightly different type of problem which nevertheless
uses the same technique of a small interval analysis to formulate the governing
differential equation.
(a) Using a small interval analysis show that the height of the water in the
tank at time t is given by
dh (a2
V.
dt b2
(b) Toricelli's law states that v2 = 2gh where g is acceleration due to gravity.
By solving the differential equation show that the time taken for the
cylinder to empty, given that it was initially full, is
b2 /2H
T=
g
(c) Design your own experiment to test the validity of this model.
270 Compartment models of mixing
VA )S;Q
zzzzz
V'0
Heat
lost
from
system
1/1
The model
The main idea of the model is to account for the flow of heat into and
out of the water. We assume that it only requires a negligible amount of
heat to raise the temperature of the metal casing to that of the water. A
compartment diagram of the heat flow is shown in Figure 13.3.2.
To arrive at a differential equation we examine the heat energy input
13.3 Modelling heat loss from a hot water tank 271
and output to the system in a very small time interval St. First, we define
u = ¢(t) = {temPerature}
at time t
(1)
heat contained
H =ye(t) = in water
at time t
We also define symbols for the following constants:
q = J rate at which heat t = 3000 watts,
is supplied to water f
us = f temperature ofl
surroundings f = 15 °C (2)
Note that this is positive for u > us. Note also that this is only an
approximate expression since the temperature u changes with time. Over
a small time interval, however, we can neglect the variation in the
temperature u.
272 Compartment models of mixing
Substituting (4) and (5) into (3) we obtain
bH = qbt - hA(u - us)bt. (6)
J=q-hA(u-us). ()7
Our objective is to obtain a differential equation for the temperature. To
do this we must relate heat H to temperature u. We saw how to do this
in Section 12.2 where we argued 6H = cm 6u, where c is the specific heat
and here m the mass of water in the tank. Dividing by b t and letting
b t --+ 0 we thus obtain the relation
dH - cm du
8
dt dt ()
which we then substitute into (7). Hence
du q hA
(u - uS)
dt cm
cm (9)
is the differential equation for the temperature of the water as a function
of the time. From (2), the initial condition for (9) is
u=uo at t=0. (10)
T = 1 In
(fl/-uo (12)
a /3/a - 60
Note that (12) is a well-defined expression since it can be shown that the
term inside the brackets is always positive and thus we never take the
log of zero or a negative number.
We now substitute the appropriate values into (12). Using (2) together
with the typical values
m = {mass of water} = 50 kg
c = {specific heat of water} = 4200 J kg -1 °C-I
A = {surface area of tank} = 1 m2
h = {heat transfer coefficient} = 10 W m-2 °C-1,
we obtain the values a = 4.76 x 10-5 and = 1.5 x 10-2 and hence
T 3413 seconds ^_, 57 minutes.
(Note that the calculation gives the time in seconds since all quantities in
the problem have been converted to SI units.) Thus our model predicts
that it takes approximately 57 minutes for the water to be heated from
15 °C to 60 °C.
In this problem we have introduced symbols for all the constants. Not
only does this make the algebra simpler but it can also be easier to
perform dimensional and physical checks on the answer. Also, now that
we have a general expression we can easily look at times for heating with
different values of the various constants. One can then easily determine
how these times change as one of the parameters (for example, the surface
area of the tank) changes.
Discussion
Is this a good model? To answer this we should check the prediction
of the model with a real water heater. One possible problem with this
model is that we assumed that the temperature was the same at all points
in the water. If the water was well stirred then there would not be a
problem. This will not be true in practice, however, since the hot water
will rise to the top. Nevertheless, the model does incorporate most of the
essential physics. It should be able to be used, albeit cautiously, to gain
some insight into the following questions :
274 Compartment models of mixing
How much better is it to have the heater inside the house rather than
outside?
What advantages are there in insulating the system?
How much saving is made by using off peak heating?
To use the model to help answer some of these questions you should
attempt some of the exercises. Further discussion may be found in Open
University module `Modelling Heat' (1975).
Exercises 13.3
1. Obtain the solution, given in the text, of the differential equation
du
+ au = fJ with u = uo when t = 0,
dt
where a and fi are constants.
2. What is the steady-state temperature for the problem in the previous exercise?
Hence argue that the expression
In
(fl/x-uo 1
fl/a - 60
is a well-defined expression.
3. Imagine that the water heater is inside the house (at temperature 20 °C).
Recalculate the time taken to heat the water.
4. Plot a graph showing the time taken to heat the water as the surface area of
the tank is varied. For all other data use the values in the text.
5. Plot a graph showing the time taken to heat the water as the heat transfer
coefficient h is varied. For all other data use the values in the text.
6. Plot a graph showing the time taken to heat the water as the outside
temperature uQ is varied. Assume that the initial temperature is the same as the
outside temperature. For all other data use the values in the text.
7. Suppose the water heater is perfectly insulated so that no heat is lost to the
surroundings. Modify the expression obtained in the text for the time to heat the
water to 600C. Calculate this time using the data given in the text.
8. Modify the model used in the text to take account of heat lost to the metal
casing. You are given that the specific heat of the metal is 400 J kg-1 °C and the
mass of the metal casing is 5 kg. How much difference does this make to the
result for the time to heat the water? Express your answer as a percentage.
9. The cost of heating is proportional to the time it takes to heat the water. Is
it cheaper to switch the heater off all night (for eight hours) and have it turn
on in the morning or is it better to use a thermostat which switches the heater
on whenever the temperature falls below 60 °C? Assume the temperature of the
surroundings is 5 °C. (You will have to find what temperature the water cools to
in eight hours overnight and how long it takes to reheat to 50 °C, amongst other
things.)
Part four
Further Mechanics
14
Motion in a fluid medium
viscosity
Gases and liquids are collectively known as fluids since they can both be
made to flow if a force is applied. While all fluids will flow, we know
that pouring water out of a bottle is a faster process than pouring, for
example, cream or honey. One obvious distinction between these three
277
278 Motion in a fluid medium
Reynolds' number
When an object falls through a fluid medium, a disturbance is caused as
the fluid is pushed away from the path of the object. The fluid dynamicist
Osborne Reynolds (1883) demonstrated that the fluid flow can change
character as the speed of the object relative to the fluid increases. This
is illustrated in Figure 14.1.1 where the curves, known as streamlines,
14.1 Some basic fluid mechanics 279
(c)
(a) (b)
Fig. 14.1.1. Three types of flow of a fluid around a sphere. Here the frame of
reference is such that the spheres are held fixed.
indicate the motion of the fluid around the sphere. For low speeds the
flow looks the same upstream and downstream as in Figure 14.1.1(a). As
the speed increases a wake is formed formed behind the object and the
fluid recirculates inside the wake, as in Figure 14.1.1(b). The type of flow
in Figure 14.1.1(a) and Figure 14.1.1(b) is called laminar. For very high
speeds the flow in the wake region is no longer smooth but turbulent as
in Figure 14.1.1(c). It might be expected that the modelling of the drag
will be different in each of these three typical situations.
Of course Reynolds realized that not only the velocity of the object
but also the properties of the fluid would affect the type of flow. To take
account of this he introduced a dimensionless parameter (now known as
the Reynolds' number of the motion). The Reynolds' number is defined
by the formula
R = pf(acld
I
where p f denotes the density (mass per unit volume) of the fluid, IM
denotes the speed of the object, q is the coefficient of viscosity of the
fluid, and d denotes a characteristic length of the object (for example,
if the object is a sphere the number d could denote the diameter of the
sphere). The formula shows that the Reynolds' number is proportional
to the speed of the object, but inversely proportional to the viscosity of
the fluid.
For a smooth, falling sphere, a Reynolds' number in the range R <
280 Motion in a fluid medium
1 typically gives a flow of the type in Figure 14.1.1(a), a Reynolds'
number in the range 1 < R < 3000 typically gives a flow of the type
in Figure 14.1.1(b) and for a Reynolds' number R > 3000 the flow is
turbulent, as in Figure 14.1.1(c). The critical Reynolds' number for when
the flow becomes turbulent is generally smaller for spheres with rough
surfaces (e.g. a baseball or cricket ball).
Example 1. Calculate the Reynolds' number for the flow past a raindrop of radius
1 mm falling in air at 5 m/s. (Data: pf = 1.23 kg/m3, ri = 1.8 x 10-5 kg/m s).
Solution. Choosing the characteristic length as the diameter of the raindrop gives
d = 0.002 m.
Substituting this value and the given data in the formula for the Reynolds' number
gives
R = 670.
Since 0 < R < 3000, we might expect that the flow around the raindrop will be like
that in Figure 14.1.1(b).
Stokes' law
For a sphere of radius r moving at velocity ac, Stokes in 1845 derived the
following expression for the magnitude of the drag force :
IFDI = 6nrnIacl.
The formula is valid only for very low Reynolds' numbers (R << 1).
This requirement puts a severe limitation on the applicability of Stokes'
law. Example 1 shows that, even for a sphere as small as a raindrop,
the Reynolds' number is much larger than 1. Stokes' law is applicable,
however, to dust-like particles, such as smoke and other pollutants in
the air, and silt in lakes and streams. In Section 14.3, an application of
Stokes' law to the calculation of the charge of the electron, will be given.
14.1 Some basic fluid mechanics 281
CD
103 +
102 +
10+
D = 2pf
FD 1 CDA.z2.
Here A denotes the cross-sectional area of the object which presents itself
to the fluid, p f is the density of the fluid and .z is the velocity of the
object. For a sphere, A is the surface area of a disk and is thus given in
terms of the radius r by A = nr2.
A plot of CD against the Reynolds' number for a sphere is given
in Figure 14.1.2. Note that the plot in Figure 14.1.2 has been done on
logarithmic graph paper. The value of CD for R between 103 and 2.5 x 105
lies in the range 0.4 to 0.5. As an approximation, the value of CD, for a
sphere, will be taken as the constant 0.45, and furthermore this will be
assumed to be the value for all values of R from 1 up to 2.5 x 105. The
magnitude of the drag force is then given by
1. (a) From the dimensions stated in the text for 1, show that the Reynolds'
number R is dimensionless.
(b) Verify that Stokes' law is dimensionally correct.
(c) Show that the drag coefficient CD is dimensionless.
2. For each of the following motions a typical speed (.z l and characteristic length
d are given. Compute the corresponding Reynolds' number. For motion in air
use pf = 1.22 kg/m3 and t = 1.8 x 10-5 kg m-1 s-1, while for motions in water use
r, = 10-3 kg m-1 s-1 and p f = 1000 kg/m3. Indicate the cases for which Stokes'
law should be applicable.
(a) A peregine falcon in a hunting dive: (.zl = 70m/s; d = 0.15 m.
(b) A minnow swimming in a quiet stream: licl = 1 m/s; d = 0.03 m.
(c) Airborne dust particles settling on a calm day : licl = 2 x 101 m/s;
d=4 x 101 m.
(d) A cruising yacht: licl = 10 m/s; d = 10 m.
(e) Silt particles settling in a still lake: ikI = 1.6 x 10-3 m/s; d = 4 x 10-5 M.
Fig. 14.2.1. Force diagram for a floating chunk of ice. B denotes the buoyant
force and W the weight force.
The story goes that Archimedes discovered this principle when observing
that the level in the municipal bath increased upon his entry. In his
excitement he ran naked through the streets shouting `Eureka' (which is
Greek for `I have found it').
It is not too difficult to understand the nature of Archimedes' principle.
Consider any undisturbed region of the fluid. There are two types of
force acting on this region : a gravitational force equal to the weight
of the particles, and the pressure force exerted at the boundary of the
region by the fluid particles outside this region. If the region of fluid
is to remain stationary, the pressure must exactly balance the weight
of the fluid particles. If a solid object displaces the fluid particles, the
same pressure forces act on the object's surface as once acted on the now
displaced fluid. Hence the object experiences a buoyant force equal in
magnitude, but opposite in direction, to the weight of fluid it displaced.
(a) (b)
Fig. 14.2.2. A hydrometer in (a) water, and (b) a liquid more dense than water.
Here BI> B'$ and x > 0.
which implies
m=mf,
where m denotes the mass of the chunk of ice.
STEP 4: Express the masses in terms of volumes by using the given densities:
In general
{mass} = {volume} x {density}.
Thus
m = 920 V, M f = 1, 000 Vs,
where V is the total volume, and VS is the submerged volume of the ice, and so
Vs/V = 0.92.
Hence 92% of the chunk of ice is below water level.
The hydrometer
A study of the previous example shows that the fraction of any particular
floating object above the surface level depends only on the density of
the fluid. This is the principle behind the hydrometer (see Figure 14.2.2)
which is used to measure the density of a liquid.
It is first necessary to mark on the stem of the hydrometer the surface
level of distilled water (density 1000 kg/m3 ), and to calculate the sub-
merged volume, Vs°) say. The hydrometer is then floated in the liquid
whose density is to be determined, and the displacement x of the surface
level, which is measured as the distance below the mark for distilled
water, is noted. If the stem of the hydrometer has cross-sectional area A,
14.2 Archimedes' Principle 285
Exercises 14.2
1. There's a well known saying: `That's just the tip of the iceberg'. Given that
the density of ice is 920 kg/m3 and the density of seawater is 1020 kg/m3, use
Archimedes' principle to calculate the fraction of the total volume of an iceberg
which is submerged under water.
2. A block of wood with dimensions 2 x 5 x 30 cm and density 400 kg/m3 is
held beneath the surface of a bucket of water by a straight string. Calculate the
tension in the string.
3. A 70 kg pig is marooned on a wooden board which is floating down a flooded
river. If the board is 15 cm thick and 2 m long and has a density of 600 kg/m3,
what is the width of the board if the top surface is level with the water?
4. Use Archimedes' principle to show that, for the hydrometer of Figure 14.2.2,
_ 1000VS°)
P1 VS°) - xA
where the symbols are defined in the text.
5. A `striped' cocktail is a drink made of three different liqueurs, which lie in
layers one on top of the other. The colours and densities of these liqueurs are
shown in the diagram below. The densities are in gram/ml, and each layer is
1 cm thick.
A small plastic cube of volume 1 cm3 and density 1.130 g/ml is carefully
lowered into the drink.
(a) Between which layers does it float?
(b) Show that 3/5 of its volume is in one layer and 2/5 in another.
6. The molecular weight of helium is 4.00 gram/mole. The molecular weight of
air is 28.9 gram/mole. A helium filled balloon rises because it displaces air in
excess of its own weight. The molar volume is approximately the same for all
gases: 0.0224 mole/m3 at a temperature of 0 °C and a pressure of 1 atm. Estimate
the volume of helium required to lift a mass of 100 kg. What is the required
radius of the balloon?
286 Motion in a fluid medium
tB FDt
FDj+w
Moving upwards Moving downwards
IFDI = 6itnr15cI.
When the sphere is moving towards the ground, .z is negative and, from
Figure 14.3.1, FD is positive. Thus
Example 1. Obtain a first-order equation for the velocity. Calculate the terminal
velocity.
Solution. Writing
v = .z (and thus v = z)
in the differential equation, and then slightly rearranging the result gives
(m-mf)g.
m m
This equation is a first-order linear inhomogeneous differential equation with con-
stant coefficients of the type studied in Chapter 11. The terminal velocity vt occurs
when v = 0. The differential equation is then simply solved for y to give
v = -(m-m1)g
` 6icrir
as the required terminal velocity.
r= t
2(P - Pf )8 /
where p denotes the density of the oil. As the characteristic time is very
small, the oil droplet can be assumed to be falling at the constant velocity
4i°). Thus, by measuring the time t it takes the droplet to fall a distance
e', the terminal velocity is determined by the simple formula
1vi°) ( = e/ t.
Suppose that now the electric field is switched on. Since the oil droplets
have acquired a charge of a few excess electrons when they were formed
by being sprayed from an atomiser, they feel an electric force FE. If
the top plate is kept at potential V, the bottom plate at zero potential
(i.e. earthed) and the oil droplet has excess charge q = ne, the laws of
electrostatics give
V
FE = nIeI d,
where d is the plate separation and the upwards direction has been taken
as positive (see Figure 14.3.2).
When the oil drop is moving, a drag force obeying Stokes' law will
act in addition to the forces in Figure 14.3.2. The equation of motion is
therefore
V
nfef -(m - mf)g - 6ngrac = mx.
d
14.3 Falling sphere with Stokes' resistance 289
Fig. 14.3.2. Forces on a charged oil drop at rest in an electric field created by the
Millikan apparatus.
Putting z = 0 and solving for z gives the terminal velocity vt, say, as
n1el
v= V + v(o)
6iuird
By measuring vt for a positive and negative voltage of the same magni-
tude, and subtracting, this formula gives a value
nIeIV
3irrird'
It is possible to vary the charge niel on the oil droplet by using radiation.
The above formula predicts that the resulting differences in upward and
downward terminal velocities will always be an integer multiple of some
constant. This was found by Millikan, and the value of the constant
accurately gave the magnitude of the electronic charge. Further details
on the Millikan oil drop experiment can be found in Melissinos (1968).
Exercises 14.3
1. It was shown in the text ttiat the equation of motion for a falling sphere
subject to Stokes' resistance is
Y--6ngry (m - mf)g
m m
(a) Identify the symbols in this equation.
(b) Find the general solution of the homogenized form of this equation.
(c) Find a particular solution and hence determine the general solution of
the original equation.
(a) Give some possible reasons why e°) is not precisely the same in each
case. What is the average value of &0.
(b) Use the average value of t(°) to calculate the radius of the drop r.
(Data: , = 1.60 x 10-5 kg/m3, c = 6.1 x 10-3 m, p = 883 kg/m3, pf =
1.29 kg/m3.)
(c) Use an appropriate formula from the text to calculate n$e$ for each line
in the above table. (Data: V = 500 volts.) You should find that, to
a good approximation, each of the values is an integer multiple of the
smallest such value.
(d) From your answer to part (c) calculate the approximate value of the
magnitude of the charge of the electron jel. The unit of charge in the
MKS system is the Coulomb.
for a sphere of radius r. The right hand side of this expression is always
positive, irrespective of the sign of the velocity z, whereas the direction of
the drag force FD must always be opposite the direction of motion (recall
Figure 14.3.1). The equation of motion can be written down correctly if
this is kept in mind. The solution of the equation of motion requires the
technique of separation of variables, discussed in Section 11.1.
V = 3 nr3
and
{mass} = {density} x {volume},
292 Motion in a fluid medium
v-axis (m/s)
4A
VtI -------------
30
20
10
I I I
t-axis (s)
1 2 3 4 5
Fig. 14.4.2. The velocity-time graph for the falling sphere of Example 1.
we see
A = 0.411 (&)
rrp
and 8= Vi_&\g].
STEP 4: Solve the differential equation using the technique of separation of
variables. It is to be shown in the exercises that this method gives for the solution
e2ABt -1
z=AB e2ABt+1
2AB
for a characteristic time. In this model, if t = r, the sphere has reached
approximately 46% of its terminal velocity. For times t > 8i, the sphere
has reached more than 99.9% of its terminal velocity.
Exercises 14.4
1. Obtain the Reynolds' number for the wooden sphere of Example 1 falling at
terminal velocity. Does the value obtained lie in the range appropriate for the
choice of drag constant CD = 0.45 (recall Figure 14.1.2)?
2. Consider the differential equation
v = -A 2V2 + B2,
with initial condition v = 0 at t = 0.
(a) State why 0 < v < B/A for t > 0.
(b) What are the dimensions of A and B, given that [v] = LT-! ?
(c) Obtain the partial fractions expansion
1 _ 1 1 1
v = B tan(C - ABt).
A
(e) How long is the ball moving upwards? For which values of t is the
solution in (d) valid?
295
296 Damped and forced oscillations
look for solutions of the form
x = eAt.
Substitution into (1) gives
(a22 + b2 + c)e2t = 0.
Since ell is never zero, this implies 2 must satisfy the quadratic equation
a22+b2+c=0. (2)
For example,
(ezt) = Aext
do
(ii) Complex conjugate roots. When the roots are complex we first get
complex solutions, using the complex exponential; real solutions can
then be obtained from the complex ones. The following example shows
how.
298 Damped and forced oscillations
(iii) Equal roots. When the roots of the characteristic equation are the
same then we only obtain one solution of the form
e2t
41(t) =
for the equation (1). It can be verified that another solution, linearly
independent from the first, can be obtained by multiplying the first
solution 01(t) by t :
02(t) = test.
Solution.
STEP 1: Consider the complex equation, with sin 2t replaced by e2U,
x-5c+6x = e2". (5)
a=2-2i=4 1 + i
1 1 1 n
W2
Hence a particular complex solution to (5) is
In (5), let x = xre + ixim where xre and xim are the real and imaginary parts of
x respectively. By equating the real and imaginary parts of (5), the two equations
xre - xre + 6xie = cos 2t
and
zim - xim + 6xim = Sin 2t
are then obtained. It thus follows that the imaginary part of ap(t) satisfies the
original equation (4).
STEP 3: Choose the imaginary part of
n
OP( t) 1 ( cos(2t + 7c4 ) + i s in (2t + 4
2,vF2
real solution
The general proof of this result is left to the exercises.
Exercises 15.1
1. Find the real and imaginary parts of the following expressions involving
complex numbers.
(a) i(3i + 1)
(b) (2i + 1)/(-2i + 1)
(c) (4i + 1)/i
1 5.1 Constant-coefficient differential equations 301
Dashpot
m
d00000000000000000000C
Equation of motion
The equation of motion for a particle on an ideally damped spring can
be determined by modifying the procedure of Section 6.2 for the case
without damping.
Example 1. Obtain and classify the equation of motion for the damped spring
system of Figure 15.2.2. The table is assumed smooth, the spring light and the
damping ideal. The natural length of the spring is ? metre and its spring constant is
k newton/metre. The particle has mass m kg. Suppose the coordinate x is measured
as the extension of the spring beyond its natural length.
Solution. In the horizontal direction there are two forces acting on the particle: Fs
due to the spring, and FD due to the dashpot. An application of Hooke's law, as
detailed in Section 6.2, gives
Fs = -kx.
The formula (1) gives the magnitude of FD. Its sign is determined from the criteria
that the damping force is always opposite in direction to the velocity. Hence
FD -y.z.
Newton's second law applied to the particle says
FD -I- Fs = mz
and thus
-y.z - kx = mz.
Rearrangement gives
z-. + -YX+kx=O,
(2)
m m
which is a homogeneous linear second-order differential equation with constant co-
efficients.
304 Damped and forced oscillations
Solution of the equation of motion
Let us suppose the spring in Example 1 is initially stretched a distance
xO metres and released from rest. It has already been remarked that
the expected consequence of a damping force in a spring system is the
gradual diminishing of the amplitude of the oscillations of the particle
about the equilibrium point. Does the solution of the differential equation
(2) make this prediction?
The solution of (2) can be obtained using the method of Section 15.1
for homogeneous equations. Recall that the first step is to try a solution
of the form e-11 and hence obtain the characteristic equation
YA
22
m
k
m
=0.
It is found that as the parameters k, y and m vary, three distinct types of
solutions are possible for the damped spring system, corresponding to the
nature of the roots of the characteristic equation. These cases are termed
underdamped, overdamped and critically damped harmonic motion. The
underdamped motion corresponds to the gradual diminishing of the
amplitude mentioned above. Each of the three cases will now be discussed
separately. Some of the mathematical details are left for the exercises.
Underdamped motion
It is to be shown in Exercise 15.2.1 that for y < 2(mk)l the solution of
(2) with the initial conditions x = xO, and is = 0, at t = 0 is
and
4
C1-1 y2
.
0.5
-0.5
1 2 3 4 5 6
From the solution (3), after some algebra, this can be written in terms of
y, k and m as
0=2n(4km-1 1 -1/2 .
\Y
A plot of A as a function of the dimensionless parameter y(km)-1/2 is
given in Figure 15.2.4.
For fixed values of k and m we see that the ratio of the first maximum
to the second maximum increases as the damping constant y increases,
and becomes infinite as the value y = 2(km) ' is reached. For y larger
than this critical value the motion is no longer underdamped.
Overdamped motion
If the damping force is sufficiently strong, it might be expected that
a particle on an extended spring, when released, will return to the
306 Damped and forced oscillations
2 y(km)-in
y > 2(mk)I.
The characteristic equation then has real, negative and distinct roots
which we denote by 2 = -A1 and 2 = -22. The general solution is of the
form
e_2, t + c2e-a,2t
x = cl
Although x is never exactly zero for any positive time t, it will approach
zero in the limit as t approaches infinity. In practice therefore, the
distance of the particle from the origin will become imperceptible after
a finite time.
Plots of (4) corresponding to the initial condition x = 1 and x = 0,
at t = 0, for various values of the dimensionless damping parameter
y(mk)-1/2, are given in Figure 15.2.5. A feature of these plots is that for
fixed values of m and k, the greater the value of the damping constant y,
the slower the return to the equilibrium position x = 0.
15.2 Damped oscillations 307
x
1
0.9-
0.8-
0.7-
0.6-
0.5
0.4
0.3
02
.
y(mk)-1n
0.1 =2
0 T r 1 I J- T T -1t
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
Once again, although x is never exactly zero for any t > 0, it approaches
zero as the time t approaches infinity. Hence the distance of the particle
from the origin will be imperceptible after some finite time. In most
applications of overdamping, it is desired to return to the equilibrium
point as fast as possible. The damping constant is thus chosen to have
the critical value (5).
308 Damped and forced oscillations
Fig. 15.2.6. Essential features of the artillery gun. The shaded region is the barrel.
Example 2. The barrel of an artillery gun (see Figure 15.2.6) weighs 500 kg and
has a recoil spring of stiffness 40000 N/m. If the gun is fired horizontally and the
barrel recoils 1 m, determine the critical damping coefficient of an attached dashpot
with ideal damping.
Solution. Critical damping occurs when y = 2(mk), . Substituting for m and k gives
y = 4J x 103 kg m/s.
as the required damping constant.
Example 3. For the problem in Example 2 find the time required for the barrel to
return to a position 5 cm from its initial position.
Solution. To answer this question, it is necessary to set up and solve the appropriate
equation of motion. Let if metres denote the natural length of the spring, and let
the displacement x metres be the extension of the spring as measured from C. Since
the displacement has reached a minimum at x = -1 The initial conditions are
x=-1, $=0, at t=0.
It is desired to calculate t such that
x = -0.05.
With this choice of coordinate system, as explained in Example 1, the equation of
motion is given by the differential equation (2). Substituting the numerical values
of y, m and k gives
x+8Jx+80=0.
The corresponding characteristic equation has a double root at --4J. The general
solution is thus
x= c2te--*%`, c1, c2 E R.
Substituting both initial conditions gives
-1 = c1
and
0=-4Jc1+c2.
Hence cl = -1 and c2 = -4J so that the solution of the equation of motion and
the initial condition is
x = -e-4j t - 4 Jtefs`.
15.2 Damped oscillations 309
From this expression for the displacement, it follows that x = -0.05 when
1 + 4Jt.
It is not possible to solve this equation using only algebra. An approximate solution
can be obtained, however, by plotting both sides of the equation and reading off the
point at which the graphs intersect (see Figure 15.2.7). It is thus found that the
required time is
t = 0.53 seconds.
Exercises 15.2
1. The differential equation
x+ myz+ kx=0
m
describes ideal damped harmonic motion.
(a) Identify the symbols in this equation, and indicate the choice of coordi-
nate system on a diagram.
(b) Use the theory of Section 15.1 to show that there are three distinct types
of solutions depending on whether
y < 2(mk )' , y = 2(mk) I or y > 2(mk) i ,
and obtain the general solution in each case. When the roots are complex,
write the solution in terms of the exponential and cosine function only.
(c) In each of these cases, determine the arbitrary constants in the general
solution so that the initial conditions x = xO and ac = 0 when t = 0 are
satisfied. Compare your answers with (3), (4) and (6) in the text.
310 Damped and forced oscillations
(d) Give the name of the type of motion in each case.
2. Consider the spring system of Figure 15.2.2. The table is assumed smooth,
the spring light and the damping ideal. Suppose the natural length of the spring
is 0.5 m and its stiffness is 10 N/m. Let the mass of the attached particle be 5 kg
and the damping constant 10 kg/s.
(a) If the displacement x metres is measured as the extension of the mass
from its equilibrium point, write down the equation of motion.
(b) With the initial conditions
x=0.1 and is=0.2, at t=0,
obtain the solution of the equation of motion in terms of the exponential
and cosine functions only.
(c) Suppose instead that the displacement x metres is measured from the
wall. Show that the equation of motion in this case is
5z = -10ic + 10(0.5 - x).
Use the workings of (b) to write down the general solution of the
homogenized version of this equation. Find a particular solution and
thus obtain the solution of the equation of motion when the initial
conditions are as in (b).
(d) Compare your answers obtained in parts (b) and (c) and comment. What
is the time interval between any two successive maxima?
3. (a) Prove that for ideal underdamped harmonic motion, the ratio of the
height of any two successive maxima is a constant. Use this fact to show
that the logarithmic decrement A can be written as
1 xi
A= In
n
x'
Fig. 15.3.1. The spring system and coordinates for Example I.
Equation of motion
An equation of motion with a forcing term of the form (1) can be
obtained in a spring system by periodically varying the position of the
attachment of the spring to the wall.
Example 1. Suppose a light spring and ideal dashpot are attached at one end to
a particle of mass M which moves on a smooth table. The other end of the spring
and dashpot are forced to oscillate with displacement given by
y = a sin cwt
as shown in Figure 15.3.1.
If x is the displacement of the particle as measured from the point with coordinate
t (the natural length of the spring), and z = x - y, show that
z+ j±+wz =Fosincwt (2)
where
k
Fo=a(cw)2, w2= M
and k is the stiffness of the spring and y is the damping constant.
Solution. Let the direction to the right of the origin in Figure 15.3.1 be positive.
There are two forces acting on the particle: Fs due to the spring, and FD due to the
dashpot. To calculate Fs, note from Figure 15.3.1 that the extension of the spring
beyond its natural length is x - y, assuming x > y. Hence, from Hooke's law, since
the positive direction is to the right,
Fs = -k(x - y).
This formula holds if y > x also, since then the spring is compressed so the force
is positive.
The force due to an ideal dashpot is directly proportional to the difference be-
tween the velocity of the two ends of the dashpot. Since the rightmost end is moving
at velocity z, and the leftmost end at velocity jy, this difference is x - Y. If x > jy,
the force is in the opposite direction to the particle motion, so
FD = -Y(x - Y)
This formula also applies if ac < y.
15.3 Forced harmonic motion 313
Since the origin for the x-coordinate is fixed, the acceleration of the particle is
x. Hence, by Newton's second law,
Mx=Fs+FD-
Changing from x to the relative coordinate z = x - y gives
Fs = -kz, FD =-y2 and z=z+y.
Hence the equation of motion M(z + y) = Fs + FD becomes
Mz=-kz-y2-My
and so
Mz = -kz - y2 + Ma(w1)2 sin w ft,
since y = a sin(w f t). Hence
z+ y 2+
M
kM z= Fo sin(wf t) (3)
as required.
z -. + - y ac + k x = F° sin(cot
f).
M M
This equation is exactly the same as (3), but with x in place of z.
is the complex solution of (5). To get the real solution of (5) we need to take the
imaginary part of (5). (Note that if the right-hand side of (4) had been cos(coft)
then we would then need to take the real part of the particular solution (6).)
However we are only interested here in the amplitude A of the real solution to
(4). This can be obtained (see the discussion at the end of Section 15.1) directly
from the complex solution (6) as A = jai. Hence the required amplitude A is
1 1
A= - (c)f)2 + O.lic) 19
- (c)f)2 + 0.1ic)f
Hence
-1/2 .
A = ((9 - w2)2 + O.O1102)
A plot of the amplitude A against the forcing frequency coif is given in Figure 15.3.2.
Figure 15.3.2 shows that the amplitude of the oscillations of the particu-
lar solution peaks very close to cof = 3 (i.e. the frequency cof/2ir = 3/27x).
The value 3/27r is the frequency of the system with damping and forcing
removed that is, the natural frequency of the system. The frequency
cof/27r at which this peak occurs is called the resonant frequency of the
system. In general, for small damping, when the frequency of the forcing
term approximately equals the natural frequency of the spring system,
the amplitude of the non-transient solution peaks. The phenomenon is
known as resonance.
Resonance is responsible for some surprising happenings. A well-
known example is a singer holding a note at a certain frequency and
shattering a glass. This occurs when the motion of the natural frequency
15.3 Forced harmonic motion 315
Exercises 15.3
2. (a) Recall that the natural frequency of a system is the frequency of the
oscillations when the damping force and the external force are removed.
What is the natural frequency of the system in Exercise I?
(b) What is the resonant frequency for the system in Exercise I?
316 Damped and forced oscillations
3. Repeat Exercise 1 for the equation
z + 0.051 + 4z = cos cwt,
here sketching the amplitude in the range 0 <_ c of <_ 4.
4. Suppose a particle rests on a light spring. If the weight of the particle causes
the spring to be compressed a distance of 5 cm, use Newton's and Hooke's laws
to deduce the ratio of the mass of the particle to the spring constant, and thus
determine the resonant frequency, assuming small damping.
5. Consider a particle resting on a smooth table and attached to a spring of
stiffness 10 N/m and an ideal dashpot. The spring and dashpot are attached to
a wall.
(a) Suppose the spring is extended and released. You are given that the
time interval between successive maxima is 1.5 s and that the ratio of
the amplitude of successive maxima is 2/7. Use appropriate formulae in
the text to calculate the mass of the particle and the damping constant.
(b) Determine the amplitude and phase of the non-transient part of the
motion if an external force Ff(t) = 2 sin(4t) acts on the system.
6. The diagram below shows the essential elements of a vibration-measuring
device (seismometer or accelerometer). If in the y-direction there is a displacement
of the form
y = Y sin(wft)
(due to an earthquake for example), the aim is to deduce Y from the recorded
amplitude of the relative displacement
mg
z = x-Y+ k
of the particle.
T
A -----------------------------
(a) By following the working of Example 1, show that the equation of
motion for the particle can be written
mi + yi + kz = mYwf sin(wft) - mg.
(b) Show that the amplitude A of the non-transient part of the solution to
this equation is
Y(wf/(0)2
A
[1 - (2Cwf/(o)
where co = (k/m) 7 and 2C = y/(mk) 2 .
15.3 Forced harmonic motion 317
318
16.1 Kinematics in a plane 319
y-axis
y ---------(x,y)
yj
x-axis
X
X = xi -}- A (1)
where j and j are the vectors of unit length pointing in the directions of
the x- and y-axes respectively. Note that the cartesian coordinates of a
point determine its position vector uniquely, and conversely.
Curves
As the time varies, the particle traces out a curve in the (x, y)-plane, as
illustrated in Figure 16.1.2.
The coordinates of the particle are thus functions of time given by,
say,
x = 4(t) and y = ip(t) (2)
x-axis
The following examples show how the curve on which the particle is
moving may be obtained once these functions are known.
Solution. Eliminating t from these two equations gives the single equation relating
x and y,
y=x2.
This is the equation of a parabola in the (x, y)-plane, as shown in Figure 16.1.3.
Since x increases as t increases, the particle moves along the parabola in the direc-
tion shown by the arrow.
16.1 Kinematics in a plane 321
y-axis
. x-axis
Particle
Fig. 16.1.4. Particle moving along a straight line in the (x, y)-plane.
Example 2. Describe the curve along which the particle moves in the (x, y)-plane
when its position vector is given as a function of time by
X = -ti + (2t + 1)j.
Solution. The cartesian coordinates x and y of the particle are just the respective
coefficients of i and j on the RHS of the formula for X. Hence
x=-tand y=2t+1.
Elimination oft gives
y=-2x+1
and so the curve along which the particle moves is a straight line, as shown in
Figure 16.1.4. The particle moves along the line in the direction indicated by the
arrow because the parametric equations show that x decreases as t increases.
322 Motion in a plane
Velocity and acceleration
Our approach to these ideas is motivated by Galileo's work on projec-
tiles. He imagined the motion in the vertical plane to consist of two
independent components : one vertical and the other horizontal. Ver-
tically the motion was as for free fall, while horizontally the motion
was with constant velocity. By arguing in this way, Galileo was able to
deduce the correct result that the path traced out by the projectile was a
parabola.
We now apply this idea quite generally to a particle moving in a
plane in such a way that its cartesian coordinates x and y are twice
differentiable functions of time. The velocity consists essentially of two
components : ac along the direction of the x-axis and y along the direction
of the y-axis. Just as we combined the two coordinates in equation (1)
to give a single position vector
X= xi+ A 1W
X=aci+.Yj (4)
X=zi+
00
(5)
1W
The formulae (4) and (5) may be regarded as the definitions of velocity
and acceleration for a particle moving in a plane, with position vector
X. given by (1), at time t. The speed of the particle is defined as the
magnitude of the velocity vector, is given by
IIXII = V-t,+y2
The advantage of combining the components to produce a single
vector will be apparent later when we give the geometrical interpretation
of the velocity vector. Meanwhile, the following example shows how
to calculate the velocity and acceleration vectors using familiar rules of
differentiation.
16.1 Kinematics in a plane 323
x-axis
Example 3. A particle moving in the (x, y)-plane has position vector at time t given
by
X = -ti + t2j.
Find the velocity and acceleration vectors and at time t. When t = -1, show
X, X and on the curve traced out by the particle. Find also the speed of the
particle at this time, if distance is in metres and time in seconds.
Solution. By our definitions, the velocity and acceleration are given as functions of
t by
-1+26,
X= 2j.
II X II =
VI"(- + (-2)2 = m/s
is the speed of the particle when t = -1.
AX
N
makes it seem plausible that this will always be the case : by definition
,..
bx by
= lim i lim
(it--+o b t + bt
has the same direction as the vector 5X. Hence Figure 16.1.6 makes it
plausible that
bX
lim =X
bt
has the same direction as the tangent to the curve along which the particles
moves, at the point with position vector X.
The notations
dX d2X
and
dt dt2
Exercises 16.1
1. In each case state what is wrong with the equation involving vectors :
(a) a= b+ c (b) a= b+ c
(c) a = b c (d) X = t + 2t3j.
2. Plot each of the following points in the (x, y)-plane and then express the
position vector of the point as a linear combination of the vectors i and j.
(a) (0,1) (b) (1, 0)
(c) (1, 1) (d) (cos(ir/6), sin(ir/6)).
3. In each case sketch the curve traced out in the (x, y)-plane by a particle when
its coordinates are given as functions of time by the stated formulae. Show the
direction of motion along the curve.
(a) x=t2 andy=t (b) x=t2 andy=t2
(c) x= 1 andy=t (d) x=et and y=e t.
4. In each case sketch the curve traced out in the (x, y)-plane by a particle when
its position vector is given as a function of time by the stated formula. Calculate
the velocity and acceleration vectors and the speed of the particle.
(a) X = 2ti + 2tj (b) X = -ti - t2j
(c) X=ti+etj (d) X = eti + e2tj.
7. A particle moves around a circle. Its position and velocity vectors at time
t are X and X. What is the angle between X and X? (A careful sketch should
reveal the answer!) What is the value of the dot product X X?
8. Suppose that a particle moves in such a way that its position vector at time t
is given by
X = eti+e-tj.
Calculate the velocity X and verify that it is perpendicular to X when t = 0.
^1 ^1
326 Motion in a plane
16.2 Motion down an inclined plane
Galileo tested his hypothesis that bodies fall towards the Earth with
uniform acceleration by rolling a brass ball down a plane inclined to the
horizontal at various angles. The inclined plane `diluted' the effect of
gravity and, by slowing the motion, made it possible to measure accu-
rately the times at which the balls passed through various points along its
path. Galileo found that, for a given inclination, the distance which a ball
rolled down the plane when released from rest was proportional to the
square of the time taken - just as would be expected if the acceleration
were uniform.
The aim here is to study motion down an inclined plane using Newton's
laws. To simplify our model we shall replace the brass ball by a particle
and shall ignore the effect of friction. The problems which arise in trying
to make a more realistic mathematical model of Galileo's experiment will
be discussed later.
To find the equations of motion for the particle, we shall perform the
following steps :
STEP 1: Choose a coordinate system for the vertical plane containing the
line along which the particle moves.
STEP 2: Express the forces acting on the particle as a linear combination
of the vectors i and j.
STEP 3: Apply Newton's second law in its vector form to get the equations
of motion of the particle.
Similar steps apply to other problems involving motion of a particle in a
plane.
Since we assume the particle stays on the plane, it follows that y = 0 throughout
the motion. Hence y = 0.
STEP 2: Write the force as a vector. Suppose the particle has mass m. There are
two forces which act on it: its weight, of magnitude mg, and the normal reaction
of the plane, of magnitude N, say. These forces are shown in Figure 16.2.2.
These forces are to be expressed in terms of the vectors i and j. Since j is the
^1 ^1
unit vector with the same direction as the normal reaction,
{normal reaction) = Nj.
To express the weight as a vector, we first find a unit vector pointing vertically
downwards. So consider the right angle triangle in Figure 16.2.3 with one side
parallel to the inclined plane and with hypotenuse of length 1.
The angle at the bottom of the triangle is 7r/2 -a; hence the angle at the top is
it/2 - (ir/2 - a) = a. Thus the remaining sides of the triangle have lengths sin(a)
328 Motion in a plane
and cos(a) as shown. Hence the unit vector in the vertically downwards direction is
sin(a)i - cos(a)j.
Hence, in vector form,
{weight} = mg(sin(a)i - cos(a)j).
The net force F acting on the particle is now given by
F = {normal reaction} + {weight}
= Nj + mg(sin(a)i - cos(a)j)
= mg sin(a)i + (N - mg cos(a))j
STEP 3: Apply Newton's second law. The vector form of this law, F = mX, now
shows that
mg sin(a)i + (N - mg cos(a))j = mzi + myj.
Equating coefficients of i and j on each side and using y = 0 gives
z = g sin(s), (1)
0 = N - mg cos(a). (2)
Equation (1) shows that the acceleration down the plane is the constant g sin(s).
Equation (2) shows that the normal reaction has magnitude mg cos(a).
Effect of friction
It is relatively easy to include in the above model the effect of friction
between the plane and the particle. (See Chapter 4 for the basic ideas of
friction.) Exercises to work out the details are set later.
Even when friction is included, however, the particle model is still
16.2 Motion down an inclined plane 329
Exercises 16.2
1. Recall from the text that, for a particle moving down a smooth plane inclined
at an angle a to the horizontal, the acceleration is given by the formula
x=gsin(a) 0<a<ir/2.
(a) Find the limit of the RHS as a -- 0.
(b) Find the limit of the RHS as a -- it/2.
(c) What are the physical interpretations of your answers to (a) and (b)?
(d) Suggest a further check which can be applied to the answer in the text.
2. Recall from the text that, for a particle moving down a smooth plane inclined
at an angle a to the horizontal, the normal reaction on the particle has magnitude
given by the formula
N = mg cos(a).
(a) Find the limit of the RHS as a -- 0.
(b) Find the limit of the RHS as a -- it/2.
(c) What are the physical interpretations of your answers to (a) and (b)?
7. Solve the differential equation in Exercise 6(a), assuming that the particle was
initially at rest.
8. Investigate how the results stated in Exercise 6 are affected if the particle has
been initially projected so as to slide up the plane.
9. Find the time taken by a skier to ski 1 km down a slope making an angle of
30° to the horizontal
16.3 Projectiles 331
16.3 Projectiles
In the study of projectiles, the aim is to set up a mathematical model for
the motion of a body projected near the Earth's surface, like a cricket
ball or an artillery shell. In the simple model investigated in this section,
the Earth is regarded as an inertial frame and hence the projectile is
assumed to move in the verticle plane containing its initial direction of
projection. While this simplification is appropriate for a cricket ball, it is
less so for a shell. The only force taken into account, moreover, is gravity
(assumed constant in magnitude and direction).
This model predicts that the path traced out by the particle (its
trajectory) is a parabola as was already known to Galileo, who
realized that the horizontal and vertical components of the motion could
be analysed separately. We achieve the same end by the use of vectors,
obtaining a pair of uncoupled simultaneous differential equations to
describe the motion.
If the model is modified to include the effect of air-resistance, the use of
vectors makes it easy to obtain the modified equations of motion, as will
be seen from a later exercise. The modified equations are coupled and,
while it is possible to uncouple them by a suitable change of coordinates,
the easiest way to get useful information about the behaviour of the
solutions is by approximate numerical techniques.
The steps to follow in deriving the equations of motion are much the
same as in the previous section.
I Weight
x-axis
STEP 3: Apply Newton's second law. The vector form of this law shows that
-mgjN = mX = mxi + myj.
Equating the coefficients of i and of j on each side gives, after cancellation of m,
x = 0, (1)
(2)
y = -g.
The simultaneous differential equations (1) and (2) are the desired equations of
motion.
The subsequent motion of the particle can be found from these dif-
ferential equations once the point from which the particle is projected
and the velocity of projection are specified. The initial velocity can be
obtained from the initial speed and angle of projection by elementary
trigonometry, as in the following example.
Example 2. Let the x-axis and the y-axis be horizontal and vertically upwards
respectively, as in the previous example. Suppose that a particle is projected up from
the origin at ground level with an initial speed of 10 m/s at an angle of n/4 radians
to the x-axis. Find the coordinates x and y of the particle as functions of the time
t after projection, and describe the trajectory of the particle in the (x, y)-plane.
(Assume x and y are in metres, t in seconds.)
Solution. To write the initial vector in terms of i and j, we first get a vector of unit
length in the direction of projection.
By the construction given in Figure 16.3.2, this unit vector is
n/4
. x-axis
cos(ir/4) cos(n/4) i
Equating coefficients of i and j in this vector equation and using the fact that the
particle starts at the origin gives
x=0 and
is5-,12
when t = 0. (3)
Y= 0 512-
Solving the differential equations (1) and (2) subject to the initial conditions (3)
gives the solutions
x=5/2-t,
(4)
1
y = _ gt2 + 512- t,
2
Exercises 16.3
1. Verify the claims made in the solution of Example 2 in the text concerning
the solutions of the simultaneous differential equations
x0
Y=-g
with the relevant initial conditions.
2. Use the solutions obtained in the text to Example 2 to determine:
(a) how long the particle takes to return to ground level,
(b) how far it had then moved horizontally,
(c) the maximum height reached by the particle.
3. Let the x- and y-axes be horizontal and vertically upwards respectively, with
the origin at ground level. Suppose a particle is projected from the origin in the
(x, y)-plane with the speed u > 0 at an angle a to the x-axis (0 < a < n/2).
(a) By following the method used in Example 2, show that the equation of
the trajectory of the particle is
g
y=- 2 x2 + tan(a)x (y >_ 0).
2 cos (a)u
(b) Verify that this equation is dimensionally correct.
4. In each of the following cases find the limiting form of the equation given in
Exercise 3 and describe the limiting trajectory :
(a) when u approaches oo
(b) when a approaches 0.
(c) Do your results seem plausible physically? Give reasons.
5. In Exercise 3, show that the particle returns to ground level at a point whose
distance from the origin is
sin(2a)u2/g.
16.3 Projectiles 335
Apply some checks to this answer. For which angle of projection is this distance
a maximum?
6. Show that, if 0 < f3 < a, the time t which elapses before the particle in
Exercise 3 again crosses through the line through 0 making an angle J3 with the
horizontal is given by
Zu
t= cos(a)(tan(a) - tan(J3)).
g
Apply some checks to this answer.
7. A projectile of mass m moves in a vertical plane subject to the forces of
gravity and air-resistance. The magnitude of the air-resistance is assumed to be
proportional to the square of the speed of the particle. The direction in which it
acts is opposite to the velocity vector. Suppose that the Cartesian coordinates of
the particle at time t are x and y where the x-axis is horizontal and the y-axis is
vertically upwards.
(a) In terms of ac and y find
(i) the magnitude of the velocity vector zi + yj,
(ii) a vector of unit length with the direction of
(iii) the speed of the projectile,
(iv) the force due to air-resistance.
(b) Deduce that the equations of motion of the projectile are
mx = -kac .z2 + y2,
my= -g - ky .z2+y2,
for some constant k > 0.
[These differential equations can be uncoupled, and hence solved in
closed form (see Synge and Griffiths, 1959); however, the resulting
formulae are complicated. Alternatively, numerical approximations to
the solutions can be obtained, on a computer, and hence graphs can be
sketched.]
17
Motion on a circle
This chapter explains the mathematical ideas needed to study the motion
of a particle moving on a circle. These ideas enable us to formulate
mathematical models for a number of interesting problems. For example,
as an introduction to the mechanics of the solar system you are shown
how to derive one of Kepler's laws for the special case of a planet which
moves in a circular orbit. In another application you are shown how
to set up the equation of motion for a pendulum, from which useful
information can be obtained by linearizing the differential equation.
336
17.1 Kinematics on a circle 337
y-axis
x-axis
x-axis
reaches (-a, 0). After this, 9 drops to values near -n and then increases
up to 0 as the particle passes through (a, 0) once more. Note the sudden
jump, or discontinuity, in the values of 9 as the particle passes through
(-a, 0).
338 Motion on a circle
is called the average angular velocity of the particle during this interval.
The angular velocity at time t is defined to be
limber fl(t)
and we denote it by 9. Thus the angular velocity is the rate of increase
of the angular coordinate with respect to time. It measures the rate at
which the particle is spinning around the origin. If the angular velocity
is positive, the particle is moving anti-clockwise around the circle; if it is
negative, however, the particle is moving clockwise. If a is constant, the
motion of the particle around the circle is said to be uniform.
In a similar way, the ratio
69 f (t + bt) - f (t)
bt bt
is called the average angular acceleration of the particle during the time
17.1 Kinematics on a circle 339
y-axis
a sin($) j
x-axis c
a cos(O) i
Cartesian coordinates
As can be seen from Figure 17.1.4, the cartesian coordinates of the
particle are given in terms of its angular coordinate by the formulae
x=acos9, y=asin9.
Hence the position vector of the particle is given in terms of its angular
coordinate by the formula
while in equation (3) expand and then interchange the second and third
terms to get
X = -a cos(e)e2i + -a sin(9)92jN - a sin(O)81 + a cos(O)BjN
and hence
X = -a92 (cos(O)i + sin(O)j) + ae (- sin(e)i + cosce . (5)
vectors
have length 1 and, at the point on the circle with angular coordinate 0,
v is normal to the circle, pointing outwards from the circle, while
i is tangent to the circle and points anti-clockwise.
and
X = -a62v + aei. (8)
Proof The vectors v and have length 1; this follows directly from (6).
i
The direction of v is the same as that of X, as comparison of (1) and (6)
shows. Thus v points out radially from the circle.
On the other hand it follows directly from (6) that the dot product of
v and is zero. Hence i is perpendicular to v and so must be tangent to
the circle.
Finally, the formulae (7) and (8) follow immediately from (4) and (5)
on use of (6). D
The geometric significance of the formulae (7) and (8) for the velocity
and acceleration is now clear. Since v and i are the unit vectors normal
1%0 1%0
and tangent to the circle, these formulae express the velocity and accel-
eration in terms of normal and tangential components, as illustrated in
Figure 17.1.6.
Thus (7) shows that the velocity is in the direction of the tangent to
342 Motion on a circle
Fig. 17.1.6. Velocity and acceleration in terms of tangent and normal vectors.
the circle and has magnitude I a6 I = a161 which is also called the speed of
the particle.
On the other hand, (8) shows that the acceleration has both a tangen-
tial component of magnitude I a6 I = a101 and a normal component, of
magnitude a62, directed towards the centre of the circle.
The formulae (7) and (8) are the key to solving problems involving
motion in a circle and they should be remembered.
Exercises 17.1
1. Each of the following points lies on the circle of radius 2 and centre (0, 0) in
the (x, y)-plane:
(0, -2), (290)9 (092)9 (-1, N/3-), (-290).
Plot each of these points on the circle and state the value of its angular coordinate.
2. A particle moves around the circle in Exercise 1 in such a way that its position
vector at time t is given by the formula
X = 2 cos(t)i + 2 sin(t)j.
(a) Express the angular coordinate 0 as a function of t when -n < t <_ n.
Hence find the values of the angular velocity 0 and angular acceleration
0 at time t. Find also the speed of the particle.
(b) Now use appropriate formulae in the text to express the velocity and
the acceleration vector X as functions of t for -n < t < n.
(c) State briefly why the formulae obtained in (b) are still valid when
n<t<_3n.
3. By using the formulae given in the text for the unit tangent and unit normal
vectors, verify that they each have length I.
4. By using the formulae given in the text for the unit tangent and unit normal
17.2 Uniform circular motion 343
vectors i and v at a point on the circle, verify that their dot product i v is zero.
To which geometrical fact does this correspond?
5. The bob P of a simple pendulum oscillates in a vertical plane along the arc
of a circle between two points A and B (at the same height). Let 0 be the angle
which the pendulum makes with the vertical at time t, as shown below.
A0
(a) Which point does the particle P occupy when its angular coordinate 0
assumes (i) its maximum value? (ii) its minimum value?
(b) What is the value of the angular velocity 9 at these points?
(c) Show the direction of the acceleration vector X at each of these points.
6. In Exercise 5, which point does the particle P occupy when its angular
acceleration is zero? Show the direction of its acceleration vector X at this point.
7. Complete the details given in the text of the derivation of formula (3) for X
from formula (2) for X.
This shows that, when circular motion is uniform, the acceleration always
points in the direction from the particle to the origin, as shown in Figure
17.2.1.
We assume that the particle starts on the x-axis so that 0 = 0 when
t = 0. The differential equation (1) together with this initial condition
has the solution
0 = wt (3)
provided t is in the interval for which -n < 0 < it. If t lies outside this
interval, then a multiple of 2n must be added to the RHS of (3). Hence
the formula
X = a cos(0)i + a sin(0)jN
Example 1. The Earth moves around the sun in an orbit which is nearly circular
of radius 1.49 x 108 km. Find
(a) the angular velocity of the Earth around the sun,
(b) the speed of the Earth,
(c) the magnitude of its acceleration towards the sun.
Solution. We choose the centre of the sun as the origin and let
position vector
X= of the Earth
at time t
outward unit normal
v= to the Earth's orbit
1 at this point
angular velocity
(00 = of the Earth
around the sun
speed
u= of the
Earth
(a) The position vector of the Earth sweeps out an angle of 2ir radians in 1
siderial year, of length 365 a days. Thus the angular velocity CO is given by
X50 metres X
w
Centre of circle
Example 2. A car is crossing a bridge whose vertical cross-section has the form of
an arc of a circle of radius 50 metres. Show that, if the car is to maintain contact
with the bridge at the highest point, then its speed u must satisfy the inequality
u s 22 m/s.
Solution. We regard the car as a particle. The solution follows the usual steps for
a dynamical problem: introduce notation; draw a force diagram; apply Newton's
second law.
We choose the centre of the circle as origin and let
X = {position vector of car at time t} ,
w = {angular velocity of car around origin} ,
V = {outward unit normal at top of circle),
i = {unit tangent at top of circle),
N = {magnitude of normal reaction on car},
f = (magnitude of friction force on car) ,
m = (mass of car).
We need only consider the forces acting on the car when it is at the highest point
of the road. The directions of the individual forces are then either horizontal or
vertical, as shown in Figure 17.2.3. Thus the net force acting on the car, at the top
of the road, is given by
F=(N-mg)v- fT.
17.2 Uniform circular motion 347
Now, while the car remains in contact with the road, it is moving around a circle
of radius a = 50 metres. Hence the formula (2) gives
X = -acw2v
where a = 50 metres. Application of Newton's second law F = mX now gives
(N - mg)v - f r = -maw2v.
Equating coefficients on each side of this equation gives
N - mg = -maw2
and hence
acw2 = g - N/m.
But, since N >_ 0, this implies that
acw2 < g.
Hence by (5)
u2 <_ ag,
and sou< V36-
It can be shown that, for the car to maintain contact with the road at
points near the top of the bridge, an even stricter speed limit must be
applied than that obtained in the above example.
Exercises 17.2
1. The moon orbits the Earth once every 27.3 days in a nearly circular orbit of
radius 3.83 x 108 metres. Find the acceleration of the moon.
2. A car of mass 1000 is travelling at a constant speed of 20 m/s as it moves
down one hill and up the next as in the diagram below. Near the lowest point,
the vertical cross-section of the road may be regarded as part of the arc of a
circle of radius 200 metres.
9Or1\_-z_1
0
348 Motion on a circle
Find (i) the normal reaction of the road on the car and (ii) the net frictional
force acting on the tyres as the car passes over the lowest point on the road.
3. The breaking strength of a 50 cm long string of a simple pendulum is 30
newton. What is the maximum mass that may be used for the bob of the
pendulum if the speed of the bob at the lowest point of the string is 1 m/s.
4. An unbanked curve on a highway has the shape of a circular arc. What is
the minimum safe radius for the arc if the coefficient of static friction between
the tyres and the road is 0.6 and the speed limit is 100 km/h?
5. Assume that a planet moves around the sun of mass M in a circular orbit of
radius r, with angular velocity to.
(a) Introduce notation for the position vector of the planet at time t and
the unit normal to the circle at the point with this position vector.
(b) From Newton's law of universal gravitation, write down the force acting
on the planet, at time t.
(c) Apply Newton's second law of motion to derive the formula
t2 = GM/r3.
Hence obtain Kepler's third law.
The model
Our model for the pendulum consists of a light rod of length I pivoted
smoothly at one end and with a particle of mass m attached to the other
end. The particle is called the bob of the pendulum. Motion is possible
in a vertical plane. We choose 0 to be the angle which the pendulum
makes with the vertical at time t, as shown in Figure 17.3.1.
Thus the bob of the pendulum is free to move in a vertical circle under
gravity. The bob is held on the circle by the force due to the rod, which
is assumed to act along the direction of the rod.
8=0
Linearization near 0 = 0
Huygens knew that during small oscillations the values of 0 stay close
to 0. So instead of using the actual RHS of (1), given by
RHS = - g sin 0,
he looked for a simpler RHS which approximates this closely when 0
is small. As we see from the graph in Figure 17.3.3, the tangent at the
point (0,0) follows the graph closely when 0 is small. So following the
tangent rather than the original graph gives a good approximation and
it has the advantage of being linear in 0.
Now the tangent to the graph is the straight line through the origin
with the same slope as the graph at (0, 0). We use differentiation to get
the slope. Thus the new RHS is given by
RHS = - g sin'(0)0,
_ - g cos(0)9,
__90.
e
352 Motion on a circle
In the differential equation (1) we now substitute this new RHS in place
of the old one. This gives the new, and simpler, differential equation
00
0=-eO. (2)
T= 2rze , (4)
V8
which is independent of the amplitude A.
Exercises 17.3
1. Show that 0 = 0 is an equilibrium point of the differential equation
00
0 = 0 -- sin(20)
and then linearise the differential equation near this equilibrium point.
Part five
Coupled Models
18
Models with linear interactions
355
356 Models with linear interactions
Dye - - Water
Assume that initially the two vats each contain 100 litres of pure wa-
ter. Pure dye is then pumped into the first vat at a fixed rate of 1
litre/minute, while pure water is pumped into the second vat at the
same rate. Pumps exchange the mixtures between the two vats - at
a rate of 4 litres/minute from vat 1 into vat 2, and 3 litres/minute
from vat 2 into vat 1. The diluted mixture is drawn off from vat 2 at a
rate of 2 litres/minute. Derive a pair of differential equations for the
concentrations of dye in each vat.
The steps involved in the solution follow closely those used in the
example in Section 13.1. First we draw a diagram and show the rates of
flow of the liquids into the vats, as in Figure 18.1.2.
We then note that, for vat 1, liquid (pure dye and mixture 2) flows in
at the rate of 4 litres/minute. This is exactly balanced by the outflow
18.1 Two-compartment mixing 357
Dye Water
1 Umin 1 Umin
i
:. Concentration is Concentration is
c1= x1/l00 c2 = X211 00
Vat 1 Vat 2
(2)
Here the first factor in each term is the volume of the appropriate mixture
flowing into or out of vat 1 during the time 8t. These factors are each
multiplied by the fraction of dye (concentration) in that mixture to give
the volume of the dye. The final result is only an approximation because
the concentrations change slightly during the time interval.
Similarly, 8x2 is the change in volume of the dye in vat 2 during this
time interval and hence
volume of dye volume of dye
8x2 = flowing - flowing
into vat 2 out of vat 2
(48t) x
i - (38t) x j-(2Ot)x ice (3)
As in Section 13.1, we can show that the errors involved in (2) and (3)
are small compared with 6 t. Hence dividing each of (2) and (3) by 8 t,
and then letting b t approach 0 gives
dx1=1-
jXl+X2
4
(4)
dx2_ 4 _ 5
1(-)()XI
dt
100x2'
Steady-state solutions
How to solve pairs of simultaneous differential equations like (4) and
(6) will be explained in the next section. One thing we can do straight
away, however, is to look for the constant solutions of the differential
equations (which correspond to steady-states or equilibrium states of the
mixing problem). These solutions can be found by setting
dci
=0 and dc2 =0
dt dt
in (6) to get the simultaneous pair of linear algebraic equations
--4ci + 3c2 = -1,
(8)
4ci - 5c2 = 0.
These equations have the solution
5 1
ci = and c2 =
5 2'
which are therefore the steady-state concentrations for the mixing prob-
lem.
The next question to ask is whether the actual concentrations approach
these steady-state concentrations as time goes on. We will be able to
answer this in the next section, by taking limits as t --+ oo, after we have
found the time-dependent solutions of the intial-value problem (6)-(7).
360 Models with linear interactions
Exercises 18.1
1. Initially two vats each contain 50 litres of pure water. Pure dye is then
pumped into the first vat at a fixed rate of 2 litres/minute, while pure water is
pumped into the second vat at a rate of 2 litres/minute. Pumps exchange the
mixtures between the two vats - at a rate of 6 litres/minute from vat 1 into vat
2, and 4 litres/minute from vat 2 into vat 1. The diluted mixture is drawn off
from vat 2 at a rate of 4 litres/minute.
Show that the concentrations of dye in the vats at time t minutes after the
start satisfy the following simultaneous differential equations:
dcl_
dt
1
25
- 3 2
25 cl + 25 c2'
dc2_ 3 4
T5 -C1 - T5-C2-
dt
What are the initial conditions for these differential equations?
2. Find the steady-state concentrations for the mixing problem in Exercise 1.
3. As in Exercise 1 except that vat 1 contains 50 litres of pure water initially
and vat 2 contains 100 litres of a 25% mixture of dye initially.
4. As in Exercise 1 except that dye is now pumped into the first vat at the
rate of 3 litres/minute. Show now that the concentrations satisfy the following
differential equations until such time as the first vat overflows :
cl = (3 - 7c1 +4C2)1(50 + t)
c2 = (6c1 - 8c2)/50.
100 + 100
cl (by (4)).
Thus
9 8 5
c'
100
c' - 1002 c' + 1002
or
.
c'+100 e'
9 8
c'__ 5
(6)
+1002 1002
which is a second-order linear constant-coefficient differential equation.
STEP 2: We solve (6) for cl as a function of t, by using the method from Chapter
15. Substituting
cl = eAt
into the homogenized version of (6) gives the characteristic equation
8
22+ 100
9 A+ =0
1002
which has the roots Al = -100 and 22 = -1L. Looking for a constant solution of
(6) itself gives cl = g . Hence, by the method given in Chapter 15, we obtain
sr 5
+ ate _ T +
ale- i r
ci = (7)
8
where ai and a2 are constants whose values are to be found from the initial condi-
tions.
STEP 3: We find c2 as a function of t, by substituting (7) back into (4). This
gives
100 4 _ 1
cl +
c2 = 3 100c' 100
100
3
8
(_jaieT& - 100 8 1 it + 4
100
t
+ (aiei&st + ate- T + 8 100
364 Models with linear interactions
Thus
4 _8t _I t
c2=-3aIeI +a2eTM+2. (8)
C2 =
14
It is easy to verify that cl and c2, given as functions of t by (9), satisfy both the
differential equations (4) and (5) and the initial conditions.
Exercises 18.2
1. In each of the following cases, state if the pair of simultaneous differential
equations is linear, constant-coefficient, or homogeneous.
(b) dx
(a) dx = 3x - 7y = 7x - 2y
dt dt
dy =x-ty2 dy =x-ty2
dt dt
dxl du _
(c) = 2x1 - 2X2 (d) 2u + v + cos(t)
dt dt
dx2 = 2x 1 - 2x2 dv
= v + 2u + sin(t)
dt dt
2. Use the procedure described in the text to find the solution of the pair of
simultaneous linear constant-coefficient differential equations
x=x+2y
y = 2x + y
18.2 Solving constant-coefficient equations 365
cl-axis c2-axis
5
8
1
3. Show that the procedure given in the text may be used to solve pairs of
simultaneous differential equations of the form
ic=ax+by+p
y=cx+dy+q
where a, b, c, d are constants and p, q are given functions of t (thereby extend-
ing the procedure to linear constant-coefficient equations which may not be
homogeneous).
dt 25
dc2 3 4
C2-
dt 25 c' 25
5. The procedure in the text shows how to solve the pair of linear constant
coefficient differential equations
is=ax+by
y=cx+dy
when b * 0. Explain how you would solve these equations in each of the
remaining cases (i) b = c = 0 and (ii) b = 0 but c * 0.
366 Models with linear interactions
18.3 A model for detecting diabetes
Glucose, an end product of cabohydrate digestion, is converted into
energy in the cells of the body. A hormone insulin, secreted by the
pancreas, facilitates the absorption of glucose by cells other than those
of the brain and nervous system.
A delicate balance is normally maintained between the amounts of
glucose and insulin in the bloodstream. If the insulin concentration is
too low, then too little glucose is absorbed from the bloodstream; the
unabsorbed glucose is then lost in the urine along with other nutrients.
If, on the other hand, the insulin concentration is too high, then too
much glucose is absorbed by cells other than those of the brain and
nervous system; lack of glucose available to the cells of the brain then
impairs its function. The end result in either case, whether too little or
too much insulin, can be coma and even death.
In the medical disorder Diabetes Mellitus, not enough insulin is secreted
by the pancreas. People suffering from this require supplements of insulin
in the form of regular injections, together with a modification of their
diet to regulate glucose input. In this section a simple model of the
interaction between glucose and insulin in the body is presented; we then
use this model to discuss a clinical test for the detection of mild forms
of diabetes.
The model
The main features that a model of the glucose-insulin regulation system
must take into account are as follows.
(a) A rise in the concentration of glucose in the bloodstream results in
the liver absorbing more of the glucose, which it converts and stores
as glycogen; a drop in the concentration of glucose reverses the
process.
(b) A rise in the concentration of insulin in the bloodstream enables the
glucose to pass more readily through the membranes of the cells in
skeletal muscle, resulting in greater absorption of glucose from the
bloodstream.
(c) A rise in the concentration of glucose in the bloodstream stimulates
the pancreas to produce insulin at a faster rate; a drop in the glucose
concentration lowers the rate of insulin production.
(d) Insulin, produced by the pancreas, is constantly being degraded by
the liver.
18.3 A model for detecting diabetes 367
The model omits details of the biochemistry involved and ignores the
effects of other hormones. It treats the bloodstream, moreover, as if
it were contained in a single compartment throughout which concen-
trations of glucose and insulin are uniform at each instant. In spite
of these simplifications, the model is nonetheless suitable as a basis for
understanding what is, in reality, a complicated situation.
Provided there has been no recent digestion, glucose and insulin con-
centrations will be in equilibrium. We are interested in how the system
responds to a change in that equilibrium. Thus we put
g = {excess glucose concentration),
h = {excess insulin concentration},
at time t. We use `h' because insulin is a hormone. Equilibrium occurs
for g = h = 0. Positive values of g or h corresponds to concentrations
greater than the equilibrium values and negative values to concentrations
less than the equilibrium values.
If either of g or h is given a non-zero value, then the body tries to
restore the equilibrium. We assume that the rates of change of these
quantities depend only on the values of g and h so that
dg
dt
- F i(g' h)'
dh
T = F2(gg h)9
Our model for the glucose and insulin concentrations is thus the solution
of the pair of differential equations (1), (2) with the initial conditions (3).
To solve the differential equations (1), (2) we apply the procedure from
the previous section and so obtain the second-order equation
g + (a + d)g + (ad + bc)g = 0 (4)
The initial conditions for (4), which can be found from (3) and (5), are
This in turn will depend on the values of a, b, c and d. The solutions for
g in the various cases are listed in Table 18.3.1
From the fact that a, b, c and d are all positive it can be shown that
the solutions of (7) for A are negative or have negative real part. Hence
the factors eAlt, eA2t, elt and a«t which occur in the above table must all
decay exponentially with time. Thus our model predicts that the glucose
concentration will approach its original undisturbed value with sufficient
lapse of time. This is just as we would expect.
In these solutions g is a linear function of go. This is a consequence of
our assumption that the differential equations are linear. It can be shown
that these solutions are good approximations to those of any smooth
non-linear model of the problem, if go is sufficiently small.
Experimental results
The model can be used to make numerical predictions about the con-
centration of glucose at various times once the constants occurring in
370 Models with linear interactions
the differential equtions (1) (2) are known. In order to determine these
constants, Bolie (1961) used three different methods, based on data from
previous experiments with dogs, which he extrapolated to humans. Mea-
suring glucose in grams, insulin in `units' and time in hours, he obtained
the following averaged values for normal individuals :
measured in units corresponding to grams for mass and hours for time.
Substituting these values into the characteristic equation (7) and then
solving for 2 gives
Al =-1.36, 22=-2.34.
Thus the characteristic equation has two real roots. Hence we can
substitute the numerical values into the first row of Table 18.3.1 to get g
as a function of t and then use (5) to get h. This gives
g = go(-0.56eAl' + 1.56e22`)
,,t - A2t (8)
h = 0.202go(e a ).
g = A sin(cot)e` (10)
Burghes and Borrie (1981) switch the roles of w and wo and then state a
conclusion which is different from that given in Ackerman et al. (1964).
Exercises 18.3
1. The differential equations for the glucose and insulin concentrations are given
in the text as
dg = -ag - bh
dt
dh = cg - dh
dt
where a, b, c and d are constants. We showed in the text that d > 0. Show in a
similar way, by referring to the relevant principles from physiology, that each of
a, b and c is positive.
2. Derive, from the equations in Exercise 1, the following equations.
g+(a+d)g+(ad+bc)g = 0, h = -1(g+ag).
b
(Apply the solution procedure for simultaneous linear differential equations with
constant coefficients from Section 18.2.)
Show, furthermore, that the initial conditions
g=go, h=0 at t=0
are equivalent to
g = go, g = -ago at t = 0.
3. Given that a, b, c and d are all positive, show that the solutions of the
characteristic equation
A2 +(a+d)),+(ad + bc) = 0
are either negative or have negative real part.
18.4 Nutrient exchange in the placenta 373
Fig. 18.4.1. Placenta provides interface between the bloodstreams of mother and
fetus.
4. On the basis of Bolie's model for the glucose-tolerance test, we obtained the
following formula for the insulin concentration t hours after the glucose injection:
h = 0.202go (eA' ` - eA2 `)
where Al = -1.36 and 22 = -2.34. Choose any positive value for go and then
sketch the graph of h against t and find the time at which h reaches its maximum
value. (For any other choice of go, the graph can be obtained simply by scaling
in the vertical direction.)
5. In an insulin-tolerance test, an injection of insulin is given to an individual
after fasting and the level of glucose in the blood is measured at subsequent times.
Assume that that the concentrations of glucose and insulin in the bloodstream
satisfy the differential equations of Exercise 1, together with the initial conditions
g=0, h=ho at t=0.
Show that, on the basis of Bolie's estimates for the coefficients in the differential
equations,
g = 4.35ho(-eA' ` + eA2`),
h = ho(1.57e2'` - .57e22`).
Sketch the graph of h as a function of t.
Porous
membrane
Maternal blood vessel
Fig. 18.4.2. Two different types of blood flow in a placenta: (a) maternal pool,
(b) countercurrent flow.
The model
The model of the placenta we will describe is illustrated in Figure 18.4.3,
our notation being as follows. We use Q, and Q2 to denote the rates of
flow of the maternal and fetal blood respectively. We suppose that the
blood vessels of the mother and fetus stay in contact with the membrane
along a total length L.
We choose as coordinate the distance x of a typical point along the
blood vessels from the point where they first make contact with the
membrane.The concentration of nutrient in each blood vessel is then a
18.4 Nutrient exchange in the placenta 375
Q1
/000
X =O x=L
Q2 x-axis
nutrient concentration
Cl = in the maternal blood = ¢1(x),
at a distance x
(1)
nutrient concentration
c2 = { in the fetal blood 02(x).
at a distance x
Maternal
Q1 01(x) -. Q1 01 (X + Sx)
11111ATIVII
Pb (c1- c2) Sx
Fetal
-+-+ (x-axis)
x (x + Ox)
Fig. 18.4.4. (a) Planes through the blood vessels. (b) Nutrient flow in maternal
blood vessels.
= Qlo1(x)St
and similarly
mass leaving
throu gh p lane Q l ¢ l ( x + S x) St . 4)
atx+Sx
The principle which enables us to estimate the amount of nutrient
transported across the membrane, known as Fick's law, states that if the
concentrations on either side of the membrane were homogeneous then
rate of area of x difference between
transport through . p X {membrane} concentrations
the membrane on either side
and hence
mass of
nutrient leaving P x b 8x x (cl- C2) x St. (5)
across membrane
Now substitute (3), (4) and (5) into (2) and then divide by bt and
rearrange to get
Q1(41(x + 8x) - q51(x)) + Pb(cl - c2) Sx 0,
the error involved in the approximation being small compared with 8x.
Hence, dividing by 6x and then letting 6x approach 0 gives
dcl
dx
= _ai(i
c - C2)
2) ()
(6)
where al = Pb/Q1.
A similar derivation for the fetal bloodstream gives
dc2_
dx -a2(cl - c2) (7)
where a2 = Pb/Q2.
From this we can deduce that, if al *- a2, then the differential equations
(6), (7) have the solutions
al
C1 = 01(x) = 41(0) - (q51(0) - 42(0)) (e(_12)x - 11 ,
a2 - al
a2
(9)
C2 = 02(x) = 42(0) - (41(0) - 42(0)) (e(_12)x -
a2 - ai
Comparisons
The placenta modelled above is called a countercurrent type of pla-
centa because the two bloodstreams flow in opposite directions. Middle-
man (1972) gives further details, and he models other types of placenta,
obtaining solutions analogous to (9) for the concentrations of the nutri-
ents. On the basis of such solutions, he is able to make some comparisons
378 Models with linear interactions
between the efficiency of the various types of placenta in exchanging nu-
trients.
Models analogous to that of the countercurrent nutrient exchange
system also occur in other applications. These include simple models
of an artificial kidney machine,which is described in Burghes and Borrie
(1981), and oxygen exchange in the swim bladders of deep sea fish,
described in Rodin and Jacques (1989).
Exercises 18.4
1. In the text, the differential equation for the concentration cl of nutrient in the
maternal bloodstream at a distance x along the placental membrane was shown
to be
dcl_
--al (Cl - c2)
dx
for a suitable constant al.
By arguing in a similar way, derive the corresponding differential equation
dc2_
dx - -a2(c1 - c2)
for the concentration of nutrient in the fetal bloodstream.
Under which physical condition does al = a2 ?
2. Why can the procedure given in Section 18.2 be used to solve the differential
equation in Exercise 1 ? Use this procedure to solve these equations in the case
al * a2. Check your answers against those given in the text.
3. Repeat Exercise 2 but with al = a2. Interpret your answers.
4. Use the solutions (9) in the text to establish each of the following statements.
(a) If the concentration of nutrient entering via the maternal artery is
equal to that leaving via the fetal vein, then the concentrations must be
constant along the entire length of the placental membrane.
(b) If the concentration of nutrient entering via the maternal artery exceeds
that leaving via the fetal vein, then at each point along the membrane
the concentration in the maternal bloodstream exceeds that in the fetal
bloodstream.
5. Modify the model given in the text if, instead of flowing in opposite directions,
the maternal and fetal bloodstreams flow in the same direction. (This is called
concurrent exchange.)
19
Non-linear coupled models
379
380 Non-linear coupled models
Year Sharks
(% total catch)
1914 11.9
1915 21.4
1916 22.1
1917 21.2
1918 36.4
1919 27.3
1920 16.0
1921 15.9
1922 14.8
1923 10.7
of sharks to total catch. It reveals that this ratio, and thus the relative
number of sharks, increased substantially during a time of reduced fishing
(1915 to 1919 corresponding to the First World War).
How can these data be understood? Volterra set up the following model
to try to explain why the decrease in fishing increased the percentage
catch of sharks.
The model
Following Volterra, our objective is to formulate some differential equa-
tions describing the shark and fish populations, which will be done using
the small time-interval method introduced in Chapter 13. The following
influencing factors will be taken into account.
(i) Natural births and deaths of the sharks and fish in isolation from
each other.
(ii) Decline of the fish population due to the fish being the prey of
the sharks.
(rii) Increase in the shark population due to the presence of more fish.
(iv) Fishing of both sharks and fish.
For the populations, we introduce the notation
number of
x fish at time t
_J number of
y - l sharks at time t
19.1 Predator-prey interactions 381
and
extra sharks
sharks shark sharks
by = born in
isolation
- deaths in +
isolation
surviving
fisho
_ caught by
fishermen
(2)
d
where each of the quantities on the right-hand side refers to the number
of sharks and fish which are born or die during the time interval bt.
We need to relate the various quantities in (1) and (2) to the shark
and fish populations. To do this we will make the following modelling
assumptions.
(a) The change in the shark and fish populations, in isolation, is pro-
portional to the present population of sharks and fish, respectively.
(This is the same assumption as for the linear model of population
growth made in Section 9.1.) The proportionality constant for the
shark population is negative, indicating that the shark population
would decrease if isolated from the fish population. This is a
consequence of the fish being the food supply of the sharks.
(b) The number of sharks and fish caught by fishermen is directly pro-
portional to the present population of the shark and fish populations
respectively. The proportionality constant is the same in both
cases, which means that the fishing methods do not discriminate
between sharks or fish.
(c) The number of fish eaten by sharks is directly proportional to the
product of the number of fish present and the number of sharks
present. This is equivalent to assuming that each shark eats a
constant fraction of the fish population.
(d) The additional number of sharks surviving is directly proportional
to the number offish eaten.
In the exercises you are asked to show that, when written as mathemat-
ical equations and substituted into equations (1) and (2), the assumptions
382 Non-linear coupled models
(a)-(d) imply the differential equations
dx _
(r - f )x - axy,
dt =
d (3)
dty = -(s + D y + fixy,
where r, s, f, a and fi are all positive constants. The constants r and -s
are the growth rates for the fish and shark populations if they existed
in isolation, the constant a is the rate at which fish are eaten by a
single shark and fl/ce gives the fraction of a shark surviving by eating
one fish. These equations are called the Lotka-Volterra equations after
Volterra, and Lotka another mathematician who formulated them
independently.
The Lotka-Volterra equations are coupled since the unknowns x and
y appear in both of the equations. Furthermore they are non-linear due
to the presence of the terms axy and fixy. Because of the non-linear
terms, it is not possible to solve these equations explicitly for general
initial conditions. We therefore seek ways of obtaining useful information
about the solutions. The steady-state solutions provide an informative
beginning.
Steady-state solutions
The steady-state solutions (i.e. time-independent constant solutions) are
the easiest particular solutions to find for an autonomous differential
equation. To find these solutions set
x=X and y=Y,
where X and Y are constants, and substitute into the Lotka-Volterra
equations (3) to give two non-linear simultaneous equations
and
(X,Y)=
rs+P f r-f/
Ot
19.1 Predator-prey interactions 383
Effect of fishing
Consider the circumstance that there is less fishing, as in the First World
War. In the Lotka-Volterra model this means that f is decreased but all
other parameters remain the same. The second constant solution above
then gives that the fish population actually decreases while the shark
population increases. Hence the ratio of sharks to fish increases, which
is consistent with the data of Table 19.1.1. The mechanism for this result
will be discussed in the next section.
Exercises 19.1
1. Consider the statements (a)-(d) of the text which relate the quantities
influencing the shark and fish populations to the populations themselves. Suppose
that the proportionality constants are r and -s in (a) for the fish and shark
populations respectively, f in (b), a in (c) and f in (d). Write the statements as
mathematical equations, keeping in mind that each quantity on the right-hand
side of equations (1) and (2) is also directly proportional to the time interval
8t. Substitute these equations in (1) and (2) and thus derive the Lotka-Volterra
equations.
2. Factorize and thus solve the two non-linear equations given in the text which
specify the steady-state solutions of the Lotka-Volterra equations.
3. (a) Suppose there are two different species of prey x and z, which are the food
of a single predator y. Derive differential equations for the changes in
the populations of the species by using modelling assumptions analogous
to those used to derive the Lotka-Volterra equations.
384 Non-linear coupled models
(b) Find all the steady-state solutions of the differential equations found in
part (a), and thus show that at least one species of prey becomes extinct
in this instance.
The method
There are three main steps in applying a phase-plane analysis, which we
will illustrate by the following example.
Example 1. Find, and sketch, the phase-plane trajectories for the coupled system
5c=-y and y=x. (2)
19.2 Phase-plane analysis 385
y-axis
K=3
x-axis
Fig. 19.2.1. Phase-plane trajectories for the coupled system (2) in Example I.
Solution.
STEP 1: Write down the chain rule and substitute the values of is and y from the
given equations. Hence obtain a single first-order differential equation in y and x
by using the chain rule and then substituting for the derivatives from the original
coupled differential equations.
The chain rule says that
dy dx dy
dx dt dt'
or equivalently,
dy
dxx Y.
where xo and yo are the initial values of x and y. Since xo and yo are constants,
we can write
xo+yo=K2
where K is a constant. The relationship between x and y defines a circle of radius
K, and varying K gives a family of circles as sketched in Figure 19.2.1.
STEP 3: Determine the directions of the trajectories by finding if x and y increase
or decrease with t at a few selected regions in the phase-plane.
386 Non-linear coupled models
From the original differential equations, for the region x > 0, y > 0, we see that
ac < 0, which means x decreases with t;
y > 0, which means y increases with t.
Thus the trajectories are anti-clockwise around the circles, as indicated in Fig-
ure 19.2.1.
where
K =rlnyo - ayo+slnxo - fxo
and K is a constant, which depends on the initial populations xo and yo.
The notation r = r - f and s = s + f has been used, and xo and yo are
the initial values of xo and yo.
19.2 Phase-plane analysis 387
K = -2
4r
2
Fish X
(b) 4 r
12
0 2 0 2 4
(c) Fish X (d) Fish X
iii II
y<o y>0
x<0 i x<O
F
Y- a
---------- T ----------
IV I
<o
.z>O i />o
x>0
x-axis
X= s/0
Interpretation
Consideration of Figure 19.2.3 allows aspects of the predator-prey system
of sharks and fish to be better understood. Suppose initially both the
shark and fish populations are above their steady-state values and thus
in region II of Figure 19.2.3. In this circumstance the shark population
increases at the expense of the fish population until the fish population
drops below its steady-state value into region III. Now there is not
enough food for the sharks so the shark population also decreases until
it drops below its steady-state value into region IV. The fish population
can now recover as there are fewer sharks. It eventually increases to
above its steady-state value into region I. Now there is sufficient food for
the shark population to recover, so it begins to increase until region II is
again entered and the cycle repeated.
19.3 Models of combat 389
Exercises 19.2
2ydy = -x.
dx
(b) Hence obtain the equations for the family of phase-plane trajectories.
(c) Sketch the family of phase-plane trajectories in the (x, y)-plane. Remem-
ber to indicate by arrows the direction in which the solution moves along
these curves.
The firing rates RX and Ry are assumed to be constant, while the prob-
abilities PX and Py are determined according to whether the target is
exposed or hidden.
If the target is exposed (the home army soldiers), it is reasonable to
assume that each single shot has a constant probability of hitting its
target, independent of the number of home soldiers. So Py is a constant
with respect to x and y.
If the target is hidden, however, then the probability of hitting a
19.3 Models of combat 391
soldier by a shot fired at random into a given area will depend on the
concentration of hidden soldiers in the area. If there are y enemy soldiers,
and if each enemy soldier has on average an area a exposed, then
Px = aAY
where A is the total area occupied by the enemy soldiers. Note that ay
gives the total area of soldiers available to be hit by random fire.
Substituting this formula into (1) and dividing by b t gives the coupled
non-linear equations
dx
= -ay and = -bxy (2)
dt
where a = RyPy and b = Rxa/A are positive constants.
Steady-state solutions
Setting x = X and y = Y, where X and Y are constants, gives
0 = -aY and 0 = -bX Y.
Hence the steady-state solution is Y = 0, X unspecified, which corre-
sponds to the enemy (guerilla) army being defeated. However, as will be
shown below via a phase-plane analysis, certain initial conditions lead to
X = 0 and thus victory to the enemy army.
Phase-plane analysis
The phase-plane technique of Section 19.2 allows us to eliminate t in the
equations (2) to obtain the differential equation
dy b
- X.
dx a
This equation is first-order separable so the methods of Section 11.1 can
be used to show that it has solution
b
Y= 2a
x2+K, (3)
where
K= - 2a (o)
y0
x 2
()4
is a constant depending on the initial conditions.
The parabolas defined by (3) for varying K are sketched in Fig-
ure 19.3.1. Only the region of the phase-plane x, y >_ 0 is sketched since
392 Non-linear coupled models
y-axis
x-axis
the number of soldiers cannot be negative. Note also from (2) that, if
x, y > 0, then ac, y < 0 so all the trajectories point towards the axes.
Suppose that a victory to the home army occurs when x > 0 and y = 0
(although in practice the defeated army would probably surrender before
this stage). From (3) and (4) this occurs when K < 0 and thus when
2
Yo < (xo )
2a
On the other hand, a victory to the hidden army occurs when K > 0 so
that
Yo > 2a(xoZ
As an application of these results, let us suppose it appeared that
the two armies were heading for mutual annihilation. From (3) and (4)
mutual annihilation will occur if
K=0
and thus
Yo _ b
5
(xo)2 2a
However if the enemy army were to double the number of soldiers it had
initially we see that the home army would only have to increase the size
of its army by a factor of ,/2- to match the enemy army.
Generally, however, the hidden army often has an advantage over the
exposed army. If the hidden army is spread over a very large area A then
the parameter b = RXoc/A is a small number compared with a = RyPy.
Thus (5) states that the two armies are evenly matched when x0 is large
and yo is relatively small.
19.3 Models of combat 393
Exercises 19.3
1. In this exercise a model of combat in which both armies are exposed to gun
fire is to be developed.
(a) Using equation (1) of the text, obtain the coupled differential equations
dx = = -qx
-pY and dy
dt dt
and specify the constants p and q in terms of the average firing rates
and probability of hitting the target for each army.
(b) Use a phase-plane analysis to determine an inequality relating p and q
such that the home army (x) wins. Do the same for a win to the enemy
army.
4. The equations obtained in Exercise 1 are linear and can thus be solved using
the method of Section 18.2.
(a) Hence find x and y as functions of t, given that initially x = xo and
Y = Yo
(b) By eliminating t from your solutions in (a) find a relation between y and
x. You should thus reproduce the result derived in Exercise 1(b).
5. Suppose that initially there are 80 000 soldiers of the `home' army and 100 000
soldiers of the enemy, and assume that both armies are exposed as in Exercise 1
above, with soldiers on each side being equally effective. (This means that p = q
in the differential equations.)
(a) If all soldiers take part in the battle, show that the enemy army wins the
battle.
(b) Show that the home army can win by first engaging half of its army
against the enemy army and then fighting the other half against the
enemy survivors of the first battle.
6. (a) Argue that a battle with long range artillery will satisfy the differential
equations
ac = -axy and y = -fixy
where a and fi are positive constants. Express these constants in terms
of other meaningful constants such as a firing rate, number of soldiers
per missile launcher, and firing rates.
(b) Sketch the phase-plane trajectories, and give a condition for one of the
sides to win.
19.4 Epidemics
A mathematical model of a measles epidemic was presented in Chapter 9.
The model was presented as a pair or coupled non-linear difference
equations. The reason for the applicability of difference equations was
the significant latent period between catching the disease and becoming
contagious. If this latent period is small (ideally zero) a model of an
epidemic involving coupled differential equations can be formulated. This
will be done in this section. A phase-plane analysis of the model is left
to the exercises.
19.4 Epidemics 395
The model
For the purposes of formulating our model for the spread of a disease,
the population will be divided into three groups: susceptibles (who are
not immune to the disease), infectives (who are capable of infecting
susceptibles) and removed (who have previously had the disease and may
not be reinfected because they are immune, have been quarantined or
have died from the disease). The symbols S, I and R will be used to
denote the number of susceptibles, infectives and removed, respectively,
in the population at time t.
The following modelling assumptions will be made.
(a) The disease is transmitted by close proximity or contact between an
infective and susceptible.
(b) A susceptible becomes an infective immediately after transmission.
(c) Infectives eventually become removed.
(d) The population of susceptibles is not altered by emigration, immi-
gration, births or deaths.
(e) Each infective infects a constant fraction # of the susceptible pop-
ulation per unit time.
(f) The number of infectives removed is proportional to the number of
infectives present.
As mentioned earlier, it is assumption (b) which makes a formulation
involving differential equations rather than difference equations rele-
vant. Diseases for which this assumption is applicable include diphtheria,
scarlet fever and herpes. Assumption (e) is the same as that used in
Section 9.4. It is valid provided that the number of infectives is small in
comparison to the number of susceptibles.
To set up the differential equations using a 8t argument, let SS, 8I and
SR denote the changes in the population of susceptibles, infectives and
removed during a small time interval 8t. By assumptions (a), (c) and (d)
aS - - ft number of susceptiblesl
infected in time 8t
By assumptions (a), (b) and (c)
number of susceptibles number of
al - infected in time St } - infectives removed
By assumptions (a), (c) and (d)
aR = f number of infectivesl
1 removed in time St
396 Non-linear coupled models
But from assumptions (e) and (f)
number of susceptiblesl =BSI St,
infected in time 8t f
number of infectivesl = yl St.
removed in time St
Substituting the last two equations into the previous three, dividing by
b t and letting b t -+ 0 gives the coupled differential equations
dS=_SI (la
dt )
dI
= flSI-
dt y1' (1b)
dR--1 (1c)
dt y '
where fi and y are positive constants of proportionality. Here fJ is known
as the infection rate which governs how fast the disease is spread from
one infective to one susceptible and y is the removal rate which governs
how fast infectives are removed (by dying, becoming immune or by being
quarantined).
Equations (la-c) give three differential equations in three unknowns
S, I and R. Notice, however, that the variable R does not occur in (1a)
or (1b) hence equation (1c) is not coupled to the system (la-b).
Note also, that adding the three equations (1a), (1b) and (1c) gives
S +-1 R) = 0
dt( + )
where No denotes the initial population. Equation (2) is evident from the
formulae for 5S, bl and bR which in turn follow from assumptions (b),
(c) and (d) used in the formulation of the model.
The equations (la) and (lb) can be analysed using the phase-plane
technique, which is to be done in the exercises. An important consequence
of the analysis is the following result :
If the initial number of susceptibles is greater than y/fl then the number of
infectives will increase and we say that an epidemic has occurred. If it is smaller
than y/fl then the number of infectives decreases and thus no epidemic occurs.
Thus knowing # and y can help decide vaccination strategies: one would
19.4 Epidemics 397
Exercises 19.4
1. Consider the equations (1a) and (1b) in the text, which describe the model for
an epidemic with removal.
(a) Eliminate t and solve the resulting differential equation to obtain the
formula
I=K -S+ylnS
P
where K is a constant depending on the initial values of I and K.
(b) From your answer to (a), find the maximum of I regarded as a function
of S and sketch some typical trajectories. Thus determine a condition
on So for which the number of infectives will keep decreasing with time.
Also give the condition on So for which the number of infectives will
initially increase with time. In this instance an epidemic is said to occur.
(c) Find a method to determine the number of susceptibles remaining when
the disease has run its course and the number of infectives is zero.
3. Generalize the model of the text to include the possibility of those recovering
from the disease becoming reinfected.
398 Non-linear coupled models
4. With y = 0 and thus zero rate of removal, equations (1a) and (1b) of the text
read
S = -PSI and 1 = PSI.
(a) Determine, and sketch, the phase-plane trajectories.
(b) Show that
1 =,6I(No-I)
where No = So + Io
(c) Solve this differential equation, and check that the solution is consistent
with the phase-plane trajectories.
(d) Give some examples of diseases which could be modelled by these
equations.
5. The following set of coupled non-linear differential equations has been used
to model tranmission of AIDS (May, Anderson and McLean, 1988). Here X
represents a susceptible population and Y represents those infected with the HIV
virus, and N = X + Y.
dX
= vN - (A + µ)X,
dt
dY
= AX - (v + )Y,
dt
where A is the probability of acquiring infection from any one partner.
(a) Discuss why it is plausible to write A = ficY IN, where fi is the probability
of acquiring infection from one infected partner and c is the rate at which
new partners are acquired.
(b) Which terms in the equation represent
(i) input of new susceptibles,
(ii) HIV carriers who get AIDS,
(iii) those who become infected with HIV?
(c) Show that the total population N satisfies the differential equation
dN
=vN -pN - vY.
dt
References
Chapter 1
Andrade, E.N. da C. (1979) Sir Isaac Newton. Greenwood Press, Westport
Connecticut.
Cohen, I.B. (1987) The Birth of a New Physics. Penguin Books, Middlesex.
Koestler, A. (1958) The Sleepwalkers. Penguin Books, Middlesex.
Westfall, R.S. (1971) Force in Newtonian Physics. Elsevier, New York.
Chapter 4
Halliday, D. and Resnick, R. (1974) Physics Parts I and II. Wiley, New York.
Chapter 7
Gardner, M. (1981) Mathematical Circus. Penguin, Middlesex.
Chapter 8
Archibald, G.C. and Lipsey, R.G. (1973) An Introduction to a Mathematical
Treatment of Economics. 3rd edition. Weidenfeldt and Nicolson, London.
Ayres, F. (1963) Theory and Problems of Mathematics of Finance. Schaum, New
York.
Gandolfo, G.C. (1971) Mathematical Methods and Models in Economic
Dynamics. North Holland, Amsterdam.
Goldberg, S. (1958) An Introduction to Difference Equations: with illustrated
examples from Economics, Psychology and Sociology. Wiley, New York.
Kenkal, J.L. (1974) Dynamic Linear Economic Models. Gordon and Breach,
London.
Lipsey, R.G., Langley, P.C. and Mahoney, D.M. (1981) Positive Economics for
Australian Students. Weidenfeldt and Nicolson, London.
Pfouts, R.W. (1972) Elementary Economics: a Mathematical Approach. Wiley,
New York.
Chapter 9
Anderson, R. and May, R. (1982) `The logic of vaccination: New Scientist,
November.
Devaney, R.L. (1986) An Introduction to Chaotic Dynamical Systems.
Benjamin-Cummings, Menlo-Park, California.
399
400 References
Chapter 10
Edelstein-Keshet, L. (1988) Mathematical Models in Biology. Random House,
New York.
Haldane, J.B.S. (1924). `A mathematical theory of artificial and natural selection
- I.' Transactions of the Cambridge Philosophical Society, 23, 19-41.
(Reprinted in the Bulletin of Mathematical Biology, 52, 209-240. 1990.)
Hartl, D.L. (1980). Principles of Population Genetics. Sinauer Assoociates,
Sunderland, Massachusetts.
Hexter,W. and Yost, H.T. (1976) The Science of Genetics. Prentice-Hall,
Englewood Cliffs, New Jersey.
Maynard-Smith, J. (1968) Mathematical Ideas in Biology. Cambridge University
Press.
Sandfur, J.T. (1968) `Difference Equations in Genetics: The UMAP Journal 10,
257-274.
Chapter 11
Braun, M. (1975) Differential Equations and their Applications: an Introduction
to Applied Mathematics. Springer, New York.
Emlen, J.M. (1984) Population Biology: The Coevolution of Population Dynamics
and Behaviour. Macmillan, New York.
Emmel, T.C. (1976) Population Biology. Harper and Row, New York.
Giancoli, D.C. (1985) Physics. 2nd edition. Prentice Hall, Englewood Cliffs, New
Jersey.
Hutchinson, G.E. (1971) An Introduction to Population Ecology. Yale University
Press, Yale.
Keyfitz, N. (1977) Introduction to the Mathematics of Populations.
Addison-Wesley, Reading, Massachusetts.
Kormondy, E.J. (1976) Concepts of Ecology. Prentice-Hall, Englewood Cliffs,
New Jersey.
Marion, J.B. (1976) Physics in the Modern World. Academic Press, Reading,
Massachusetts.
Rubinow, S.I. (1975) An Introduction to Mathematical Biology. Wiley, New York.
References 401
Chapter 12
Holman, J.P. (1981) Heat Transfer. McGraw-Hill, New York.
Modelling Heat. Open University Module, Open University Press, Milton
Keynes.
Chapter 13
Rainey, R.H. (1967) `Natural Displacement of Pollution from the Great Lakes.'
Science, 155, 1244.
Modelling Heat. Open University Module, Open University Press, Milton
Keynes.
Chapter 11
Melissinos, A.C. (1966) Experiments in Modern Physics. Academic Press,
Reading, Massachusetts.
Streeter, V.L. (1966) Fluid Mechanics. McGraw-Hill, New York.
Chapter 15
Braun, M. (1975) Differential Equations and their Applications: an Introduction
to Applied Mathematics. Springer, New York.
Chapter 16
Cohen, I.B. (1987) The Birth of a New Physics. Penguin, Middlesex.
De Mestre, N. (1990) The Mathematics of Projectiles in Sport. Cambridge
University Press.
Chapter 17
Toeplitz, O. (1963) Calculus: A Genetic Approach. University of Chicago Press,
Chicago.
Chapter 18
Ackerman, E., Rosevar, J.W. and Mc Guckin, W.F. (1964) Phys. Med. Biol. 9,
203.
Bolie, V.W. (1960) `Coefficients of normal glucose regulation.' Journal of Applied
Physiology, 16, 783.
Braun, M. (1975) Differential Equations and their Applications: an Introduction
to Applied Mathematics. Springer, New York.
Burghes, D.N. and Borrie (1981) Modelling with Differential Equations. Ellis
Horward Limited, Chichester.
Edeistein-Keshet, L. (1988) Mathematical Models in Biology. Random House,
New York.
Middleman (1972) Transport Phenomena in the Cardiovascular System.
Wiley-Interscience, New York.
Rodin, E,.Y. and Jacques, S. (1989) `Countercurrent oxygen exchange in the
swim bladders of deep sea fish : a mathematical model.' Mathematical and
Computer Modelling, 12, 389.
402 References
Chapter 19
Bailey, T.J. (1975) The Mathematical Theory of Infectious Diseases and its
Applications. 2nd edition. Charles Griffin and Company Limited.
Braun, M. (1975) Differential Equations and their Applications: an Introduction
to Applied Mathematics. Springer, New York.
Edelstein-Keshet, L. (1988) Mathematical Models in Biology. Random House,
New York.
Giordano, F.R. & Wier, M.D. (1985) A First Course in Mathematical Modeling.
Belmont California.
Lanchester : chapter in Newman (1956) The World of Mathematics - Volume
Two. Simon and Schuster, New York.
Index
403
404 Index
CAMBRIDGE
UNIVERSITY PRESS
ISBN 9780521440691